You are on page 1of 87

INTRODUCTION TO

PROCESS DYNAMICS &


CONTROL
INTRODUCTION TO P&ID
*PPT and YT Discussion*

Process
- The conversion of feed materials to products using chemical and physical operations. In
practice, the term process tends to be used for both the processing equipment. (Seborg,
2004)
- The types of processes: Continuous, Batch, Semi-batch

Continuous - continuous stirred tank reactor, thermal cracking furnace, multi component
distillation column

Batch/Semi Batch - batch digester in a pump mill, plasma etcher in a semiconductor processing,
batch or semi batch reactor

Types of Process Variables

● Controlled (Measured) Variable


- The process variables that are controlled. The desired value of a controllable
value is referred to as its set point.

● Manipulated Variable
- The process variables that can be adjusted in order to keep the controllable
variables at or near their set points. Typically, the manipulated variables are flow
rates.

● Disturbance Variable
- Process variables that affect the controlled variables but cannot be manipulated.
Disturbances generally are related to changes in the operating environment of the
process.

Shell and Tube Heat Exchanger

A process fluid on the shell side is cooled by cooling water on the tube side. Typically,
the exit temperature of the process fluid is controlled by manipulating the cooling water flow
rate. Variations in the inlet temperatures and the process fluid flow rate affects the heat
exchanger operation. Consequently, these variables and disturbance variables.
● Controlled Variable
- Exit temperature
● Manipulated Variable
- Cooling water flow rate
● Disturbance Variable
- Inlet temperature
- Process fluid flow rate

A Blending Process - CSTR Blending System

The control objective is to blend the two inlet streams to produce an outlet stream that has
the desired composition. Stream 1 is a mixture of two chemical species, A and B. We assume
that its mass flow rate 𝑤1 is constant, but the mass fraction of A, 𝑥1 , varies the time. Stream 2
consists of pure A and thus 𝑥2 = 1 the mass flow rate of Stream 2, 𝑤2 , can be manipulated using
a control valve. The mass fraction of A in the exit stream is donated by x and the desired value
(set point) by 𝑥𝑠𝑝 .

Controlled variable x
Manipulated Variable w2
Disturbance Variable x1

Design Question: If the nominal value of 𝑥1 is 𝑥̅1 , what nominal flow rate 𝑤
̅ 2 is required to
produce the desired outlet concentration 𝑥𝑠𝑝 ?

Overall Balance; Steady-state


𝑤
̅1 + 𝑤
̅2 = 𝑤̅
𝑤
̅1 + 𝑤
̅2 + 𝑤̅ =0

Complete A Balance; Steady-state


𝑤
̅1 𝑥̅1 + 𝑤
̅ 2 𝑥̅2 − 𝑤𝑥
̅̅̅̅ = 0
𝑥
𝑠𝑝 1−𝑥̅
𝑤 ̅1 ( 1−𝑥 )
̅2 = 𝑤
𝑠𝑝

Process Dynamics

● The term refers to unsteady-state (transient) process behavior.


● Transient operation occurs during important situations such as startups and shutdowns,
unusual process disturbances, and planned transitions from one product to grade to
another. (Seborg, 2004)
A Blending Process - CSTR Blending System

Process Control

● The primary objective is to maintain a process at the desired operating conditions safely
and efficiently, while satisfying environmental and product quality requirements.
(Seborg, 2004)
● This refers to the methods that are used to control processes variables when
manufacturing a product.

Process Variables

● A condition of the process fluid (a liquid or gas) that can change the manufacturing
process in some way.
- Pressure
- Flow
- Level
- Temperature
- Density*
- pH*
- Mass*
- Conductivity*

Manufacturers control the production process for three reasons:


- Reduce variability
- Increase efficiency
- Ensure safety
Reduce Variability
● Process control can reduce variability in the end-product, which ensures a consistently
high-quality product.
● Reducing variability can also save money by reducing for product padding to meet
required product specifications.

Increase Efficiency
● Some processes need to be maintained at a specific point to maximize efficiency.
● Manufacturers save money by minimizing the resources required to produce the end-
product.

Ensure Safety
● A run-away process, such as an out-of-control nuclear or chemical reaction, may result if
manufacturers do not maintain precise control of all process variables. The consequences
of a run-away process can be catastrophic.
● Precise process control may also be required to ensure safety.
Process Control Terms

Setpoint
- A value for a process variable that is desired to be maintained.

Error
- The difference between the measured variable and the setpoint can be either positive or
negative.
- The objective of any control scheme is to minimize or eliminate error.

Magnitude
- The deviation between the values of the setpoint and the process variables.

Duration
- The length of time that an error condition has existed.

Rate of Change
- The length of time that an error condition has existed.
Offset
- A sustained deviation of the process variable from the setpoint.

Load Disturbance
- An undesired change in one of the factors that can affect the process variable.

Control Algorithm
- The mathematical expression of a control function.
- This can be used to calculate requirements of much more complex control loops.
Manual Control
- Control operations that involve human action to make an adjustment.

Automatic Control
- Control operations in which no human intervention is required
- Automatic valve actuator

Closed Control Loops


- Loop that exists where a process variable is measured, compared to setpoint, and action is
taken to correct any deviation from setpoint.
Open Control Loops
- Loop that exists where the process variable is not compared, and action is not taken not in
response to feedback on the condition of the process variables that is instead taken
without regard to process variable conditions.

Process Control Loop


PDC Book Chapter 1
CHAPTER 1- INTRODUCTION TO PROCESS CONTROL

1.1 REPRESENTATIVE PROCESS CONTROL PROBLEMS


The foundation of process control is process understanding.

Process: The conversion of feed materials to products using chemical and physical
operations. In practice, the term process tends to be used for both the processing operation and
the processing equipment.

Note that this definition applies to three types of common processes: continuous, batch,
and semi-batch. Next, we consider representative processes and briefly summarize key
control issues.

1.1.1 Continuous Processes


Four continuous processes are shown schematically in Fig. 1.1:

(a) Tubular heat exchanger.


A process fluid on the tube side is cooled by cooling water on the shell side.
Typically, the exit temperature of the process fluid is controlled by manipulating
the cooling water flow rate. Variations in the inlet temperatures and the process
fluid flow rate affect the heat exchanger operation. Consequently, these variables
are considered to be disturbance variables.
(b) Continuous stirred-tank reactor (CSTR).
If the reaction is highly exothermic, it is necessary to control the reactor
temperature by manipulating the flow rate of coolant in a jacket or cooling coil. The
feed conditions (composition, flow rate, and temperature) can be manipulated
variables or disturbance variables.

(c) Thermal cracking furnace.


Crude oil is broken down ("cracked") into a number of lighter petroleum
fractions by the heat transferred from a burning fuel/air mixture. The furnace
temperature and amount of excess air in the flue gas can be controlled by
manipulating the fuel flow rate and the fuel/air ratio. The crude oil composition and
the heating quality of the fuel are common disturbance variables.
(d) Multicomponent distillation column.
Many different control objectives can be formulated for distillation columns.
For example, the distillate composition can be controlled by adjusting the reflux
flow rate or the distillate flow rate. If the composition cannot be measured on -line,
a tray temperature near the top of the column can be controlled instead. If the feed
stream is supplied by an upstream process, the feed conditions will be dist urbance
variables.
For each of these four examples, the process control problem has been
characterized by identifying three important types of process variables.
1. Controlled variables (CVs): The process variables that are controlled. The
desired value of a con-trolled variable is referred to as its set point
2. Manipulated variables (MVs):
− can be adjusted in order to keep the con-trolled variables at or near
their set points.
− flow rates.
3. Disturbance variables (DVs): Process variables that affect the controlled
variables but cannot be manipulated.
− related to changes in the operating environment of the process, for
example, its feed conditions or ambient temperature. Some
disturbance variables can be measured on-line, but many cannot such
as the crude oil composition for Process (c), a thermal cracking
furnace.

The specification of CVs, MVs, and DVs is a critical step in developing a control
system. The selections should be based on process knowledge, experience, and control
objectives.

1.1.2 Batch and Semi-Batch Processes


Batch and semi-batch processes are used in many process industries, including
microelectronics, pharmaceuticals, specialty chemicals, and fermentation. Batch and
semi-batch processes provide needed flexibility for multiproduct plants, especially
when products change frequently and production quantities are small. Figure 1.2 shows
four representative batch and semi-batch processes:

(e) Batch or semi-batch reactor. An initial charge of reactants is brought up to reaction


conditions, and the reactions are allowed to proceed for a specified period of time
or until a specified con-version is obtained. Batch and semi-batch reactors are used
routinely in specialty chemical plants, polymerization plants (where a reaction
byproduct typically is removed during the reaction), and in pharmaceutical and
other bioprocessing facilities (where a feed stream, e.g., glucose, is fed into the
reactor during a portion of the cycle to feed a living organism, such as a yeast or
protein). Typically, the reactor temperature is controlled by manipulating a coolant
flow rate. The end-point (final) concentration of the batch can be controlled by
adjusting the de-sired temperature, the flow of reactants (for semi-batch operation),
or the cycle time.
(f) Batch digester in a pulp mill. Both continuous and semi-batch digesters are used in
paper manufacturing to break down wood chips in order to extract the cellulosic
fibers. The end point of the chemical reaction is indicated by the kappa number, a
measure of lignin content. It is con-trolled to a desired value by adjusting the
digester temperature, pressure, and/or cycle time.
(g) Plasma etcher in a semiconductor processing. A single wafer containing hundreds
of printed circuits is subjected to a mixture of etching gases under conditions
suitable to establish and maintain a plasma (a high voltage applied at high
temperature and extremely low pressure). The unwanted material on a layer of a
microelectronics circuit is selectively removed by chemical reactions. The
temperature, pressure, and flow rates of etching gases to the reactor are con-trolled
by adjusting electrical heaters and control valves. (h) Kidney dialysis unit. This
medical equipment is used to remove waste products from the blood of human
patients whose own kidneys are failing or have failed. The blood flow rate is
maintained by a pump, and "ambient conditions," such as temperature in the unit,
are controlled by adjusting a flow rate. The dialysis is continued long enough to
reduce waste concentrations to acceptable levels.

Next, we consider an illustrative example in more detail.

1.2 ILLUSTRATIVE EXAMPLE-A BLENDING PROCESS


A simple blending process is used to introduce some important issues in control system
design. Blending operations are commonly used in many industries to ensure that final
products meet customer specifications.
A continuous, stirred-tank blending system is shown in Fig. 1.3.

The control objective is to blend the two inlet streams to produce an outlet stream that
has the desired composition. Stream 1 is a mixture of two chemical species, A and B. We
assume that its mass flow rate w 1 is constant, but the mass fraction of A, x 1, varies with
time. Stream 2 consists of pure A and thus x 2 = 1. The mass flow rate of Stream 2, w 2, can
be manipulated using a control valve. The mass fraction of A in the exit stream is denoted
by x and the desired value (set point) by xsp. Thus for this control problem, the controlled
variable is x, the manipulated variable is w2, and the disturbance variable is x1.
Next we consider two questions.
Design Question. If the nominal value of x1 is ̅̅̅,
𝑥1 what nominal flow rate ̅̅̅̅
𝑤2 is required
to produce the desired outlet concentration, xsp?
To answer this question, we consider the steady-state material balances: Overall
balance:

0=𝑤
̅̅̅1̅ + ̅̅̅̅
𝑤2 − w
̅

Component A balance:
0=𝑤
̅̅̅1̅ ̅̅̅
𝑥1 + ̅̅̅̅
𝑤2 ̅̅̅
𝑥2 − w
̅ x̅ (1.2)
The overbar over a symbol denotes its nominal steady-state value, for example, the value used
in the process design. According to the process description, x2 = 1 and x̅ = xsp. Solving Eq.
1-1 for w
̅ , substituting these values into Eq. 1-2, and rearranging gives:
̅̅̅1̅
𝑥𝑠𝑝 −𝑥
𝑤2 = 𝑤
̅̅̅̅ ̅̅̅1̅ (1.2)
1− 𝑥𝑠𝑝

Equation 1-3 is the design equation for the blending system. If our assumptions are correct and
if x1 = x̅ then this value of w2 will produce the desired result, x = xsp. But what happens if
conditions change?
Control Question. Suppose that inlet concentration xl varies with time. How can we
ensure that the outlet composition x remains at or near its desired value, xsp?
As a specific example, assume that x 1 increases to a constant value that is larger than
its nominal value, 𝑥
̅̅̅.
1 It is clear that the outlet composition will also Increase due to the increase
in inlet composition.
Consequently, at this new steady state, x > xsp.
Next we consider several strategies for reducing the effects of x 1 disturbances on x.

Method 1. Measure x and adjust w 2. It is reasonable to measure controlled variable x


and then adjust w2 accordingly. For example, if x is too high, w2 should be reduced; if x is too
low, w2 should be increased. This control strategy could be implemented by a person (manual
control). However, it would normally be more convenient and economical to automate this
simple task (automatic control).
Method 1 can be implemented as a simple control algorithm (or control law),

𝑤2 (𝑡) = ̅̅̅̅
𝑤2 + 𝐾𝑐 [𝑥𝑠𝑝 − 𝑥(𝑡)] (1.4)

where Kc, is a constant called the controller gain. The symbols, w 2(t) and x(t), indicate
that w2 and x change with time. Equation 1-4 is an example of proportional control because
the change in the flow rate, w2(t) —w2, is proportional to the deviation from the set point, x(t).
Consequently, a large deviation from set point produces a large corrective action, while a small
deviation results in a small corrective action. Note that we require K c, to be positive because
w2 must increase when x decreases, and vice versa. However, in other control applications
negative values of Kc are appropriate as discussed in Chapter 8.
A schematic diagram of Method 1 is shown in Fig: 1.4.

The outlet concentration is measured and transmitted to the controller as an electrical


signal. (Electrical signals are shown as dashed lines in Fig. 1.4.) The controller executes the
control law and sends the calculated value of ur2 to the control valve as an electrical signal .
The control valve opens or closes accordingly. in Chapters 8 and 9 we consider process
instrumentation and control hardware in more detail.
Method 2. Measure x1, adjust w2. As an alternative to Method 1, we could measure disturbance
variable x1 and adjust w2 accordingly. Thus, if x1 > x̅, we would decrease w 2 so that w2 < ̅̅̅̅.
𝑤2
If x1 < x̅, we would increase w 2. A control law based on Method 2 can be derived from Eq. 1-
3 by replacing x̅ with x1(t) and ̅̅̅̅
𝑤2 with w2(t).

𝑥𝑠𝑝 −𝑥1 (𝑡)


𝑤2 (𝑡) = 𝑤
̅̅̅̅1 (1-5)
1− 𝑥𝑠𝑝

The schematic diagram for Method 2 is shown in Fig. 1.5. Because Eq. 1-3 is valid only for
steady-state conditions, it is not clear just how effective Method 2 will be during the transient
conditions that occur after an x1 disturbance.

Method 3. Measure x1 and x, adjust w2. This approach is a combination of Methods 1 and 2.
Method 4. Use a larger tank. If a larger tank is used, fluctuations in x 1 will tend to be damped
out as a result of the larger volume of liquid. However, increasing tank size is an expensive
solution due to the increased capital cost.

1.3 CLASSIFICATION OF PROCESS CONTROL STRATEGIES


Next, we will classify the four blending control strategies of the previous section and
discuss their relative advantages and disadvantages. Method 1 is an example of a feedback
control strategy. The distinguishing feature of feedback control is that the controlled
variable is measured and the measurement is used to adjust the manipulated variable. For.
feedback control, the disturbance variable is not - measured.
It is important to make a distinction between negative feedback and positive feedback.
In the engineering literature, negative feedback refers to the desirable situation where the
corrective action taken by the controller forces the controlled variable toward the set point.
On the other hand, when positive feedback occurs, the controller makes things worse by
forcing the controlled variable farther away from the set point. For example, in the blending
control problem, positive feedback takes place if KC, < 0 because w2 will increase when x
increases.' Clearly, it is of paramount importance to ensure that a feedback control system
incorporate negative feedback rather than positive feedback.

Table 1.1 concentration control strategies for the blending system


METHOD MEASURED MANIPULATED CATEGORY
VARIABLE VARIABLE
1 X W2 FB
2 X1 W2 FF
3 X1 and x W2 FF/FB
4 --- --- Design change
FB- feedback control; FF- feedforward control; FF/FB-feedforward and feedback control

An important advantage of feedback control is that corrective action occurs regardless of


the source of the disturbance. For example, in the blending process, the feedback control law in
(1-4) can accommodate disturbances in w1, as well as x1. Its ability to handle disturbances of
unknown origin is a major reason why feedback control is the dominant process control strategy.
Another important advantage is that feedback control reduces the sensitivity of the controlled
variable to unmeasured disturbances and process changes. However, feedback control does have
a fundamental limitation: no corrective action is taken until after the disturbance has upset the
process, that is, until after the controlled variable deviates from the set point. This shortcoming is
evident from the control law of (1-4).
Method 2 is an example of a feedforward control strategy. The distinguishing feature of
feedforward control is that the disturbance variable is measured, but the controlled variable is not.
The important advantage of feedforward control is that corrective action is taken before the
controlled variable deviates from the set point. Ideally, the corrective action will cancel the effects
of the disturbance so that the controlled variable is not affected by the disturbance. Although ideal
cancellation is generally not possible, feedforward control can significantly reduce the effects of
measured disturbances, as discussed in Chapter 15.
Feedforward control has three significant disadvantages: (i) the disturbance variable must
be measured (or accurately estimated), (ii) no corrective action is taken for unmeasured
disturbances, and (iii) a process model is required. For example, the feedforward control strategy
for the blending system (Method 2) does not take any corrective action for unmeasured w1
disturbances. In principle, we could deal with this situation by measuring both x1 and w1 and then
adjusting w2 accordingly. However, in industrial applications it is generally uneconomical to
attempt to measure all potential disturbances. A more practical approach is to use a combined
feedforward-feedback control system, where feedback control provides corrective action for
unmeasured disturbances, while feedforward control reacts to eliminate measured disturbances
before the controlled variable is upset. Consequently, in industrial applications feedfonvard control
is normally used in combination with feedback control. This approach is illustrated by Method 3,
a combined feedforward-feedback control strategy because both x and x1 are measured.
Finally, Method 4 consists of a process design change and thus is not really a control
strategy. The four strategies for the stirred-tank blending system are summarized in Table 1.1.

1.4 A MORE COMPLICATED EXAMPLE—A DISTILLATION COLUMN The blending


control system in the previous section is quite simple because there is only one con-trolled variable
and one manipulated variable. For most practical applications, there are multiple con-trolled
variables and multiple manipulated variables. As a representative example, we consider the
distillation column in Fig. 1.6 that has five controlled variables and five manipulated variables.
The controlled variables are product compositions, xD and xB, column pressure, P1, and the liquid
levels in the reflux drum and column base, hD and hB. The five manipulated variables are product
flow rates, D and B, reflux flow, R, and the heat duties for the condenser and reboiler, QD and QB.
The heat duties are adjusted via the control valves on the coolant and heating medium lines. The
feed stream is assumed to come from an upstream unit. Thus, the feed flow rate cannot be
manipulated, but it can be measured and used for feedforward control.
A conventional multiloop control strategy for this distillation column would consist of five
feedback control loops. Each control loop uses a single manipulated variable to control a single
controlled variable. But how should the controlled and manipulated variables be paired? The total
number of different multiloop control configurations that could be considered is 5! or 120. Many
of these control configurations are impractical or unworkable such as any configuration that
attempts to control the base level hB by manipulating distillate flow D or condenser heat duty QD.
However, even after the in-feasible control configurations are eliminated, there are still many
reasonable configurations left. Thus, there is a need for systematic techniques that can identify the
most promising configurations. Fortunately, such tools are available and are discussed in Chapter
18.
For control applications, where conventional multiloop control systems are not
satisfactory, an alternative approach, multivariable control, can be advantageous. In multivariable
control, each manipulated variable is adjusted based on the measurements of all the controlled
variables rather than only a single controlled variable, as in multiloop control. The adjustments are
based on a dynamic model of the process that indicates how the manipulated variables affect the
controlled variables. Consequently, the performance of multivariable control, or any model-based
control technique, will depend heavily on the accuracy of the process model. A specific type of
multivariable control, model predictive control, has had a major impact on industrial practice, as
discussed in Chapter 20.

1.4 THE HIERARCHY OF PROCESS CONTROL ACTIVITIES


As mentioned earlier, the chief objective of process control is to maintain a process at the
desired operating conditions, safely and efficiently, while satisfying environmental and
product quality requirements. So far, we have emphasized one process control activity, keeping
controlled variables at specified set points. But there are other important activities that we will
now briefly describe.
In Fig. 1.7 the process control activities are organized in the form of a hierarchy with
required functions at the lower levels and desirable, but optional, functions at the higher levels.
The time scale for each activity is shown on the left side of Fig. 1.7. Note that the frequency
of execution is much lower for the higher-level functions.
1. Measurement and Actuation (Level T)
Measurement devices (sensors and transmitters) and actuation equipment (for example, control
valves) are used to measure process variables and implement the calculated control actions. These
devices are interfaced to the control system, usually digital control equipment such as a digital
computer. Clearly, the measurement and actuation functions are an indispensable part of any
control system.

2. Safety and Environmental/Equipment Protection (Level 2)


The Level 2 functions play a critical role by ensuring that the process is operating safely and
satisfies environmental regulations. As discussed in Chapter 10, process safety relies on the
principle of multiple protection layers that involve groupings of equipment and human actions.
One layer includes process control functions, such as alarm management during abnormal
situations, and safety instrumented systems for emergency shutdowns. The safety equipment
(including sensors and control valves) operates independently of the regular instrumentation used
for regulatory control in Level 3a. Sensor validation techniques can be employed to confirm that
the sensors are functioning properly.

3. Regulatory Control (Level 3a)


As mentioned earlier, successful operation of a process requires that key process variables such as
flow rates, temperatures, pressures, and compositions be operated at, or close to, their set points.
This Level 3a activity, regulatory control, is achieved by applying standard feedback and
feedforward control techniques (Chapters 11-15). If the standard control techniques are not
satisfactory, a variety of advanced control techniques are available (Chapters 16-18). In recent
years, there has been increased interest in monitoring control system performance (Chapter 21).

4. Multivariable and Constraint Control (Level 3b)


Many difficult process control problems have two distinguishing characteristics: (i) significant
interactions occur among key process variables, and (ii) inequality constraints exist for
manipulated and controlled variables. The inequality constraints include upper and lower limits.
For example, each manipulated flow rate has an upper limit determined by the pump and control
valve characteristics. The lower limit may be zero or a small positive value based on safety
considerations. Limits on controlled variables reflect equipment constraints (for example,
metallurgical limits) and the operating objectives for the process. For example, a reactor
temperature may have an upper limit to avoid undesired side re-actions or catalyst degradation,
and a lower limit to ensure that the reaction(s) proceed.
The ability to operate a process close to a limiting constraint is an important objective for
advanced process control. For many industrial processes, the optimum operating condition occurs
at a constraint limit, for example, the maximum allowed impurity level in a product stream. For
these situations, the set point should not be the constraint value because a process disturbance
could force the controlled van-able beyond the limit. Thus, the set point should be set
conservatively, based on the ability of the control system to reduce the effects of disturbances.
This situation is illustrated in Fig. 1.8. For (a), the variability of the controlled variable is quite
high, and consequently, the set point must be specified well below the limit. For (b), the improved
control strategy has reduced the variability; consequently, the set point can be moved closer to the
limit, and the process can be operated closer to the optimum operating condition.
The standard process control techniques of Level 3a may not be adequate for difficult
control problems that have serious process interactions and inequality constraints. For these
situations, the advanced control techniques of Level 3b, multivariable control and constraint
control, should be considered. In particular, the model predictive control (MPC) strategy was
developed to deal with both process interactions and inequality constraints. MPC is the subject of
Chapter 20.

5. Real-time Optimization (Level 4)


The optimum operating conditions for a plant are determined as part of the process design.
But during plant operations, the optimum conditions can change frequently owing to changes in
equipment avail-ability, process disturbances, and economic conditions (for example, raw material
costs and product prices). Consequently, it can be very profitable to recalculate the optimum
operating conditions on a regular basis. This Level 4 activity, real-time optimization (RTO), is the
subject of Chapter 19. The new optimum conditions are then implemented as set points for
controlled variables.
The RTO calculations are based on a steady-state model of the plant and economic data
such as costs and product values. A typical objective for the optimization is to minimize operating
cost or maximize the operating profit. The RTO calculations can be performed for a single process
unit and/or on a plantwide basis.
The Level 4 activities also include data analysis to ensure that the process model used in
the RTO calculations is accurate for the current conditions. Thus, data reconciliation techniques
can be used to ensure that steady-state mass and energy balances are satisfied. Also, the process
model can be up-dated using parameter estimation techniques and recent plant data (Chapter 7).

6. Planning and Scheduling (Level 5)


The highest level of the process control hierarchy is concerned with planning and
scheduling operations for the entire plant. For continuous processes, the production rates of all
products and intermediates must be planned and coordinated, based on equipment constraints,
storage capacity, sales projections, and the operation of other plants, sometimes on a global basis.
For the intermittent operation of batch and semi-batch processes, the production control problem
becomes a batch scheduling problem based on similar considerations. Thus, planning and
scheduling activities pose difficult 'optimization problems that are based on both engineering
considerations and business projections.

Summary of the Process Control Hierarchy


The activities of Levels 1, 2 and 3a in Fig. 1.7 are required for all manufacturing plants,
while the activities in Levels 3b-Level 5 arc optional but can be very profitable. The decision to
implement one or more of these higher-level activities depends very much on the application and
the company. The decision hinges strongly on economic considerations (for example, a
cost/benefit analysis), and company priorities for their limited resources, both human and financial.
The immediacy of the activity de-creases from Level 1 to Level 5 in the hierarchy. However, the
amount of analysis and the computational requirements increase from the lowest level to the
highest level. The process control activities at different levels should be carefully coordinated and
require information transfer from one level to the next. The successful implementation of these
process control activities is a critical factor in making plant operation as profitable as possible.

1.5 AN OVERVIEW OF CONTROL SYSTEM DESIGN


In this section, we introduce some important aspects of control system design. However, it
is appropriate first to describe the relationship between process design and process control.
Traditionally, process design and control system design have been separate engineering
activities. Thus, in the traditional approach, control system design is not initiated until after
plant design is well underway and major pieces of equipment may even have been ordered.
This approach has serious limitations because the plant design determines the process
dynamics as well as the operability of the plant. In extreme situations, the process may be
uncontrollable, even though the design appears satisfactory from a steady-state point of view.
A more desirable approach is to consider process dynamics and control issues early in the
process design. The interaction between process design and control is analyzed in more detail
in Chapters 10, 23, and 24.
Next, we consider two general approaches to control system design:
1. Traditional Approach. The control strategy and control system hardware arc selected
based on knowledge of the process, experience, and insight. After the control system is
installed in the plant, the controller settings (such as controller gain IC, in Eq. 1-4) are
adjusted. This activity is referred to as controller tuning.
2. Model-Based Approach. A dynamic model of the process is first developed that can
be helpful in at least three ways: (i) It can be used as the basis for model-based
controller design methods (Chapters 12 and 14); (ii) the dynamic model can be
incorporated directly in the control law (for example, model predictive control); and
(iii) the model can be used in a computer simulation to evaluate alternative control
strategies and to determine preliminary values of the controller settings.

In this book, we advocate the philosophy that, for complex processes, a dynamic model of
the process should be developed so that the control system can be properly designed. Of course,
for many simple process control problems controller specification is relatively straightforward and
a detailed analysis or an explicit model is not required. For complex processes, however, a process
model is in-valuable both for control system design and for an improved understanding of the
process. As mentioned earlier, process control should be based on process understanding.
The major steps involved in designing and installing a control system using the model-
based approach are shown in the flow chart of Fig. 1.9. The first step, formulation of the control
objectives, is a critical decision. The formulation is based on the operating objectives for the plants
and the process constraints. For example, in the distillation column control problem, the objective
might be to regulate a key component in the distillate stream, the bottoms stream, or key
components in both streams. An alternative would be to minimize energy consumption (e.g., heat
input to the reboiler) while meet product quality specifications on one or both product streams.
The inequality constraints should inch upper and lower limits on manipulated variables, conditions
that lead to flooding or weeping in the column, and product impurity levels.
After the control objectives have been formulated, a dynamic model of the process is
developed. The mimic model can have a theoretical basis, for example, physical and chemical
principles such as conservation laws and rates of reactions (Chapter 2), or the model can be
developed empirically from experimental data (Chapter 7). If experimental data are available, the
dynamic model should be validated, with the d and the model accuracy characterized. This latter
information is useful for control system design and tuna.
The next step in the control system design is to devise an appropriate control strategy that
will m the control objectives while satisfying proem constraints. As indicated in Fig. 1.9, this
design activity is both an art and a science. Process understanding and the experience and
preferences of the design team are key factors. Computer simulation of the controlled process is
used to screen alternative control strategies and to provide preliminary estimates of appropriate
controller settings.

Finally, the control system hardware and instrumentation are selected, ordered, and
installed in the plant. Then the control system is tuned in the plant using the preliminary estimates
from the design step as a starting point. Controller tuning usually involves trial-and-error
procedures as described in Chapter 12.

SUMMARY
In this chapter we have introduced the basic concepts of process dynamics and process
control. The process dynamics determine how a process responds during transient conditions, such
as plant start-ups and shutdowns, grade changes, and unusual disturbances. Process control enables
the process to be maintained at the desired operating conditions, safely and efficiently, while
satisfying environmental and product quality requirements. Without effective process control, it
would be impossible to operate large-scale industrial plants.
Two physical examples, a continuous blending system and a distillation column, have been
used to introduce basic control concepts, notably, feedback and feedforward control. We also
motivated the need for a systematic approach for the design of control systems for complex
processes. Control system development consists of a number of separate activities that are shown
in Fig. 1.9. In this book we advocate the design philosophy that, for complex processes, a dynamic
model of the process should be developed so that the control system can be properly designed.
A hierarchy of process control activities was presented in Fig. 1.7. Process control plays a
key role in ensuring process safety and protecting personnel, equipment, and the environment.
Controlled variables are maintained near their set points by the application of regulatory control
techniques and advanced control techniques such as multivariable and constraint control. Real-
time optimization can be employed to determine the optimum controller set points for current
operating conditions and constraints. The highest level of the process control hierarchy is
concerned with planning and scheduling operations for the entire plant. The different levels of
process control activity in the hierarchy are related and should be carefully coordinated.
ELEMENTS OF THE
DESIGN OF A
CONTROL SYSTEM
Elements of the Design of a Control System
* PPT

PROCESS CONTROL ACTIVITIES

Hierarchy

Process:
1. Measurement and Actuation < 1 second
2. Safety and Environmental/Equipment Protection < 1 second
3a. Regulatory Control seconds – minutes
3b. Multivariable and Constraint Control minutes – hours
4. Real-Time Optimization hours – days
5. Planning and Scheduling days – months

1. MEASUREMENT AND ACTUATION


• Measurement devices (e.g sensors and transmitters) and actuation equipment (e.g.
control valves)

2. SAFETY AND ENVIRONMENTAL/EQUIPMENT PROTECTION


• Process safety relies on the principle of multiple protection layers
• Safety equipment operates independently of the regular instrumentation used for
regulatory control
• Environment regulations
• Philippine Clean Air Act

POLLUTANTS STANDARD MAXIMUM METHODS OF


APPLICABLE TO PERMISSIBLE ANALYSIS
SOURCE LIMITS
(mg/NCM)
1. Antimony Any source 10 as Sb AAS
and its
compounds
2. Arsenic Any source 10 as As AAS
and its
compounds
3. Cadmium Any source 10 as Cd AAS
and its
compound
4. Carbon Any industrial 500 as CO Orsat analysis
monoxide source
5. Copper and Any industrial 100 as Cu AAS
its source
compounds

• Continuous Emission Management System (CEMS)

Typical layers of protection in a modern chemical plant

3a. REGULATORY CONTROL


• Successful operation of a process requires that key process variables such as
FLPT and compositions be operated at or close to their set points.
• Achieved by applying standard feedback and feedforward control techniques.
3b. MULTIVARIABLE AND CONSTRAINT CONTROL
2 distinguishing characteristics
• Significant interactions occur among key process variables
• Inequality constraint exist for manipulated and controlled variables.

4. REAL-TIME OPTIMIZATION (RTO)


• The optimum operating conditions for a plant are determined as part of the process
design. But during plant operations, the optimum conditions can change frequently owing
to changes in equipment availability, process disturbance, and economic conditions.
• RTO calculations are based on steady-state model of the plant and economic data such as
costs and product values.
• Typical objective: To minimize operating cost or maximize the operating profit.
5. PLANNING AND SCHEDULING
• Highest level of the process control hierarchy
• Planning and scheduling activities pose difficult optimization problems that are based on
both engineering considerations and business projections.

CONTROL SYSTEM DESIGN

OVERVIEW OF CONTROL SYSTEM DESIGN

GENERAL REQUIREMENTS
1. SAFETY
• It is imperative that industrial plants operates safely to promote the well-being of
people and equipment within the plant and in the nearby communities.
2. ENVIRONMENTAL REGULATIONS
• Industrial plants must comply with environmental regulations concerning the
discharge of gases, liquids, and solids beyond the plant boundaries.
3. PRODUCTION SPECIFICATIONS AND PRODUCTION RATE
• In order to be profitable, a plant must make products that meet specifications
concerning product quality and production rate
4. ECONOMIC PLANT OPERATION
• It is an economic reality that the plant operation over long periods of time must be
profitable. Thus, the control objectives must be consistent with the economic
objectives.
5. STABLE PLANT OPERATION
• The control system should facilitate smooth, stable plant operation without
excessive oscillation in key process variables.

WHAT IS A CONTROL SYSTEM?


A control system is a system, which provides the desired response by controlling the output.
WHAT CONTROL SYSTEMS CAN DO?
1. Control a variable to obtain the required value
• Setting the required temperature for a room
2. Control the sequence of events
• Setting the controls on the automatic clothes washing machine
3. Control whether an event occurs or not
• Safety lock on the door of washing machine

CLASSIFICATION OF CONTROL SYSTEM


1. BASED ON TIME
• CONTINUOUS TIME CONTROL SYSTEM- all the signals are continuous in
time

• DISCRETE TIME CONTROL SYSTEM- there exists one or more discrete time
signals

2. BASED ON NUMBER OF INPUTS/OUTPUTS


• SISO – Single Input Single Output- have one input and one output

• MIMO – Multiple Input Multiple Output- more than one input and more than one
output

3. BASED ON FEEDBACK PATH


• OPEN LOOP CONTROL SYSTEM- output is not fed-back to the input. So, the
control action is independent of the desired output.
➢ Control Element - This determines the action to be taken as a result of the
input of the required value signal to the system.
➢ Correction Element - This has an input from the controller and gives an
output of some action designed to change the variable being controlled
➢ Process - This is the process of which a variable is being controlled

• CLOSED LOOP CONTROL SYSTEM- output is fed back to the input. So, the
control action is dependent on the desired output

➢ Comparison Element - Compares the required value of the variable


being controlled with the measured value of what is being achieved
and produces an error signal
➢ Control law implementation element - Determines what action to take
when an error signal is received.
− Proportional-Integral-Derivative
− Control Mode Controller: sometimes used for combination of
the comparison element
➢ Correction Element - Final control element
− Produces a change in the process which aims to correct or
change the controlled condition
➢ Process- The system in which there is a variable that is being
controlled
➢ Measurement element - Produces a signal related to the variable
condition of the process that is being controlled

CLASSIFICATION OF CONTROL SYSTEM

Open Loop Control Systems Closed Loop Control Systems


Control action is independent of the desired Control action is dependent of the desired
Output. output.
Feedback path is not present. Feedback path is present.
These are also called as non-feedback These are also called as feedback control
control systems. systems.
Easy to design. Difficult to design.
These are economical. These are costlier.
Inaccurate. Accurate.

CLASSIFICATION OF CONTROL STRATEGIES

1. FEEDBACK CONTROL SYSTEM


− responsible for processing the feedback signals which further act as an input to
the system
− Feature: controlled variable is measured and the measurement is used to adjust
the manipulated variable.
− Disturbance variable is not measured.
− Corrective action occurs regardless of the source of the disturbance

2. FEEDFORWARD CONTROL SYSTEM


− Feature: the disturbance variable is measured, but the controlled variable is not.
− Corrective action is taken before the controlled variable deviates from the set
point. It rejects the disturbances before they affect the controlled variable

No. FEEDBACK CONTROL FEED FORWARD CONTROL


SYSTEM SYSTEM
1 In feedback system output depends on In feed forward system the signal is
the generated feedback signal. passed to some external load.
2 Measure of disturbances in the system Measure of disturbances in the
is not needed system is needed
3 All the disturbances are detected All the disturbances are not
detected
4 The loop in a feedback system is a The loop in a feedback system is a
closed loop. open loop.
5 It focuses on the output of the system. It focuses on the input of the
system.
6 The variables are adjusted on the basis The variables are adjusted on the basis of
of errors. knowledge.

CONTROL SYSTEM DESIGN

3 MAIN STEPS
1. Select controlled, manipulated, and measured variables.
• For purposes of control system design, it is convenient to classify process
variables as either input variables or output variables
• There are criteria considered in selection (Chapter 10).

2. Choose the control strategy and control structure


• Most widely used process control strategy is multi-loop control.
• The key design decision for multi-loop control is to determine an appropriate
control structure, that is, to find a suitable pairing of controlled and manipulated
variables.

3. Specify controller settings


• Final step in the control system design
• Proportional-Integral-Derivative Control Models
UNDERSTANDING CONTROL SYSTEMS, PART 1: OPEN-LOOP CONTROL
SYSTEMS
*YT VIDEO*

One of the devices you use daily is your toaster. The bread’s color has changed based on
the timer setting you chose. Here, the toaster with the bread is an open-loop system that takes an
input, time, and gives an output, bread color.
Considering that don’t know what setting the timer should be at to get your desired bread color.
You can do a couple of experiments.
1. input different timer settings, and wait until your output settles to a value,
2. then mark your findings on this plot.

If you now fit a curve through these points, this represents the model of the toaster with the
bread at steady state. Remember that you want to find the time setting for your desired bread color.

To find this mathematically, let’s equate the input to u and the output to y. In the
experiments, for different values of u, you found the corresponding values of y. So you can write
y as a function of u.

But now you want to do the opposite. Given the desired value of y, you want to calculate
the value of u. In mathematical terms, this corresponds to taking the inverse of the function.
Therefore, to calculate the time you need for your desired bread color, you take the inverse of it.

In this open-loop system, the handle position is the input and water temperature is the
output.
You turn the handle to different positions, and based on this the water temperature changes.

Open-loop control seems to work perfectly. All you need is to do a couple of experiments
and, once you have an idea of the model of your system, you can easily find the input needed to
get a desired output.
But are there any situations where open-loop control may fail? Reconsider the toaster
example.
Through trial and error, you found that setting the timer to three gives you your desired
bread color. But what happens when you use a different type of bread, like a bagel? You may end
up having a burnt bagel because the time settings you found by trial and error were for a slice of
bread.

Mathematical interpretation of this is through experiments you found the function f that
determines the relationship between time and color. But now that you use a different type of bread,
this relationship is not represented by f anymore, but another function, g.

Therefore, if you use the inverse of f, you’re not going to get your desired color, because the input
needed to get the desired output is now calculated by taking the inverse of g.
This shows that open-loop control fails if you have variations in your system.
Now let’s go back to the shower example. You positioned the handle such that little Timmy
gets a nice warm shower, but what happens when someone runs the dishwasher? Water gets
freezing cold. The reason is that when the dishwasher is running, the hot water supply is used up
and therefore less hot water is available for the shower. So, the water temperature drops.

This is another shortcoming of open-loop control. Unexpected environmental changes


acting on the system affect the output.
Open-loop control is easy and conceptually simple. Through experiments, you find the
model of your system. If there are no variations or unexpected events, you know what input to
give to the system to get a desired output. However, when there are variations in the system or
unexpected events, open-loop control is unreliable.
UNDERSTANDING CONTROL SYSTEMS, PART 2: FEEDBACK CONTROL
SYSTEMS
*YT VIDEO
Today we’ll talk about feedback control. Let's go to your kitchen and see how you can use
feedback control to toast bread perfectly. You put a slice of bread in the toaster, set its timer level,
and then turn it on. Depending on how long you toast the bread, you can get different colors. But
you don't want just any color, you want to start your morning with this crispy, yummy toast. There
are two reasons this might be hard: If it's your first time using the toaster, you don't know how
long to toast the bread. Or, assuming you do know how long to toast it, next time you might open
the fridge and find a bagel or frozen bagel.

These are variations that you may face, but regardless of these variations, you would still
want to make perfect toast. So what if, instead of toasting the bread based on a timer setting, you
toast it based on its color. “But how?” You might wonder. If you continuously monitor the color
of the bread, you'll know when exactly to turn off the toaster. This is the basic idea behind a
feedback control system.

Let's try this on a slice of bread and a frozen bagel. You turn on the toaster and start
monitoring your bread. When the toast reaches the color you want, you turn off the toaster. Notice
that you didn’t have prior information on how long to toast the bread. Monitoring them allowed
you to tell when they have reached your desired color and when to turn off the toaster.

We will now read your mind. While you are monitoring the bread, you draw a plot in your
mind. On the y-axis you have the bread color that you're watching, and on the x-axis you have the
time.

This is what you want. Then, you start toasting, and this is what you see.
At each time instant, you compute an error between what you see and what you want. If
this error is not zero, you keep toasting. When ‘what you see’ overlaps with ‘what you want,’ the
error becomes zero. Your yummy toast is ready, so you turn off the toaster.

If we now project what you think in your mind onto the closed-loop structure here, we get
the complete feedback loop.

This part represents the comparison you make between ‘what you see’ and ‘what you
want’. You compute the difference between monitored and desired bread color, and this gives you
the error. Then, based on the error, you decide whether to keep the toaster on or turn it off.

Next, we will switch rooms to see another example of feedback control and how it
compensates for unexpected events. After eating your yummy toast, you're ready to take a warm
shower. Similar to the previous example, you have a desired water temperature. By trial and error,
you find the right position for the shower handle. You're planning to use this handle position for
future showers, as well. But what happens when someone runs the dishwasher the next time you’re
taking a shower? In this situation, the hot water is used up, and therefore the shower gets freezing
cold. Let’s go back to the time where the dishwasher isn’t running yet and see how feedback
control can compensate for this unexpected event.

The water temperature is at your desired value; someone runs the dishwasher. Through
your skin you sense that the water temperature drops. The error is now greater than zero. To
compensate, you turn the shower handle towards the hot side and, as the temperature increases to
the desired value, the error gets smaller. And the smaller the error gets, the smaller adjustments
you make to the shower handle. If you now want to fully automate this process, you can use a
thermocouple that measures the water temperature, you can use a thermocouple that measures the
water temperature and then, based on the error, a controller can adjust the shower handle.

To summarize, we’ve seen how feedback control works, how it handles variations in the
system, and how it compensates for unexpected events.
UNDERSTANDING CONTROL SYSTEMS, PART 3: COMPONENTS OF A
FEEDBACK CONTROL SYSTEM
*YT VIDEO
You’re invited to a party and don’t want to miss a minute of it. You want to get there as
fast as possible but without exceeding the speed limit, which is 50 mph. In order to get to the party
as soon as possible, you’ll need to maintain your speed at 50 mph. You may not have noticed it
before but actually you drive your car in closed loop.
Let’s first identify the components of this closed- loop system and then we’ll learn the basic
terminology. When you step on the gas pedal, your car’s speed starts to change. When you step on
the gas pedal, your car’s speed starts to change. You know that you want to be going 50 mph, let’s
show your desired speed on the plot with green. Notice that it remains constant over time. Next,
you look at the speedometer to see at what speed you are. Let’s show the monitored speed on this
plot. Then you compare it to what you want. This is done by finding the difference between the
desired and monitored speed. And it is called the error.
Based on the error you make a decision. In this case, the error is not zero, therefore you
decide to apply more gas to get to your desired speed. And this in turn influences the speed of the
car. The smaller the error gets, the less gas you apply. And once you get to the desired speed, you
keep your foot still on the gas pedal to maintain your speed. Oh, I see, you’ve already started
dreaming. I’m sorry to interrupt but let’s go back to reality.
Now that we identified the components, let’s talk about how they are referred in control
theory. In a feedback control system, the goal is to control the output to a desired value and that
value is referred as desired output.
Other common terms are reference and setpoint. The object whose output you try to
control is called the plant. The output of the plant is measured by the component called the sensor,
and the measured output is then compared to the desired output. The difference between the desired
and measured output is the error.
Based on the computed error, the controller decides on the control action. And it commands
an input, which is then sent to the actuator. The actuator is how the controller influences the plant.
In some situations, you’ll see that the actuator, sensor, and plant dynamics are merged into one
and referred to as the “plant”. This may sound pessimistic but life is far from perfect, sometimes
things may go wrong.
For our example here, it is the road grade. Imagine you’re going uphill. Your speed will
start to drop due to the hill and you will need to press down further on the gas pedal to maintain
your speed. Unexpected environmental changes causing the system output to behave in an
undesired way, such as the road grade in this example, is called “disturbance”. Similarly, a sticky
gas pedal, icy roads, or windy weather act as disturbance on the car.
Another thing that may go wrong is related to measurement. While you’re monitoring the
speed, what if your vision is a little blurry because you forgot your glasses at home? The speed
you’re monitoring becomes noisy and fluctuates between values. This is referred as “noise,” which
is an unwanted signal affecting your measured output value. Driving manually makes you tired?
Cruise control can make your life easier.
Let’s see how this system can be converted to an automatic control system. Now instead
of monitoring the speed manually, you can have a device that provides an electronic measurement
of the speed. Your brain was acting as a controller. Now, the engine control unit becomes the
controller that computes a control action based on the error. The actuator now becomes a
mechanical device that does what your foot used to do. In this case, the actuator is a motor that
moves the gas pedal up and down based on the voltage computed by the controller. Now that you
have the cruise controller fully engaged, enjoy your ride to the party!
LECTURE 3: ELEMENTS OF THE DESIGN OF A CONTROL SYSTEM
*YT VIDEO

Hierarchy of the process control activities


1. measurement and actuation
2. safety and environmental or equipment protection
3a. regulatory control
3b. multivariable and constraint control
4. real-time optimization
5. planning and scheduling

✓ alongside with these activities are the time scale from which the activity can be done.

1. Measurement and actuation

• Measurement devices such as sensors and transmitters and actuation equipment


such as control bar control valves are used to measure process variables and implement
the calculated control.
• Measurement devices (e.g sensors and transmitters) and actuation equipment (e.g.
control valves)
• the measurement and actuation functions are indispensable part of any control
system

The orifice plate is connected to a sensor that measures the pressure difference of the flow. After
the flow transmitter finishes measuring the flow it will then be passed to the controller. The
controller will then apply the corresponding action before transferring it to the control valve
(actuator), which will give the corresponding action to the control valve (partially open or partially
closed).

2. Safety and environmental or Equipment protection


• Process safety relies on the principle of multiple protection layers
• Safety equipment operates independently of the regular instrumentation used for
regulatory control
• Environment regulations
• Basic function is to measure process variables.
• Ensure the safety of the plant and the workers

Environment Regulations in the Philippines

A. Philippine Clean Air Act

B. Continuous Emission Management System (CEMS)


• Present in refinery and power plants
• Analyzer shelter (instrument for measuring CO, CO2 and other pollutants)

Typical layers of protection in a modern chemical plant


• Protection layers for a typical process are shown in the order of activation expected as a
hazardous condition is approached.
• ESD (Emergency Shutdown)
• SIS (Safety Interlock System)
• Basic controls, process alarms and operators to supervise the whole operations.
• Critical alarms, operator supervision and manual interventions. (If a certain process
variable reaches a certain condition, it is possible for an emergency to occur. Before it
happens, there will be alarms that will signal for the need of a corrective action to prevent
a scenario from happening.)
• In addition to alarms and manual interventions, there are also SIS and ESD.
• During the design, it can be determined what part of the plants can have the possibility of
having critical conditions and emergency.
• Pumps, motors and compressors typically have SIS and ESD.
• Physical Protection or relief devices, high pressure vessels typically have relief devices.
• Physical protection (dikes) Dikes are usually used for tanks. Dikes will help in containing
the spilled (e.g. crude oil).
o Unsure if spilled products are recoverable.
• Plant emergency response (e.g fire respondents)
• Community emergency response (in case the scenario worsens)

3a. Regulatory Control


• Successful operation of a process requires that key process variables such as FLPT (Flow
level, Pressure and Temperature) and compositions be operated at or close to their set
points.
• Achieved by applying standard feedback and feedforward control techniques.

3b. Multivariable and Constraint Control


• 2 distinguishing characteristics
1. Significant interactions occur among key process variables. (disturbances)
2. Inequality constraint exist for manipulated and controlled variables. (upper and
lower limits)

• E.g. Boiling of water, Reactor temperature (upper limit to avoid the undesired side reaction
or catalyst degradation and lower limit to ensure that the reactions proceed)
• Upper and lower limit are crucial.

4. Real-Time Optimization (RTO)


• The optimum operating conditions for a plant are determined as part of the process design.
But during plant operations, the optimum conditions can change frequently owing to
changes in equipment availability, process disturbance, and economic conditions.
• Equipment availability, maintenance of equipment
• Process Disturbance
• Economic Conditions, always considered in RTO. Net sales. Demands.
• RTO calculations are based on steady-state model of the plant and economic data such as
costs and product values.
• Typical objective: To minimize operating cost or maximize the operating profit.

5. Planning and Scheduling



Highest level of the process control hierarchy.

Planning and scheduling activities pose difficult optimization problems that are based on
both engineering considerations and business projections.
OVERVIEW OF CONTROL SYSTEM DESIGN
Management Computer Computer Vendor
objectives Simulation Simulation information

Formulate Develop Devise Select control Install Adjust FINAL


Control process control hardware Control controller CONTROL
Objectives control strategy and software System settings SYSTEM

Information from Physical and Process control


existing plants (if Chemical theory
available) Principles

Plant data (if Experience


available) with existing
plants (if
available)

Aspen HYSYS - computer software used for process simulation.


GENERAL REQUIREMENTS - CONTROL SYSTEM DESIGN
1. Safety
It is imperative that industrial plants operates safely to promote the well-being of people and
equipment within the plant and in the nearby communities.
2. Environmental Regulations
Industrial Plants must comply with environmental regulations concerning the discharge of
gases, liquids, and solids beyond the plant boundaries.
Philippine Clean Air Act, Clean Water Act, Hazardous and Toxic Waste Management, Solid
Waste Management.
3. Product Specifications and Production Rate
In order to be profitable, a plant must make products that meet specifications concerning
product quality and production rate.
4. Economic Plant Operation
It is an economic reality that the plant operation over long periods of time must be profitable.
Thus, the control objectives must be consistent with the economic objectives.
5. Stable Plant Operation
The control system should facilitate smooth, stable plant operation without excessive
oscillation in key process variables.
WHAT IS A CONTROL SYSTEM?
A control system is a system which provides the desired response by controlling the output.

WHAT CONTROL SYSTEMS CAN DO?


• Control a variable to obtain the required value.
- Setting the required temperature for a room.
• Control the sequence of events
- Setting the controls on the automatic clothes washing machine.
• Control whether an event occurs or not
- Safety lock on the door of washing machine.

CLASSIFICATION OF CONTROL SYSTEM


BASED ON TIME

• Continuous Time Control System - all the signals are continuous in time
• Discrete Time Control System - there exists one or more discrete time signals
BASED ON NUMBER OF INPUTS/OUTPUTS

• SISO - Single Input Single Output - have one input and one output
• MIMO - Multiple Input Multiple Output - more than one input and more than one
output

There are also instances of multiple inputs and single output. Can be encountered in cascading
controls.
BASED ON FEEDBACK PATH
Open Loop Control System
- Output is not fed-back to the input. So, the control action is independent of the desired
output.

Control Element – Correction Element – Process


BASIC ELEMENTS OF AN OPEN LOOP CONTROL SYSTEM
- Control Element
o This determines the action to be taken as a result of the input of the required
value signal to the system.
- Correction Element
o This has an input from the controller and gives an output of some action
designed to change the variable being controlled.
- Process
o This is the process of which a variable is being controlled.
Closed Loop Control System
- Output is fed back to the input. So, the control action is dependent on the desired output.
Elements of a Closed Loop
- Comparison Element
- Control Law Implementation Element
- Correction Element
- Process
- Measurement Element
Comparison Element
– compares the required value of the variable being controlled with the measured value of
what is being achieved and produces an error signal
Control Law Implementation Element
– determines what action to take when an error signal is received.
- Proportional-Integral-Derivative Control Mode
- Controller: sometimes used for combination of the comparison element
Correction Element
– final control element
- Produces a change in the process which aims to correct or change the controlled
condition
Process
– the system in which there is a variable that is being controlled
Measurement Element
– produces a signal related to the variable condition of the process that is being controlled
Difference Between Open and Closed Loop
Open Loop Control Systems Closed Loop Control Systems
- Control action is independent of the - Control action is dependent of the
desired output desired output
- Feedback path is not present - Feedback path is present
- These are also called as non- - These are also called as feedback
feedback control systems. control systems
- Easy to design - Difficult to design
- These are economical - These are costlier
- Inaccurate - Accurate

CLASSIFICATION OF CONTROL STRATEGIES


Feedback Control System
- Responsible for processing the feedback signals which further act as an input to the
system.
- Feature: controlled variable is measured and the measurement is used to adjust the
manipulated variable
- Disturbance variable is not measured
- Corrective action occurs regardless of the source of the disturbance

Feedforward Control System


- Feature: the disturbance variable is measured, but the controlled variable is not.
- Corrective action is taken before the controlled variable deviates from the setpoint. It
rejects the disturbances before they affect the controlled variable.

Difference of Feedback Control System and Feedforward Contol System


No. Feedback Control Feed Forward
System Control System
1 In feedback system In feed forward
output, depends on system, the signal is
the generated passed to some
feedback signal external load
2 Measure of Measure of
disturbances in the disturbances in the
system is not needed system is needed
3 All the disturbances All the disturbances
are detected are not detected
4 The loop in the The loop in a feed
feedback system is a forward is an open
closed loop loop
5 It focuses on the It focuses on the
output of the system input of the system
6 The variables are The variables are
adjusted on the basis adjusted on the basis
of errors of knowledge

Control System Design


3 Main Steps:

1. Select controlled, manipulated, and measured variables. For purposes of control system design,
it is convenient to classify process variables as either input variables or output variables. There are
criteria considered in selection (Chapter 10).
2. Choose the control strategy and control structure. Most widely used process control strategy is
multi-loop control. The key design decision for multi-loop control is to determine an appropriate
control structure, that is, to find a suitable pairing of controlled and manipulated variables.

Example of Multi-Loop Control (Closed Loop/ Feedback Control System)

3. Specify controller settings. Final step in control system design. Proportional-Integral-


Derivative Control Models
[Process Dynamics and Control 2E, Seborg]
Chapter 10 | OVERVIEW OF CONTROL SYSTEM DESIGN

10.1. INTRODUCTION
The development of a control system begins with a critical decision, the formulation of the
control objectives. As indicated in Chapter 1, control objectives are based on
management/financial objectives, process knowledge, and the operational requirements for the
plant. Although the specific control objectives vary from plant to plant, there are a number of
general requirements:
1. Safety. It is imperative that industrial plants operate safely so as to promote the well-
being of people and equipment within the plant and in the nearby communities. Thus,
plant safety is always the most important control objective and is the subject of Section
10.5.
2. Environmental Regulations. Industrial plants must comply with environmental
regulations concerning the discharge of gases, liquids, and solids beyond the plant
boundaries.
3. Product Specifications and Production Rate. In order to be profitable, a plant must
make products that meet specifications concerning product quality and production rate.
4. Economic Plant Operation. It is an economic reality that the plant operation over long
periods of time must be profitable. Thus, the control objectives must be consistent with
the economic objectives. For example, if a plant can sell all its product, then it is
desirable to maximize its production rate. But in a market-limited situation, it might be
more important to increase product quality, improve yields, or reduce manufacturing
costs such as utility costs.
5. Stable Plant Operation. The control system should facilitate smooth, stable plant
operation without excessive oscillation in key process variables. Thus, it is desirable to
have smooth, rapid set-point changes and rapid recovery from plant disturbances such
as changes in feed composition.

10.1.1 Steps in Control System Design


After the control objectives have been formulated, the control system can be
designed. The design procedure consists of three main steps:
1. Select controlled, manipulated, and measured variables.
2. Choose the control strategy and control structure.
3. Specify controller settings.
Criteria for the selection of controlled, manipulated, and measured variables are
considered in Section 10.4. The controlled and manipulated variables are usually measured
on-line. If it is not feasible to measure a controlled variable such as a chemical composition,
it may be possible to estimate it from measurements of other variables such as
temperatures, pressures, and flow rates. This approach; referred to as inferential control or
soft sensor, is considered in Chapter 16. Measurements of other process variables can
sometimes be used to good advantage. In particular, measurements of important
disturbance variables can provide the basis for feed-forward control, as described in
Chapter 15.
The most widely used process control strategy is multiloop control. A multiloop
control system consists of a set of PI or PID controllers, one for each controlled variable.
The key design decision for multiloop control is to determine an appropriate control
structure, that is, to find a suitable pairing of controlled and manipulated variables. This
important topic is considered in detail in Chapter 18. The traditional multiloop control
structure has proved to be satisfactory for a wide variety of control problems. But there are
classes of process control problems where an advanced control strategy can provide much
better control and also be more profitable. Some common examples are:
1. Processes with slow dynamics and measurable disturbances. For this class of
problems, the addition of feed forward control to the multiloop control scheme
should be considered.
2. Processes with strong interactions between process variables. If key process
variables strongly interact, multivariable control strategies can be quite effective.
In multivariable control, each manipulated variable is adjusted based on
measurements of two or more controlled variables. By contrast, in multiloop control
each manipulated variable is paired with a single controlled variable. Multivariable
control strategies are considered in Chapters 18 and 20.
3. Processes that exhibit strongly nonlinear behavior. For this class of control
problems, nonlinear control techniques can be highly effective and are considered
in Chapter 16.
4. Processes where constraints on process variables must be satisfied. Upper and
lower limits are imposed on controlled and manipulated variables for a variety of
reasons that include equipment limitations and operating objectives. For example,
the maximum flow rate may be limited by the pump and control valve
characteristics, while product specifications and environmental regulations place
restrictions on impurity levels in process streams. Inequality constraints can be
accommodated by using model predictive control (see Chapter 20).
The final step in the control system design is to specify the controller settings. For
example, if PI control is considered, the controller gain Kc and the integral time τf will be
assigned nominal values. The nominal settings are often adjusted (or tuned) after the
control system is installed, using the controller tuning strategies described in Chapter 12.

10.2. THE INFLUENCE OF PROCESS DESIGN ON PROCESS CONTROL


Traditionally, process design and control system design have been separate engineering
activities. Thus, in the traditional approach, control system design is not initiated until after the
plant design is
well underway and major pieces of equipment may even have been ordered. This approach has
serious limitations because the plant design determines the process dynamic characteristics, as well
as the operability of the plant. In extreme situations, the plant may be uncontrollable even though
the process design appears satisfactory from a steady-state point of view (Shinskey, 1982).
A more desirable approach is to consider process dynamics and control issues early in the
plant design. This interaction between design and control has become especially important for
modern processing plants, which tend to have a large degree of material and energy integration
and tight performance specifications. As Hughart and Kominek (1977) have noted: ''The control
system engineer can make a major contribution to a project by advising the project team on how
process design will influence the process dynamics and the control structure." The interaction of
the process design and control system design teams is considered in Chapters 23 and 24.

10.2.1 Heat Integration of Process Units


We now consider two specific examples of how process design affects process
dynamics and control. Because of the rise in fuel prices, there has been considerable interest
in reducing energy costs of distillation trains by heat integration or thermal coupling of two
or more columns. Figure 10.1 compares a· conventional distillation system with a heat-
integrated scheme. Heat integration reduces energy costs by allowing the overhead stream
from Column 1 to be used as the heating medium in the reboiler for Column 2. However,
this column configuration is more difficult to control for two reasons. First, the process is
more highly interacting because process upsets in one column affect the other column.
Second, the heat integration configuration has one less manipulated variable available for
process control because the reboiler heat duty for Column 2 can no longer be independently
manipulated. But these disadvantages can be reduced by adding a small trim reboiler to
Column 2, or by utilizing advanced control strategies (Shinskey, 1984). The trim reboiler
acts in parallel with the main reboiler but has a separate supply of the heating medium that
can be manipulated. Thus, adding the trim reboiler regains the control degree of freedom
that was originally lost as a result of the heat integration.

A second type of heat integration is shown in Fig. 10.2b for a packed-bed reactor;
the more conventional design is shown in Fig. 10.2a. If the chemical reaction is exothermic,
energy costs can be reduced by using the hot product stream to heat the cold feed in a heat
exchanger. However, this reactor configuration has the same disadvantages as the energy-
integrated distillation system in Fig. 10.1, namely, one less manipulated variable and
unfavorable dynamic interactions. In particular, the feed-effluent heat exchanger
introduces positive feedback and the possibility of a thermal runaway because temperature
fluctuations in the effluent are transmitted to the feed stream. Thus, the rate of reaction and
the heat generation due to the reaction increase. In analogy with the previous distillation
column example, these disadvantages can be reduced by adding a trim heat exchanger (that
has a separate supply of the heating medium) after the feed-effluent heat exchanger. Again,
this modification regains the control degree of freedom that was lost as a result of heat
integration.

Example 10.1:
Two alternative temperature control schemes for a jacketed batch reactor are shown in Fig. 10.3.
Discuss the relative merits of the two strategies from a process operation perspective.

Solution:
The configuration in Fig. 10.3a has the serious disadvantage that the coolant circulation
rate varies, and thus the corresponding time delay for the coolant loop also varies. As Shinskey
(1996) has noted, when the time delay varies with the manipulated variable, a nonlinear oscillation
can develop. H the reactor temperature increases, the controller increases the coolant flow rate,
which reduces the time delay and causes a sharp temperature decrease. On the other hand, when
the reactor temperature is too low, the controller reduces the coolant flow rate, which increases the
time delay and results in a slow response. This nonlinear cycle tends to be repeated.
Shlnskey (1996) points out that this control problem can be solved by making a simple
equipment design change, namely, by adding a recirculation pump as shown in Fig. 10.3b. Now
the recirculation rate and process time delay are kept constant and thus are independent of the flow
rate of fresh cooling water. Consequently, the nonlinear oscillations are eliminated.

10.3. DEGREES OF FREEDOM FOR PROCESS CONTROL


The important concept of degrees of freedom was introduced in Section 2.3, in connection
with process modeling. The degrees of freedom Np is the number of process variables that must
be specified in order to be able to determine the remaining process variables. If a dynamic model
of the process is available, NF can be determined from a relation that was introduced in Chapter 2
𝑁𝐹 = 𝑁𝑉 + 𝑁𝐸
It is where NV is the total number of process variables, and NE is the number of independent
equations. For process control applications, it is very important to determine the maximum number
of process variables that can be independently controlled, that is, to determine the control degrees
of freedom, NFC:
Definition: The control degrees of freedom NFC is the number of process variables (e.g.,
temperatures, levels, flow rates, compositions) that can be independently controlled.
In order to make a clear distinction between NF and NFC, we will refer to NF as the model
degrees of freedom and NFC as the control degrees of freedom. Note that NF and NFC are related by
the following equation
𝑁𝐹 = 𝑁𝐹𝐶 + 𝑁𝐷
It is where ND is the number of disturbance variables (i.e., input variables that cannot be
manipulated). Common disturbance variables include ambient temperature and the feed
conditions, when the latter are determined .by an upstream process. Equation 10-2 can be
interpreted as a partitioning of the model degrees of freedom NF into manipulated inputs (the
control degrees of freedom, NFC) and disturbance inputs, ND.

10.3.1 Calculation of the Control Degrees of Freedom


We have noted that NFC can be calculated from Eqs. 10-1 and 10-2 providing that
a physical process model is available. But for complex processes, a physical process model
may not be available, and thus NF cannot be determined from Eq. 10-1. But even when an
unsteady-state model is available, calculating NF from Eq. 10-1 is error prone because it is
easy to write too many or too few equations, if the model contains hundreds of variables
and equations. Consequently, alternative approaches have been proposed for determining
the control degrees of freedom (Hanson et al., 1962; Tyreus, 1992; Ponton, 1994; Luyben,
1996; Larsson and Skogestad, 2000). In process control applications, the manipulated
variables are usually flow rates that are adjusted via control valves, pumps, compressors,
or connveyer belts (for solid materials). Energy inputs such as electrical heater duties are
also adjusted in some control applications. Thus, NFC is closely related to the number of
material and energy streams that can be adjusted. These observations lead to the following
general rule.
General Rule: For many practical problems, the control degrees of freedom NFC is equal
to the number of independent material and energy streams that can be manipulated.
Note that it is important that the manipulated streams be independent. For example,
if a process stream splits, or if two process streams merge to form a third stream, it is not
possible to adjust all three flow rates independently. These situations are shown in Fig.
10.4. Two examples illustrate the general rule.

10.3.2 Effect of Feedback Control


Next, we consider the effect of feedback control on the control degrees of freedom.
In general, adding a feedback controller (e.g., PI or PID) assigns a control degree of
freedom because a manipulated variable is adjusted by the controller. However, if the
controller set point is continually adjusted by a higher-level (or supervisory) control
system, then neither NF nor NFC change. To illustrate this point, consider the feedback
control law for a standard PI controller:
1 𝑡 ∗
𝑢(𝑡) = 𝐾𝑐 [𝑒(𝑡) + ∫ 𝑒(𝑡 ) 𝑑𝑡 ∗ ]
𝜏𝑙 0
It is where 𝑒(𝑡) = 𝑦𝑠𝑝 (𝑡) − 𝑦(𝑡) and 𝑦𝑠𝑝 is the set point. We consider two cases:
Case 1. The set point is constant, or only adjusted manually on an infrequent basis.
For this situation, ysp is considered to be a parameter instead of a variable.
Introduction of the control law adds one equation but no new variables because u and y are
already included in the process model. Thus, NE increases by one, NV is unchanged, and
Eqs. 10-1 and 10-2 indicate that NF and NFC decrease by one.
Case 2. The set point is adjusted frequently by a higher-level controller.
The set point is now considered to be a variable. Consequently, the introduction of
the control law adds one new equation and one new variable, ysp. Equations 10-1 and 10-2
indicate that NF and NFC do not change. The importance of this conclusion will be more
apparent when cascade control is considered in Chapter 16.

10.4. SELECTION OF CONTROLLED, MANIPULATED, AND MEASURED


VARIABLES
For purposes of control system design, it is convenient to classify process variables as
either input variables or output variables. The output variables (or outputs) are process variables
that typically are associated with exit streams or conditions within a process vessel (e.g.,
compositions, temperatures, levels, and flow rates). The important output variables should be
measured. Unfortunately, there are usually not enough degrees of freedom available to control all
of the measured variables. Consequently, a subset is selected to be the controlled variables.
Fortunately, for many complex processes having a large number of process variables, the process
can be reasonably well controlled by considering only a small number of controlled variables and
manipulated variables. This concept is referred to as partial control (Kothare et al., 2000).

By definition, the input variables (or inputs) are physical variables that affect the output
variables. For control system design and analysis, it is convenient to divide the input variables into
manipulated variables that can be adjusted and disturbance variables that are determined by the
external environment. The manipulated variables are typically flow rates. Common disturbance
variables include the feed conditions to a process and the ambient temperature. Typically, input
variables are associated with inlet streams (e.g., feed composition or feed fl.ow rate) or
environmental conditions (e.g., ambient temperature). However, an exit flow rate from a process
can also be an "input variable" (from a control point of view), if the flow rate is a manipulated
variable. For example, this situation occurs when the liquid level in a tank is controlled by adjusting
an exit flow rate.
A general representation of a control problem is shown in Fig. 10.7. In general, it is
desirable to have at least as many manipulated variables as controlled variables. But sometimes
this is not possible and special types of control systems need to be considered (see Chapter 16). It
may not be feasible to control all of the output variables for several reasons (Newell and Lee,
1989):
1. It may not be possible or economical to measure all of the outputs, especially
compositions.
2. There may not be enough manipulated variables (refer to the degrees of freedom
discussion in Section 10.3).
3. Potential control loops may be impractical because of slow dynamics, a low sensitivity
to available manipulated variables, or interactions with other control loops.
In general, controlled variables are measured on-line, and the measurements are used for
feedback control. But it is sometimes possible to control a variable that is not measured on-line by
using a process model (a soft sensor) to estimate its value from other measurements.

10.4.1 Selection of Controlled Variables


The consideration of plant and control objectives has led to a number of suggested
guidelines for the selection of controlled variables from the available output variables
(Hougen, 1979; Newell and Lee, 1989).
Guideline 1.
All variables that are not self-regulating must be controlled. In Chapter 5 a non-self-
regulating variable was defined to be an output variable that exhibits an unbounded
response after a sustained disturbance such as a step disturbance. A common example is
liquid level in a tank that has a pump on an exit line, as will be shown in Chapter 11. Non-
self-regulating variables must be controlled in order for the controlled process to be stable.
Guideline 2.
Choose output variables that must be kept within equipment and operating
constraints (e.g., temperatures, pressures, and compositions). The constraints originate
from safety, environmental, and operational requirements.
Guideline 3.
Select output variables that represent a direct measure of product quality (e.g.,
composition, refractive index) or that strongly affect it (e.g., temperature or
pressure).
Guideline 4.
Choose output variables that seriously interact with other controlled variables. The
steam header pressure that supplies steam to downstream units is a good example. If the
supply pressure is not well regulated, it can act as a significant disturbance to the
downstream units.
Guideline 5.
Choose output variables that have favorable dynamic and static characteristics. Thus,
output variables that have large measurement time delays, or are insensitive to the
manipulated variables, are poor choices. ·
Except for Guideline 1, these guidelines are not strict rules. Also, for a specific
situation the guidelines may be inconsistent and thus result in a conflict. For example, if
one output variable must be kept within specified limits for safety reasons (Guideline 2),
while a second interacts strongly with other output variables (Guideline 4), Guideline 2
would prevail due to safety considerations. Thus, the first output variable would be
controlled.

10.4.2 Selection of Manipulated Variables


Based on the plant and control objectives, a number of guidelines have been
proposed for the selection of manipulated variables from among the input variables
(Hougen, 1979; Newell and Lee, 1989):
Guideline 6.
Select inputs that have large effects on controlled variables. For conventional feedback
control systems (e.g., several PID controllers), we would like each manipulated variable to
have a significant, rapid effect on only one controlled variable. Thus, the corresponding
steady-state gain should be large. Ideally, the effects of the manipulated variable on the
other controlled variables should be negligible (that is, these steady-state gains should be
approximately zero).
It is also important that each manipulated variable can be adjusted over a
sufficiently large range of conditions. For example, if a distillation column has a steady-
state reflux ratio of five, it will be much easier to control the reflux drum level by
manipulating the reflux flow rather than the distillate flow rate because a larger range of
disturbances in the vapor flow rate to the reflux drum (five times larger) can be handled.
However, the effect of this choice on the product compositions must also be considered in
making the final decision.
Guideline 7.
Choose inputs that rapidly affect the controlled variables. For multiloop control, it is
desirable that each manipulated variable have a rapid effect on its corresponding controlled
variable. Thus, any time delay and time constant(s) should be small, relative to the
dominant process time constant.
Guideline 8.
The manipulated variables sliould affect the controlled variables directly rather than
indirectly. Compliance with this guideline usually results in a control loop with favorable
static and dynamic characteristics. For example, consider the problem of controlling the
exit temperature of a process stream that is heated by steam in a shell and tube heat
exchanger. It is preferable to throttle the steam flow to the heat exchanger rather than the
condensate flow from the shell because the steam flow rate has a more direct effect on the
steam pressure and the rate of heat transfer.
Guideline 9.
Avoid recycling of disturbances. As Newell and Lee (1989) have noted, it is preferable not
to manipulate an inlet stream or a recycle stream, because disturbances tend to be
propagated forward or recycled back to the process. This problem can be avoided by
manipulating a utility stream to absorb disturbances or an exit stream that allows the
disturbances to be passed downstream, provided that the exit stream changes do not unduly
upset downstream process units.
Note that these guidelines may be in conflict. For example, a comparison of the
effects of two inputs on a single controlled variable may indicate that one has a larger
steady-state gain {Guideline 6) but slower dynamics (Guideline 7). In this situation, a
tradeoff between static and dynamic considerations must be made in selecting the
appropriate manipulated variable from the two candidates.

10.4.3 Selection of Measured Variables


Safe, efficient operation of processing plants requires on-line measurement of key
process variables. Clearly, the controlled variables should be measured. Other output
variables are measured to provide additional information to the plant personnel or for use
in model-based control schemes such as inferential control. It is also desirable to measure
manipulated variables because they provide useful information for tuning controllers and
troubleshooting control loops, as will be discussed in Chapter 12. Measurements of
disturbance variables provide the basis for feedforward control schemes (see Chapter 15).
In choosing sensor locations, both static and dynamic considerations are important,
as discussed in Chapter 9.
Guideline 10.
Reliable, accurate measurements are essential for good control. There is ample
evidence in the literature (Hughart and Kominek, 1977; Hougen, 1979) that inadequate
measurements are a key contributor to poor control. Hughart and Kominek (1977) cite
common measurement problems that they observed in distillation column control: orifice
runs without enough straight piping, analyzer sample lines with too much time delay,
temperature probes located in insensitive regions, and flow rate measurement of liquids
that are at, or near, their boiling points can lead to liquid flashing at the orifice plate. They
note that these measurement problems can be readily resolved during the process design
stage but that improving a measurement location after the process is operating is extremely
difficult.
Guideline 11.
Select measurement points that have an adequate degree of sensitivity. As an
example, consider product composition control in a tray distillation column. If the product
composition cannot be measured on-line, it is often controlled indirectly by regulating a
tray temperature near that end of the column. But for high-purity separations, the location
of the temperature measurement point can be quite important. If a tray near an end of the
column is selected, the tray temperature tends to be insensitive because the tray
composition can vary significantly even though the tray temperature changes very little.
For example, suppose that the impurity in the vapor leaving the top tray has a nominal
value of 20 ppm. A feed composition change could cause the impurity level to change
significantly (for example, from 20 to 40 ppm) while producing only a negligible change
in the tray temperature. By contrast, if the temperature measurement point were moved to
a tray that is closer to the feed tray, the temperature sensitivity would be improved, but
now disturbances entering the column at either end (e.g., condenser or reboiler) would not
be detected as quickly.
Guideline 12.
Select measurement points that minimize time delays and time constants.
Reducing dynamic lags and time delays associated with process measurements improves
closed-loop stability and response characteristics. Hughart and Kominek {1977) have
observed distillation columns with the sample connection for the bottoms analyzer located
200 feet downstream from the column. This distance introduces a significant time delay
and makes the column difficult to control, particularly because the time delay varies with
the bottoms flow rate.

10.4.4 Control System Design Issues


Next, we briefly introduce a few key issues for control system design. These topics
will be considered in more detail in subsequent chapters.
In order to illustrate the development of a control strategy, consider the tray
distillation control example shown in Fig. 10.5. Suppose that the chief control objectives
are to control both product compositions, xD and xB. However, the liquid levels in the reflux
drum ho and the column base (or "sump") hD must be kept between upper and lower limits.
Also, the column pressure P must be controlled. We further assume that five process
variables can be manipulated: product flow rates, D and B, reflux flow rate R, and the flow
rates for the condenser and reboiler, qc and qh. We assume that the feed flow rate cannot be
manipulated because it is an exit stream from an upstream process.
1. Should xD and xB be measured on-line? On-line composition measurement is
often difficult, expensive, or not even feasible. As an alternative, we could measure
tray temperatures near the top and bottom of the column and control these
temperatures, instead of xD and xB. In this indirect approach, it is hoped that if the
tray temperatures (and column pressure) are tightly controlled, then the unmeasured
product compositions will stay close to their set points. This assertion is correct for
the distillation of binary mixtures, but it is not necessarily true for multicomponent
separations. An alternative approach is to use inferential control based on an
estimate of the unmeasured composition, as mentioned earlier.
2. Choice of control structure. A key decision is whether a conventional multiloop
control strategy will provide satisfactory control. If not, an advanced control
strategy is required. This decision should be based primarily on the control
objectives and knowledge of the static and dynamic behavior of the column. These
important issues are considered in Chapter 18.
3. Pairing of controlled and manipulated variables. Suppose that a multiloop
control strategy has been selected and that there are controlled variables and n
manipulated variables. The next step is to determine how the controlled and
manipulated variables should be paired. For the distillation column in Fig. 10.5, it
is reasonable to pair the controlled variables at one end of the column with
manipulated variables at the same end, based on dynamic considerations (Guideline
7). Thus, xB should be paired with either B or qc rather than with R or D. But it is
less clear whether B or qh should be selected as the manipulated variable. Similarly,
is it better to control XD by adjusting R or D? These issues are dealt with in Chapter
18.

10.5. PROCESS SAFETY AND PROCESS CONTROL


Process safety has been a primary concern of the process industries for decades. But in
recent years, safety issues have received increased attention for several reasons that include
increased public awareness of potential risks, stricter legal requirements, and the increased
complexity of modem industrial plants. Chemical engineers have a special role to perform in
assuring process safety. As Turton et al. (1998) have noted, "As the professional with the best
knowledge of the risks of a chemical processing operation, the chemical engineer has a
responsibility to communicate those risks to employers, employees, clients, and the general
public." Furthermore, in the AIChE Code of Ethics, the first responsibility of chemical engineers
is to "Hold paramount the safety, health, and welfare of the public in performance of their
professional duties." Professional societies, including the American Institute of Chemical
Engineers (AIChE), the Institution of Chemical Engineers (London), and the Instrument Society
of America (ISA), have played a leading role in developing safety standards and reference
materials. For example, the AIChE Center for Chemical Process Safety (CCPS) has published a
number of books (AIChE, 1993 a, b) and a journal devoted to safety, Process Safety Progress.
The overall safety record of the process industries has been quite good. However, it is not
possible to eliminate risk entirely, and serious accidents occasionally do occur (Kletz, 1985; Lees,
1986; Crowl and Louvar, 1990; Banerjee, 2003).

10.5.1 Overview of Process Safety


Process safety is considered at various stages in the lifetime of a process:
1. An initial safety analysis is performed during the preliminary process design.
2. A very thorough safety review is conducted during the final stage of the process
design using techniques such as hazard and operability (HAZOP) studies, failure
mode and effect analysis, and fault tree analysis (AIChE, 1993b; Crowl and Louvar,
1990).
3. After plant operation begins, HAZOP studies are conducted on a periodic basis in
order to identify and eliminate potential hazards.
4. Many companies require that any proposed plant change or change in operating
conditions require formal approval via a Management of Change process that
considers the potential impact of the change on the safety, environment, and health
of the workers and the nearby communities. Proposed changes may require
governmental approval, as occurs for the U.S. pharmaceutical industry, for
example.
5. After a serious accident or plant "incident," a thorough review is conducted to
determine its cause and to assess responsibility.
In modern chemical plants, process safety relies on the principle of multiple
protection layers (AIChE, 1993b; ISA, 1996). A typical configuration is shown in Figure
10.11. Each layer of protection consists of a grouping of equipment and/or human actions.
The protection layers are shown in the order of activation that occurs as a plant incident
develops. In the inner layer, the process design itself provides the first level of protection.
The next two layers consist of the basic process control system (BPCS) augmented with
two levels of alarms and operator supervision or intervention. An alarm indicates that a
measurement has exceeded its specified limits and may require operator action. The fourth
layer consists of a safety interlock system (SIS), which is also referred to as a safety
instrumented system or as an emergency shutdown (ESD) system. The SIS automatically
takes corrective action when the process and BPCS layers are unable to handle an
emergency. For example, the SIS could automatically turn off the reactant pumps after a
high temperature alarm occurs for a chemical reactor. Relief devices such as rupture discs
and relief valves provide physical protection by venting a gas or vapor if overpressurization
occurs. As a last resort, dikes are located around process units and storage tanks to contain
liquid spills. Emergency response plans are used to address emergency situations
and to inform the community. The functioning of the multiple layer protection system can
be summarized as follows (AIChE, 1993 b ): "Most failures in well-designed and operated
chemical processes are contained by the first one or two protection layers. The middle
levels guard against major releases and the outermost layers provide mitigation response
to very unlikely major events. For major hazard potential, even more layers may be
necessary."
It is evident from Fig. 10.11 that process control and automation play important
roles in ensuring process safety. In particular, many of the protection layers in Fig. 10.11
involve instrumentation and control equipment. Furthermore, the process and instrument
dynamics are key considerations in safety analysis. For example, after a major incident
develops, how much time elapses before it is reflected in the available measurements? If
the incident remains undetected, how long will it take for an emergency situation to result?

10.5.2 The Role of the Basic Process Control System


The basic process control system (BPCS) consists of feedback control loops that
regulate process variables such as temperatures, flow rates, liquid levels, and pressures.
Although the BPCS typically provides satisfactory control during routine process
operation, it may not do so during abnormal conditions. For example, if a controller output
saturates (reaches a maximum or minimum value), the controlled variable may exceed
allowable limits. Similarly, the failure or malfunction of a component in the feedback loop
such as a sensor, control valve, or transmission line could cause the process operation to
enter an unacceptable region. Typical component failure rates are shown in Table 10.1 for
1986 data.
10.5.3 Process Alarms
The second and third protection layers rely on process alarms to call attention to
abnormal situations. A block diagram for an alarm system is shown in Fig. 10.12. An alarm
is generated automatically when a measured variable exceeds specified high or low limits.
The logic block is programmed to take appropriate corrective action when one or more
alarm switches are triggered. After an alarm occurs, the logic block activates a final control
element or an annunciator, either a visual display or an audible sound such as a horn or
bell. For example, if a reactor temperature exceeds a high alarm limit, a light might flash
on a computer screen, with the color indicating the alarm priority (e.g., yellow for a less
serious situation, red for a critical situation). An alarm continues until it is acknowledged
by an operator action such as pressing a button or a key on a computer keyboard. If the
alarm indicates a potentially hazardous situation, then an automated corrective action is
initiated by the SIS. In computer control systems, two types of high- and low-alarm limits
are widely employed. Warning limits are used to denote minor excursions from nominal
values, while alarm limits indicate larger, more serious excursions.
Connell (1996) has proposed the following classification system for process alarms.
Type I Alarm: Equipment status alarm. Indicates equipment status, for example,
whether a pump is on or off or whether a motor is running or stopped.
Type 2 Alarm: Abnormal measurement alarm. Indicates that a measurement is
outside of specified limits.
Type 3 Alarm: An alarm switch without its own sensor. Directly activated by
the process rather than by a sensor signal. Type 3 alarms are used.for situations where it is
not necessary to know the actual value of the process variable, only whether it is above (or
below) a specified limit.
Figure 10.13 shows typical configurations for Type 2 and 3 alarms. In the Type 2
alarm system, the flow sensor/transmitter (FT) signal is sent to both a flow controller (FC)
and a flow switch (FSL refers to "flow-switch-low"). When the measurement is below the
specified low limit, the flow switch sends a signal to an alarm which activates an
annunciator in the control room (FAL refers to "flow-alarm-low").

By contrast, for the Type 3 alarm system in Fig. 10.13b, the flow switch is self-
actuated and thus does not require a signal from a flow sensor/transmitter.
Type 3 alarms are also used to indicate that an automatic shutdown system has
"tripped." They are widely employed on automobile instrument panels (Connell, 1996).
Type 4 Alarm. An alarm switch with its own sensor. Serves as a backup in case
the regular sensor fails.
Type 5 Alarm. Automatic Shutdown or Start-up System. Important and widely
used systems; described in the next section on Safety Interlock Systems.
It is tempting to specify tight alarm limits for a large number of process variables,
but one should resist this temptation because au excessive number of unnecessary alarms
could result. Furthermore, too many alarms can be as detrimental as too few alarms, for
several reasons. First, frequent "nuisance alarms" tend to make the plant operators less
responsive to important alarms. Second, in an actual emergency, a large number of
unimportant alarms tends to obscure the root cause of the problem. Third, the relationships
between alarms needs to be considered. Thus, the design of an appropriate alarm
management system is a challenging task (Connelly, 1997).

10.5.4 Safety Interlock Systems (SIS)


The SIS in Fig. 10.11 serves as an emergency backup system for the BPCS. The
SIS automatically starts when a critical process variable exceeds specified alam1 limits that
define the allowable operating region. Its initiation results in a drastic action such as
starting or stopping a pump or shutting down a process unit. Consequently, it is used only
as a last resort to prevent injury to people or equipment.
It is very important that the SIS function independently of the BPCS; otherwise,
emergency protection will be unavailable during periods when the BPCS is not operating
(e.g., due to a malfunction or power failure). Thus, the SIS should be physically separated
from the BPCS (AIChE, 1993b) and have its own sensors and actuators. Sometimes
redundant sensors and actuators are utilized (Englund and
Grinwis, 1992). For example, triply redundant sensors are used for critical measurements,
with SIS actions based on the median of the three measurements. This strategy prevents a
single sensor failure from crippling SIS operation. The SIS also has a separate set of alarms
so that the operator can be notified when the SIS initiates an action (e.g., turning on an
emergency cooling pump), even if the BPCS is not operational.

10.5.5 Interlocks and Automatic Shutdown Systems


The SIS operation is designed to provide automatic responses after alarms indicate
potentially hazardous situations. The objective is to have the process reach a safe condition.
The automatic responses are implemented via interlocks and automatic shutdown and start-
up systems. Distinctions are sometimes made between safety interlocks and process
interlocks; process interlocks are used for less critical situations to provide protection
against minor equipment damage and undesirable process conditions such as the
production of off-spec product (Perry and Green, 1997).
Two simple interlock systems are shown in Fig. 10.14. For the liquid storage
system, the liquid level must stay above a minimum value in order to avoid pump damage
such as cavitation. If the level drops below the specified limit, the low-level switch (LSL)
triggers both an alarm (LAL) and a solenoid (S), which acts as a relay and turns the pump
off. For the gas storage system in Fig. 10.14b, the solenoid-operated valve is normally
closed. But if the pressure of the hydrocarbon gas in the storage tank exceeds a specified
limit, the high-pressure switch (PSH} activates an alarm (PAH) and causes the valve to
open fully, thus reducing the pressure in the tank. For interlock and other safety systems, a
switch can be replaced by a transmitter if the measurement is required. Also, transmitter
tend to be more reliable. Another common interlock configuration is to place a solenoid
switch between a controller and a control valve. When an alarm is activated, the switch
trips and causes the air pressure in the control valve to be vented. Consequently, the control
valve reverts to either its fail-open or fail-close position. Interlocks have traditionally been
implemented as "hard-wired systems" that are independent of the control hardware. But for
most applications, software implementation of the interlock logic via a digital computer or
a programmable logic controller is a viable alternative. Programmable logic controllers
(PLCs) are considered in Chapter 22 and Appendix A.
If a potential emergency situation is very serious, the SIS system will automatically
shut down or start up equipment. For example, a pump would be turned off (or tripped) if
it overheats or loses lubricant pressure. Similarly, if an exothermic chemical reaction starts
to "run away," it may be possible to quickly add a quench material that stops the reaction.
For some emergency situations, the appropriate response is an automatic start-up of
equipment rather than an automatic shutdown. For example, a backup generator or a
cooling water pump can be started if the regular unit shuts down unexpectedly.
Although the SIS is essential for safe process operation, unnecessary plant
shutdowns and startups should be avoided for several reasons. First, they result in loss of
production and generate off-spec product during the subsequent plant start-up. Second, the
emergency shutdowns and start-ups for a process unit involve risks and may activate
additional safety systems that "trip" other process units. Thus, "nuisance shutdowns" can
create additional hazards. The use of redundant sensors can reduce unnecessary shutdowns.
In summary, the safety-related aspects of instrumentation and process control are
of paramount importance. Some recommended safety guidelines for process design and
process control are shown in Table 10.2. As Rinard (1990) has poignantly noted, "The
regulatory control system affect the size of your paycheck; the safety control system affects
whether or not you will be around to collect it."
SUMMARY
This chapter has provided a broad overview of important issues in control system design.
The process design itself is a major factor in determining how well the process can be controlled.
Consequently, it is highly recommended that process dynamics and control issues be considered
early in the process design. Ignoring these important considerations can result in a plant that is
very difficult to control, or even inoperable.
The control system design is strongly influenced by the control degrees of freedom that are
available, NFC. In most situations, NFC is simply the number of independent process variables that
can be manipulated. In general, NFC < NF where NF is the model degrees of freedom that was
introduced in Chapter 2. The selection of the controlled, manipulated, and measured variables is a
key step in the control system design. These choices should be based on the guidelines presented
in Section 10.4. Because process safety is of paramount concern in manufacturing plants, it is also
a primary concern in control system design. In fact, automation and process control play key roles
in ensuring safe process operation.
[Instrumentation and Control Systems 1E, Bolton]
CHAPTER 4 | CONTROL SYSTEMS

4.1 Introduction
The term automation is used to describe the automatic operation or control of a process. In
modem manufacturing there is an ever-increasing use of automation, e.g. automatically operating
machinery, perhaps in a production line with robots, which can be used to produce components
with virtually no human intervention. Also, in appliances around the home and in the office, there
is an ever-increasing use of automation. Automation involves carrying out operations in the
required sequence and controlling outputs to required values. The following are some of the key
historical points in the development of automation, the first three being concerned with
developments in the organization of manufacturing which permitted the development of automated
production:
1. Modern manufacturing began in England in the 18th century when the use of water wheels
and steam engines meant that it became more efficient to organize work to take place in
factories, rather than it occurring in the home of a multitude of small workshops. The
impetus was thus provided for the development of machinery.
2. The development of powered machinery in the early 1900s meant improved accuracy in
the production of components so that instead of making each individual component to fit a
particular product, components were fabricated in identical batches with an accuracy which
ensured that they could fit any one of a batch of a product. Think of the problem of a nut
and bolt if each nut has to be individually made so that it fitted the bolt and the advantages
that are gained by the accuracy of manufacturing nuts and bolts being high enough for any
of a batch of nuts to fit a bolt.
3. The idea of production lines followed from this with Henry Ford, in 1909, developing them
for the production of motor cars. In such a line, the production process is broken up into a
sequence of set tasks with the potential for automating tasks and so developing an
automated production line.
4. In the 1920s developments occurred in the theoretical principles of control systems and the
use of feedback for exercising control. A particular task of concern was the development
of control systems to steer ships and aircraft automatically.
5. In the 1940s, during the Second World War, developments occurred in the application of
control systems to military tasks, e.g. radar tracking and gun control.
6. The development of the analysis and design of feedback amplifiers, e.g. the paper by Bode
in 1945 on Network Analysis and Feedback Amplifier design, was instrumental in further
developing control system theory.
7. Numerical control was developed in 1952 whereby tool positioning was achieved by a
sequence of instructions provided by a program of punched paper tape, these directing the
motion of the motors driving the axes of the machine tool. There was no feedback of
positional data in these early control systems to indicate whether the tool was in the correct
position, the system being open-loop control.
8. The invention of the transistor in 1948 in the United States led to the development of
integrated circuits, and, in the 1970s, microprocessors and computers which enabled
control systems to be developed which were cheap and able to be used to control a wide
range of processes. As a consequence, automation has spread to common everyday
processes such as the domestic washing machine and the automatic focusing, automatic
exposure, camera.
The automatic control of machines and processes is now a vital part of modem industry.
The benefits of such control systems include greater consistency of product, reduced operating
costs due to improved utilization of plant and materials and a reduction in manpower, and greater
safety for operating personnel.
This chapter is an introduction to the basic idea of a control system and the elements used.

4.2 Control Systems


As an illustration of what control systems can do, consider the following:
Control a variable to obtain the required value
1. You set the required temperature for a room by setting to the required temperature
the room thermostat of a central heating system. This is an example of a control
system with the variable being controlled being the room temperature.
2. In a bottling plant the bottles are automatically filled to the required level. The
variable being controlled is the liquid level in a bottle and control is exercised to
ensure no difference between the required level and that which occurs.
3. A computer-numerical-control (CNC) machine tool is used to automatically
machine a workpiece to the required shape, the control system ensuring that there
is no difference between the required dimensions and that which occurs.
4. Packets of biscuits moving along a conveyor belt have their weights checked and
those that are below the required minimum weight limit are automatically rejected.
Control is being exercised over the weight.
Control the sequence of events
5. A belt is used to feed blanks to a pressing machine. As a blank reaches the machine,
the belt is stopped, the blank positioned in the machine, the press activated to press
the required shape, then the pressed item is ejected from the machine and the entire
process repeated. A sequence of operations is being controlled with some
operations controlled to occur only if certain conditions are met, e.g. activation of
the press if there is a blank in place.
6. You set the dials on the automatic clothes washing machine to indicate that 'whites'
are being washed and the machine then goes through the complete washing cycle
appropriate to that type of clothing. This is an example of a control system where a
controlled sequence of events occurs.
Control whether an event occurs or not
7. The automatic clothes washing machine has a safety lock on the door so that the
machine will not operate if the power is off and the door open. The control is of the
condition which allows the machine to operate.

A control system can be thought of as a system which for some particular input or inputs
is used to control its output to some particular value (Figure 4.1(a)), give a particular sequence of
events (Figure 4.1(b)) or give an event if certain conditions are met (Figure 4.1(c)).
As an example of the type of control system described by Figure 4.1(a), a central heating
control system has as its input the temperature required in the house and as its output the house at
that temperature (Figure 4.2). The required temperature is set on the thermostat and the control
system adjusts the heating furnace to produce that temperature.
The control system is used to control a variable to some set value.
As an example of the type of control system described by Figure 4.1(b), clothes washing
machine has as its input a set of instructions as to the sequence of events required to wash the
clothes, e.g. fill the drum with cold water, heat the water to 40''C, tumble the clothes for a period
of time, empty the drum of water, etc. The manufacturers of the machine have arranged a number
of possible sequences which are selected by pressing a button or rotating a dial to select the
appropriate sequence for the type of wash required. Thus, the input is the information determining
the required sequence, and the output is the required sequence of events (Figure 4.3). The control
system is used to control a sequence of events.

4.2.1 Open- and Closed-Loop Control


Consider two alternative ways of heating a room to some required temperature. In
the first instance there is an electric fire which has a selection switch which allows a 1-kW
or a 2-kW heating element to be selected. The decision might be made that to obtain the
required temperature it is only necessary to switch on the 1-kW element. The room will
heat up and reach a temperature which is determined by the fact the 1-kW element is
switched on. The temperature of the room is thus controlled by an initial decision and no
further adjustments are made. This is an example of open-loop control. Figure 4.4
illustrates this. If there are changes in the conditions, perhaps someone opening a window,
no adjustments are made to the heat output from the fire to compensate for the change.
There is no information fed back to the fire to adjust it and maintain a constant temperature.

Now consider the electric fire heating system with a difference. To obtain the
required temperature, a person stands in the room with a thermometer and switches the 1
kW and 2 kW elements on or off, according to the difference between the actual room
temperature and the required temperature in order to maintain the temperature of the room
at the required temperature. There is a constant comparison of the actual and required
temperatures. In this situation there is feedback, information being fed back from the output
to modify the input to the system. Thus, if a window is opened and there is a sudden cold
blast of air, the feedback signal changes because the room temperature changes and so is
fed back to modify the input to the system. This type of system is called closed-loop. The
input to the heating process depends on the deviation of the actual temperature fed back
from the output of the system from the required temperature initially set, the difference
between them being determined by a comparison element. In this example, the person with
the thermometer is the comparison element. Figure 4.5 illustrates this type of system.

Fig. 4.5. The electric fire closed-loop system


Note that the comparison element in the closed-loop control system is represented
by a circular symbol with a + opposite the set value input and a - opposite the feedback
signal. The circle represents a summing unit and what we have is the sum
+ set value – feedback value = error
This difference between the set value and feedback value, the so-called error, is the
signal used to control the process. If there is a difference between the signals then the actual
output is not the same as the desired output. When the actual output is the same as the
required output then there is zero error. Because the feedback signal is subtracted from the
set value signal, the system is said to have negative feedback.

Consider an example of a ball valve in a cistern used to control the height of the
water (Figure 4.6). The set value for the height of the water in the cistern is determined by
the initial setting of the pivot point of the lever and ball float to cut the water off in the
valve. When the water level is below that required, the ball moves to a lower level and so
the lever opens the valve to allow water into the tank. When the level is at the required
level the ball moves the lever to a position which operates the valve to cut off the flow of
water into the cistern. Figure 4.7 shows the system when represented as a block diagram.

In an open-loop control system the output from the system has no effect on the input
signal to the plant or process. The output is determined solely by the initial setting. In a
closed-loop control system the output does have an effect on the input signal, modifying it
to maintain an output signal at the required value.

Open-loop systems have the advantage of being relatively simple and consequently
cheap with generally good reliability. However, they are often inaccurate since there is no
correction for errors in the output which might result from extraneous disturbances. Closed-
loop systems have the advantage of being relatively accurate in matching the actual to the
required values. They are, however, more complex and so more costly with a greater
chance of breakdown as a consequence of the greater number of components.

4.3 Basic Elements


Figure 4.8 shows the basic elements of an open-loop control system. The system has three
basic elements: control, correction and the process of which a variable is being controlled.

1. Control element. This determines the action to be taken as a result of the input of
the required value signal to the system.
2. Correction element. This has an input from the controller and gives an output of
some action designed to change the variable being controlled.
3. Process. This is the process of which a variable is being controlled.
There is no changing of the control action to account for any disturbances which change
the output variable.

4.3.1 Basic Elements of a Closed-Loop System


Figure 4.9 shows the general form of a basic closed-loop system.
The following are the functions of the constituent elements:
1.Comparison element
This element compares the required value of the variable being
controlled with the measured value of what is being achieved and produces an
error signal:
error = required value signal - measured actual value signal
Thus, if the output is the required value then there is no error and so no
signal is fed to initiate control. Only when there is a difference between the
required value and the actual values of the variable will there be an error signal
and so control action initiated.
2.Control law implementation element
The control law element determines what action to take when an error
signal is received. The control law used by the element may be just to supply a
signal which switches on or off when there is an error, as in a room thermostat,
or perhaps a signal which is proportional to the size of the error so that if the
error is small a small control signal is produced and if the error is large, a large
proportional control signal is produced. Other control laws include integral
mode where the control signal continues to increase as long as there is an error
and derivative mode where the control signal is proportional to the rate at which
the error is changing.
The term control unit or controller is often used for the combination of
the comparison element, i.e. the error detector, and the control law
implementation element. An example of such an element is a differential
amplifier which has two inputs, one the set value and one the feedback signal,
and any difference between the two is amplified to give the error signal. When
there is no difference there is no resulting error signal.
3.Correction element
The correction element or, as it is often called, the final control element,
produces a change in the process which aims to correct or change the controlled
condition. The term actuator is used for the element of a correction unit that
provides the power to carry out the control action. Examples of correction
elements are directional control valves which are used to switch the direction
of flow of a fluid and so control the movement of an actuator such as the
movement of a piston in a cylinder. Another example is an electric motor where
a signal is used to control the speed of rotation of the motor shaft.
4.Process
The process is the system in which there is a variable that is being
controlled, e.g. it might be a room in a house with the variable of its temperature
being controlled.
5.Measurement element
The measurement element produces a signal related to the variable
condition of the process that is being controlled. For example, it might be a
temperature sensor with suitable signal processing.
The following are terms used to describe the various paths through the system taken
by signals:
1. Feedback path
Feedback is a means whereby a signal related to the actual condition
being achieved is fed back to modify the input signal to a process. The feedback
is said to be negative when the signal which is fed back subtracts from the input
value. It is negative feedback that is required to control a system. Positive
feedback occurs when the signal fed back adds to the input signal.
2. Forward path
The term forward path is used for tlie path from tlie error signal to the
output. In Figure 4.9 these forward path elements are the control law element,
the correction element and the process element.
The term process control is often used to describe the control of variables, e.g.
liquid level or the flow of fluids, associated with a process in order to maintain them at
some value. Note also that the term regulator is sometimes used for a control system for
maintaining a plant output constant in the presence of external disturbances. Hence the
term regulator is sometimes applied to the correction unit.

4.4 Case Studies


The following are examples of closed-loop control systems to illustrate how, despite the
different forms of control being exercised, the systems all have the same basic structural elements.
4.4.1 Control of the speed of rotation of a motor shaft
Consider the motor system shown in Figure 4.10 for the control of the speed of
rotation of the motor shaft and its block diagram representation in Figure 4.11.

The input of the required speed value is by means of the setting of the position of
the movable contact of the potentiometer. This determines what voltage is supplied to the
comparison element, i.e. the differential amplifier, as indicative of the required speed of
rotation. The differential amplifier produces an amplified output which is proportional to
the difference between its two inputs. When there is no difference then the output is zero.
The differential amplifier is thus used to both compare and implement the control law. The
resulting control signal is then fed to a motor which adjusts the speed of the rotating shaft
according to the size of the control signal. The speed of the rotating shaft is measured using
a tachogenerator, this being connected to the rotating shaft by means of a pair of bevel
gears. The signal from the tachogenerator gives the feedback signal which is then fed back
to the differential amplifier.

4.4.2 Control of the position of a tool


Figure 4.12 shows a position control system using a belt driven by a stepper motor
to control the position of a tool and Figure 4.13 its block diagram representation.

The inputs to the controller are the required position voltage and a voltage giving a
measure of the position of the workpiece, this being provided by a potentiometer being
used as a position sensor. Because a microprocessor is used as the controller, these signals
have to be processed to be digital. The output from the controller is an electrical signal
which depends on the error between the required and actual positions and is used, via a
drive unit, to operate a stepper motor. Input to the stepper motor causes it to rotate its shaft
in steps, so rotating the belt and moving the tool.

4.4.3 Power steering


Control systems are used to not only maintain some variable constant at a required
value but also to control a variable so that it follows the changes required by a variable
input signal. An example of such a control system is the power steering system used with
a car. This comes into operation whenever the resistance to turning the steering wheel
exceeds a predetermined amount and enables the movement of the wheels to follow the
dictates of the angular motion of the steering wheel. The input to the system is the angular
position of the steering wheel. This mechanical signal is scaled down by gearing and has
subtracted from it a feedback signal representing the actual position of the wheels. This
feedback is via a mechanical linkage. Thus, when the steering wheel is rotated and there is
a difference between its position and the required position of the wheels, there is an error
signal. The error signal is used to operate a hydraulic valve and so provide a hydraulic
signal to operate a cylinder. The output from the cylinder is then used, via a linkage, to
change the position of the wheels. Figure 4.14 shows a block diagram of the system.

4.4.4 Control of fuel pressure


The modem car involves many control systems. For example, there is the engine
management system aimed at controlling the amount of fuel injected into each cylinder and
the time at which to fire the spark for ignition. Part of such a system is concerned with
delivering a constant pressure of ftiel to the ignition system. Figure 4.15(a) shows the
elements involved in such a system. The fuel from the fuel tank is pumped through a filter
to tlie injectors, the pressure in the fuel line being controlled to be 2.5 bar (2.5 x 0.1 MPa)
above the manifold pressure by a regulator valve. Figure 4.15(b) shows tlie principles of
such a valve. It consists of a diaphragm which presses a ball plug into the flow path of the
fuel. The diaphragm has the fuel pressure acting on one side of it and on the other side is
the manifold pressure and a spring. If the pressure is too high, the diaphragm moves and
opens up the return path to the fuel tank for the excess fuel, so adjusting the fuel pressure
to bring it back to the required value.
The pressure control system can be considered to be represented by the closed-loop
system shown in Figure 4.16. The set value for the pressure is determined by the spring
tension. The comparator and control law are given by the diaphragm and spring. The
correction element is the ball in its seating, and the measurement is given by the diaphragm.

4.4.5 Antilock brakes


Another example of a control system used with a car is the antilock brake
system (ABS). If one or more of the vehicle's wheels lock, i.e. begins to skid, during
braking, then braking distance increases, steering control is lost and tyre wear
increases. Antilock brakes are designed to eliminate such locking. The system is
essentially a control system which adjusts the pressure applied to the brakes so that
locking does not occur. This requires continuous monitoring of the wheels and
adjustments to the pressure to ensure that, under the conditions prevailing, locking
does not occur. Figure 4.17 shows the principles of such a system.

The two valves used to control the pressure are solenoid-operated directional-
control valves, generally both valves being combined in a component termed the
modulator. When the driver presses the brake pedal, a piston moves in a master cylinder
and pressurizes the hydraulic fluid. This pressure causes the brake caliper to operate and
the brakes to be applied. The speed of the wheel is monitored by means of a sensor. When
the wheel locks, its speed changes abruptly and so the feedback signal from the sensor
changes. This feedback signal is fed into the controller where it is compared with what
signal might be expected on the basis of data stored in the controller memory. The
controller can then supply output signals which operate the valves and so adjust the
pressure applied to the brake.

4.4.6 Thickness control


As an illustration of a process control system. Figure 4.18 shows the type of system
that might be used to control the thickness of sheet produced by rollers. Figure 4.19
showing the block diagram description.
The thickness of the sheet is monitored by a sensor such as a linear variable
differential transformer (LVDT). The position of the LVDT probe is set so that when the
required thickness sheet is produced, there is no output from the LVDT. The LVDT
produces an alternating current output, the amplitude of which is proportional to the error.
This is then converted to a DC error signal which is fed to an amplifier. The amplified
signal is then used to control the speed of a DC motor, generally being used to vary the
armature current. The rotation of the shaft of the motor is likely to be geared down and
then used to rotate a screw which alters the position of the upper roll, hence changing the
thickness of the sheet produced.

4.4.7 Control of liquid level


Figure 4.20 shows a control system used to control the level of liquid in a tank using
a float-operated pneumatic controller. Figure 4.21 showing a block diagram of the system.

When the level of the liquid in the tank is at the required level and the inflow and
outflows are equal, then the controller valves are both closed. If there is a decrease in the
outflow of liquid from the tank, the level rises and so the float rises. This causes point P to
move upwards. When this happens, the valve connected to the air supply opens and the air
pressure in the system increases. This causes a downward movement of the diaphragm in
the flow control valve and hence a downward movement of the valve stem and the valve
plug. This then results in the inflow of liquid into the tank being reduced. The increase in
the air pressure in the controller chamber causes tlie bellows to become compressed and
move that end of the linkage downwards. This eventually closes off the valve so that the
flow control valve is held at the new pressure and hence the new flow rate.
If there is an increase in the outflow of liquid from the tank, the level falls and so
the float falls. This causes point P to move downwards. When this happens, the valve
connected to the vent opens and the air pressure in the system decreases. This causes an
upward movement of the diaphragm in the flow control valve and hence an upward
movement of the valve stem and the valve plug. This then results in the inflow of liquid
into the tank being increased. The bellows react to this new air pressure by moving its end
of the linkage, eventually closing off the exhaust and so holding the air pressure at the new
value and the flow control valve at its new flow rate setting.
4.4.8 Robot gripper

The term robot is used for a machine which is a reprogrammable multi-function


manipulator designed to move tools, parts, materials, etc. through variable programmed
motions in order to carry out specified tasks. Here just one aspect will be considered, the
gripper used by a robot at the end of its arm to grip objects. A common form of gripper is
a device which has 'fingers' or 'jaws. The gripping action then involves these clamping on
the object. Figure 4.22 shows one form such a gripper can take if two gripper fingers are
to close on a parallel sided object. When the input rod moves towards the fingers they pivot
about their pivots and move closer together. When the rod moves outwards, the fingers
move further apart. Such motion needs to be controlled so that the grip exerted by the
fingers on an object is just sufficient to grip it, too little grip and the object will fall out of
the grasp of the gripper and too great might result in the object being crushed or otherwise
deformed. Thus, there needs to be feedback of the forces involved at contact between the
gripper and the object. Figure 4.23 shows the type of closed-loop control system involved.

The drive system used to operate the gripper can be electrical, pneumatic or
hydraulic. Pneumatic drives are very widely used for grippers because they are cheap to
install, the system is easily maintained and the air supply is easily linked to the gripper.
Where larger loads are involved, hydraulic drives can be used. Sensors that might be used
for measurement of the forces involved are piezoelectric sensors or strain gauges. Thus,
when strain gauges are stuck to the surface of the gripper and forces applied to a gripper,
the strain gauges will be subject to strain and give a resistance change related to the forces
experienced by the gripper when in contact with the object being picked up.
The robot arm with gripper is also likely to have further control loops to indicate
when it is in the right position to grip an object. Thus, the gripper might have a control loop
to indicate when it is in contact with the object being picked up; the gripper can then be
actuated and the force control system can come into operation to control the grasp. The
sensor used for such a control loop might be a microswitch which is actuated by a lever,
roller or probe coming into contact with the object.

4.4.9 Machine Tool Control


Machine tool control systems are used to control the position of a tool or workpiece
and the operation of the tool during a machining operation. Figure 4.24 shows a block
diagram of the basic elements of a closed-loop system involving the continuous monitoring
of the movement and position of the work tables on which tools are mounted while the
workpiece is being machined.
The amount and direction of movement required in order to produce the required
size and form of workpiece is the input to the system, this being a program of instructions
fed into a memory which then supplies the information as required. The sequence of steps
involved is then:
1. An input signal is fed from the memory store.
2. The error between this input and the actual movement and position of the work
table is the error signal which is used to apply the correction. This may be an
electric motor to control the movement of the work table. The work table then
moves to reduce the error so that the actual position equals the required position.
3. The next input signal is fed from the memory store.
4. Step 2 is then repeated.
5. The next input signal is fed from the memory store and so on.

4.4.10 Fluid Flow Control


Figure 4.25 shows the elements of a control system used to control the rate of flow
of liquid to some required value, regardless of any fluctuations in supply pressure or back
pressure. Figure 4.26 shows a block diagram of the system.

4.5 Discrete-Time Control Systems


Discrete-time control systems are control systems in which one or more inputs can change
only at discrete instants of time, i.e. the inputs are effectively on-off signals and so in digital form
rather than the analogue form which has been discussed earlier in this chapter. This form of control
is often called sequential control. It describes control systems involving logic control functions,
e.g. is there or is there not a signal from sensor A or perhaps an AND logic system where the issue
is whether there is an input from sensor A and an input from sensor B, in order to determine
whether to give an output and so switch some device on or off.

Discrete-time control systems are control systems in which one or more inputs can change
only at discrete instants of time and involve logic control functions.

As a simple illustration of sequential control, consider the automatic kettle. When the kettle
is switched on, the water heats up and continues heating until a sensor indicates that boiling is
occurring. The sensor is just giving an on-off signal. The kettle then automatically switches off.
The heating element of the kettle is not continuously controlled but only given start and stop
signals.
As another example, the filling of a container with water might have a sensor at the bottom
which registers when the container is empty and so gives an input to the controller to switch the
water flow on and a sensor at the top which registers when the container is full and so gives an
input to the controller to switch off the flow of water (Figure 4.27). We have two sensors giving
on-off signals in order to obtain the required sequence of events.
As an illustration of the type of control that might be used with a machine consider the
system for a drill which is required to automatically drill a hole in a workpiece when it is placed
on the work table (Figure 4.28). A switch sensor can be used to detect when the workpiece is on
the work table, such a sensor being an on-off sensor. This then gives an on-input signal to the
controller and it then gives an output signal to actuate a motor to lower the drill head and start
drilling.
When the drill reaches the full extent of its movement in the workpiece, the drill head
triggers another switch sensor. This provides an on input to the controller and it then reverse the
direction of rotation of the drill head motor and the drill retracts.
INTRODUCTION
TO
MODELLING
Theoretical vs. Empirical Process Models
*YT Video

Two Types of Process Models:

Empirical Process Model


• are developed through experimenting on the actual system you’re studying (or something
physically close to the actual system).
• Not all processes can be experimented on

Theoretical Process Model


• Strictly rely on the 1st principles and physical laws (conservation of mass, momentum, and
energy)
• The math using theoretical models can quickly becomes very complex.
Steady State Model and Dynamic Model
*YT Video

Steady State Model


• Describes the process outputs under stationary conditions of input variables

Dynamic Model
• Describes the output response when there is change in one or more inputs.

Use of developing dynamic model?


• To exercise control system design
• Diagnose operational fault in plant

Example to differentiate SSM and DM


Consider: Tee Joint

Q1 Q2
X1 TEE JOINT X2

Q3
X3

Assumptions: Negligible time and volume

Under steady state:


Mass balance:
Q1+Q2=Q3 (Eq. 1)
Solute mass balance
Q1X1+ Q2X2= Q3 X3 (Eq.2)

Under dynamic state:


Eq. 1 and 2 are still valid.
Blending Process: Dynamic Modeling
*YT Video

𝑥 = mass fraction of species A


𝑤 = total mass flowrate leaving the tank
𝑉 = volume of the tank
𝜌 = density that is assumed as constant
Assumptions:
• Density is constant
• Liquid is incompressible
• Species A is dilute
𝑥1 = mass fraction of Stream 1
𝑤1 = flowrate of Stream 1
𝑥2 = mass fraction of Stream 2
𝑤2 = flowrate of Stream 2
CV = control variable
DV = disturbance variable
MV = manipulated variable

Total Mass:
𝑑
(𝑚) = 𝑤1 + 𝑤2 − 𝑤
𝑑𝑡
𝑑 𝑑
(𝜌𝑉) = 𝜌 (𝑉)
𝑑𝑡 𝑑𝑡
𝑑𝑉
𝜌 = 𝑤1 + 𝑤2 − 𝑤
𝑑𝑡
Mass of A:
𝑑
(𝜌𝑉𝑥) = 𝑤1 𝑥1 + 𝑤2 𝑥2 − 𝑤𝑥
𝑑𝑡
𝑑𝑉 𝑑𝑥
𝜌( 𝑥 + 𝑉)
𝑑𝑡 𝑑𝑡
𝑑𝑥 𝑑𝑉
𝜌𝑉 = 𝑤1 𝑥1 + 𝑤2 𝑥2 − 𝑤𝑥 − 𝜌𝑥
𝑑𝑡 𝑑𝑡

𝑑𝑉
𝜌 = 𝑤1 + 𝑤2 − 𝑤
𝑑𝑡
𝑑𝑥 1
𝜌𝑉 = 𝑤1 𝑥1 + 𝑤2 𝑥2 − 𝑤𝑥 − 𝜌𝑥 (𝑤1 + 𝑤2 − 𝑤)
𝑑𝑡 𝜌
𝑑𝑥 1
= [𝑤 (𝑥 − 𝑥) + 𝑤2 (𝑥2 − 𝑥)]
𝑑𝑡 𝜌𝑉 1 1
𝑑𝑉 1
= (𝑤 + 𝑤2 − 𝑤)
𝑑𝑡 𝜌 1
Blending Process: Steady State
*YT Video

𝑥 = mass fraction of species A


𝑤 = total mass flowrate leaving the tank
𝑉 = volume of the tank
𝜌 = density that is assumed as constant
𝑥1 = mass fraction of Stream 1
𝑤1 = flowrate of Stream 1
𝑥2 = mass fraction of Stream 2
𝑤2 = flowrate of Stream 2

Total Mass Balance:


𝑑𝑉 1
= (𝑤 + 𝑤2 − 𝑤)
𝑑𝑡 𝜌 1
𝑑𝑉 1
0= = (𝑤 + 𝑤2 − 𝑤)
𝑑𝑡 𝜌 1
𝑑𝑉 1
0= = (𝑤
̅ +𝑤 ̅2 − 𝑤
̅)
𝑑𝑡 𝜌 1

Mass Balance based on Species Balance A for the tank:


𝑑𝑥 1
= [𝑤 (𝑥 − 𝑥) + 𝑤2 (𝑥2 − 𝑥)]
𝑑𝑡 𝜌𝑉 1 1
𝑑𝑥 1
0= = [𝑤 (𝑥 − 𝑥) + 𝑤2 (𝑥2 − 𝑥)]
𝑑𝑡 𝜌𝑉 1 1
𝑑𝑥 1
0= = [𝑤
̅ (𝑥̅ − 𝑥̅ ) + 𝑤
̅ 2 (𝑥̅2 − 𝑥̅ )]
𝑑𝑡 𝜌𝑉̅ 1 1

𝑤
̅=𝑤 ̅1 + 𝑤 ̅2
̅1 (𝑥̅1 − 𝑥̅ ) + 𝑤
𝑤 ̅ 2 (𝑥̅2 − 𝑥̅ ) = 0
𝑤
̅1 𝑥̅ + 𝑤̅ 2 𝑥̅ = 𝑤̅1 𝑥̅1 + 𝑤 ̅ 2 𝑥̅2
𝑤̅1 𝑥̅1 + 𝑤̅ 2 𝑥̅2
𝑥̅ =
𝑤
̅1 + 𝑤 ̅2

Example: 𝑥̅1 = 0.1 𝑥̅2 = 0


̅1 = 1 𝑘𝑔⁄𝑠
𝑤 ̅ 2 = 3 𝑘𝑔⁄𝑠
𝑤

𝑤
̅=𝑤 ̅1 + 𝑤̅ 2 = 1 + 3 = 4 𝑘𝑔⁄𝑠
(1 𝑥 0.1) + (3 𝑥 0)
𝑥̅ = = 0.025
1+3
BLENDING PROCESS
LINEARIZATION EXAMPLE
*YT Video

We are going to consider the blending process, which is a nonlinear process and we are
going to calculate the approximate linear model for that process. We are going to consider the
dynamic equation for mass fraction x leaving the tank making an equation of:

In this case we are going to assume that our density () is constant and inlet mass fraction
(x₁ and x₂) are constant in this example. And now we are going to calculate the linearization of this
nonlinear dynamic equation. For the left hand side we are going to substitute the deviation
variables

Where:

𝑥̅ is the steady state value

X’ is the deviation from steady state

Then simplify

The approximate the right hand side by the first order of series of approximation
Then plug-in the function f and its derivative in this approximate linear model. First we
need to evaluate this function at a steady state value but the function is just the time derivative of
x and in this steady state we can plug in zero and the bar variables on the right hand side.

Then evaluate the function of f

Then use

The calculate the steady state value through this condition


REMEMBER:

This model is a linear approximate model and the approximation would only be good for
small deviations from steady state, small value of our deviation variables that’s because the first
order of series of approximation is going to be accurate for small deviation terms.
Introduction to Degrees of Freedom Analysis

*YT Video

Degree of freedom analysis

-
Done to find whether we have enough or too much information to solve a particular
problem
DOF = Number of unknowns – number of independent balances – number of other
equations

Where:

# of independent balances = # of species that are present in a particular system (mass or energy)

# of other equations= process specification (relationship or ratio of different flowrates), physical


property data (specific density, gravity, temperature or pressure)

First Scenario

2x + y = 7

- One equation
- Can’t solve because there are 2 unknowns
- Number of unknown > number of equation
- UNDERSPECIFIED SYSTEM
- DOF > 0
- Without further information we can’t solve all the unknown
Second Scenario

2x + y = 7

x+ 3y = 11

- Two equations
- Number of unknown = number of equations
- There is a zero degree of freedom
- Can solve the system of equation
- SOLVABLE
- DOF = 0
- We have the necessary equations to relate the unknowns that we have
Third Scenario

2x + y = 7

x+ 3y = 11

x+y=4

- Three equations
- Number of unknown < number of equations
- OVERSPECIFIED SYSTEM
- DOF < 0
- Can get answers of unknown but their answers are inconsistent or different
In analyzing engineering problems we are trying to find a system that has zero degree of
freedom and provides enough information so that we can solve for the unknown for the
particular problem.

EXAMPLE # 1

In the first example we have single unit process with two inputs and two outputs. If we
want to calculate the degree of freedom we need to know the number of unknown, number of
independent balances and other equations that we can write.

Unknown = 3

Independent material balances = 3 species

Other equation = 0
DOF = Number of unknowns – number of independent balances – number of other
equations

DOF= 3-3-0

DOF= 0 ( The system can be solved)

EXAMPLE #2

In the example we have single unit process that has one input and two outputs. Again, if
we want to calculate the degree of freedom we need to know the number of unknown, number of
independent balances and other equations that we can write.

Unknown = 5

Independent material balances = 3 species


Other equation = 2

DOF = Number of unknowns – number of independent balances – number of other


equations

DOF= 5-3-2

DOF= 0 ( The system can be solved)

You might also like