28.12.2014 Views

process instrumentation i module code: eipin1b study ... - Home Mweb

process instrumentation i module code: eipin1b study ... - Home Mweb

process instrumentation i module code: eipin1b study ... - Home Mweb

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

PROCESS INSTRUMENTATION I<br />

MODULE CODE: EIPIN1B<br />

STUDY PROGRAM: UNIT 2<br />

VUT<br />

Vaal University of Technology<br />

2/10


EIPINI Chapter 4: Level Measurement Page 4-1<br />

4. LEVEL MEASUREMENT<br />

This chapter aims to introduce students to some of the level measurement technologies<br />

used to measure the level of liquids and granulars in a container.<br />

4.1 INTRODUCTION<br />

On the 28th of March, 1979, thousands of people fled from Three Mile Island when<br />

the cooling system of a nuclear reactor failed. This dangerous situation developed<br />

because the level controls turned off the coolant flow to the reactor when they detected<br />

the presence of cooling water near the top of the tank. Unfortunately, the water<br />

reached the top of the reactor vessel not because there was too much water in the tank,<br />

but because there was so little that it boiled and swelled to the top. From this example,<br />

we can see that level measurement is more complex than simply the determination of<br />

the presence or absence of a fluid at a particular elevation.<br />

Most level measurement techniques can be categorized into one of two groups –<br />

direct and indirect (inferred) methods. Direct methods involve measuring the height of<br />

fluid directly as for example with a dipstick, overflow pipe, float or sight glass. With<br />

indirect methods, another variable is measured that correlate to the liquid level.<br />

Measuring the weight of a substance in a container or the pressure exerted on the<br />

bottom of a tank or transmitting an ultrasonic beam to the level surface and measuring<br />

the time of flight of the transmitted and received signal, is indirectly related to the<br />

level height.<br />

Level measurement may also be continuous or “point-level”. A continuous method<br />

refers to a technique whereby the device measures level on a constant basis, displaying<br />

or transmitting the actual level of the liquid as it changes. Point-level devices measure<br />

liquid at specific points within the tank. As the liquid level rises and falls, it passes<br />

through definite points during it's transit. Continuous methods may however be<br />

programmed to output alarms at specific points also, in addition to transmitting a<br />

continuous level measurement.<br />

Level measuring devices are also described as either a contact or non-contact type of<br />

instrument. A contact type device, such as a float, makes physical contact with the<br />

liquid in the container, in order to determine the level. A non-contact device, such as<br />

ultrasonic or radar, does not require contact with the material in the container to<br />

measure the level.<br />

4.2 THE SIGHT GLASS<br />

The sight glass consists of a strengthened glass tube, attached to the container as<br />

shown in Figure 4-1, through which the fluid level in the container can easily be<br />

observed by the operator. Monitoring the level from a distance, is facilitated when a<br />

magnetic float inside the tube, is allowed to rise and fall with the liquid level, causing<br />

metallic flags (or louvers) to flip and expose a different colour, indicating the level.<br />

These devices are available with continuous monitoring equipment, allowing both a<br />

local visible indication as well as an external signal for remote monitoring.


EIPINI Chapter 4: Level Measurement Page 4-2<br />

Valve<br />

Sight glass<br />

Magnetic<br />

float<br />

Metallic<br />

flaps<br />

Valve<br />

Container with sight glass<br />

Sight glass with magnetic indicator<br />

Figure 4-1<br />

4.3 FLOAT TYPE LEVEL INDICATORS<br />

4.3.1 Chain Float<br />

This type of float is linked to a<br />

rotating drum, by means of a<br />

chain, as shown in Figure 4-2.<br />

The chain engages a sprocket,<br />

which turns the drum, and with<br />

it the level indicator. A tape,<br />

that wraps around the drum, is<br />

also used, instead of a chain. A<br />

weight is attached to the other<br />

end of the chain or tape, to<br />

keep the chain pulled straight<br />

while the float moves up or<br />

down with the changing level.<br />

Level indicator<br />

Drum<br />

Seal<br />

Chain<br />

Weight<br />

Float<br />

Figure 4-2<br />

4.3.2 Ball Float<br />

The float is attached to a rod<br />

and rotary shaft, operating<br />

through a packing and bearing<br />

in the container wall, to the<br />

level indicator (Figure 4-3).<br />

Practical considerations, limit<br />

the shaft rotation to ± 30º from<br />

the horizontal, and therefore the<br />

range of the instrument as well.<br />

Figure 4-3<br />

Level<br />

indicator<br />

Rotary shaft<br />

Packing and<br />

bearing<br />

Float


EIPINI Chapter 4: Level Measurement Page 4-3<br />

4.3.3 Magnetic Float (Magnetic Coupled Float and Follower) Level Meter<br />

For this type of indicator, a dip<br />

pipe, made of non-magnetic<br />

material, is permanently installed in<br />

the container and sealed off, as<br />

shown in Figure 4-4. A doughnut<br />

shaped float with high strength<br />

magnet, is fitted around the dip<br />

pipe, to rise and fall vertically with<br />

the liquid level. A similar magnet<br />

that magnetically bonds to the<br />

outer magnet, is suspended within<br />

the dip pipe and attached to a<br />

rod, chain and pulley or similar<br />

arrangement, to the level indicator.<br />

This system is advantageous where<br />

leakages cannot be tolerated – such<br />

as with toxic, explosive or<br />

flammable liquids, as no packing<br />

or stuffing box is required and<br />

only the dip tube is required to<br />

withstand vessel pressure and<br />

temperature conditions.<br />

Figure 4-4<br />

Level indicator<br />

Indicator rod<br />

Non-magnetic<br />

dip tube<br />

Doughnut float<br />

with<br />

outer magnet<br />

Follower<br />

with<br />

inner magnet<br />

4.3.4 Magnetic Float Switch<br />

The magnetic float switch<br />

(Figure 4-5) is a point level device.<br />

When the level reaches a certain<br />

point, the float magnet activates<br />

the magnetic reed switch. The<br />

electric contacts are safely isolated<br />

from the inside (wet side) of<br />

the container by non-magnetic<br />

material that allows magnetic<br />

interaction between the float<br />

magnet and magnetic switch. The<br />

contacts may be used to switch a<br />

pump on or off, to sound an alarm<br />

or for other control purposes.<br />

Float<br />

Figure 4-5<br />

Float magnet<br />

Non-magnetic<br />

housing<br />

Magnetic<br />

reed switch<br />

Swivel pin


EIPINI Chapter 4: Level Measurement Page 4-4<br />

4.3.5 Flexure Tube Displacer (Torque tube) Level Meter<br />

The displacer type liquid level measuring instrument is not a float as such, for the<br />

displacer is heavier than the <strong>process</strong> fluid and the displacer moves very little during<br />

changes in tank level (a definite advantage over other float types). According to<br />

Archimedes’s law (illustrated in the figure below), the apparent weight of the<br />

displacer when immersed in a liquid, is its nominal weight in air minus the weight<br />

of the displaced liquid. The weight of the displacer will thus vary linearly from its<br />

weight in air (when the tank is empty) to its apparent weight when fully immersed<br />

in the liquid (when the tank is full). The weight of the displacer acting on the<br />

torque arm, will cause an angular displacement of the free end of the flexible torque<br />

tube and this movement will be transmitted to the outside world, by the torque rod.<br />

Torque tube<br />

Torque tube<br />

flange<br />

Level indicator<br />

Torque rod<br />

Archimedes’s Principle<br />

Torque arm<br />

Chain<br />

Displacer<br />

Figure 4-6<br />

Note: This instrument may also<br />

be fitted in a separate<br />

measuring chamber, fitted<br />

to the main vessel.<br />

Main vessel<br />

Measuring<br />

chamber with<br />

displacer, torque<br />

tube and measuring<br />

equipment<br />

Example 4-1<br />

During calibration of a displacer type level meter, it is found that the torque registered<br />

by the meter when the tank is empty, is 10 N-m. If the torque arm is 0.1 m, calculate<br />

the torque that will be generated when the displacer, with volume 0.002 m 3 , is fully<br />

immersed in the <strong>process</strong> fluid with density 1000 kg/m 3 .<br />

Remembering that torque is given by T = Fr, the weight of the displacer in air is<br />

10/0.1 = 100 N. Using Archimedes’ law, the apparent weight of the displacer when it<br />

is completely immersed in the liquid, is 100 – 0.002×1000×9.81 = 80.38 N. The torque<br />

generated then, will be 80.38×0.1 = 8.038 N-m.


EIPINI Chapter 4: Level Measurement Page 4-5<br />

4.4 DIFFERENTIAL PRESSURE MEASUREMENT OF LEVEL<br />

Perhaps the most frequently used technique of measuring level, is the method of<br />

measuring differential pressure. The primary benefit of this method is that the<br />

equipment measuring the differential pressure, can be externally installed or<br />

retrofitted to an existing vessel. It can also be isolated safely from the <strong>process</strong><br />

using block valves for maintenance and testing.<br />

There are, however, some disadvantages to this method. One vessel penetrations<br />

near the bottom of the vessel is needed, where leak paths could be the cause of<br />

many problems. Measurement errors also occur due to changes in liquid density.<br />

Density variations are caused by temperature changes or change of product. These<br />

variations must always be compensated for, to maintain accurate measurements.<br />

4.4.1 Open Containers<br />

To determine the fluid level H (Figure 4-7) in an open container, it is only necessary to<br />

measure the difference between the pressure at the bottom of the container and<br />

atmospheric pressure. In Figure 4-7 (a), the pressure difference is measured with a<br />

differential pressure transmitter while in Figure 4-7 (b), a u tube manometer is used.<br />

The pressure instruments may be installed with their zero lines level with the bottom<br />

of the container or at a distance z, below the bottom of the container.<br />

P atm<br />

P atm<br />

P atm<br />

H<br />

P atm<br />

H<br />

z<br />

HP<br />

LP<br />

DP<br />

transmitter<br />

Zero line<br />

z<br />

HP<br />

LP<br />

h<br />

U tube<br />

manometer<br />

Figure 4-7 (a)<br />

Figure 4-7 (b)


EIPINI Chapter 4: Level Measurement Page 4-6<br />

Example 4-2<br />

A mercury (density 13600<br />

P<br />

kg/m 3 atm P atm<br />

) u-tube manometer is<br />

used to measure the level of a<br />

liquid (density 1000 kg/m 3 ) in<br />

5 m<br />

a 5 meter high container, with 3 m<br />

the zero level of the<br />

manometer, exactly in line<br />

h Zero line<br />

with the bottom of the tank.<br />

Calculate the manometer<br />

X<br />

Y<br />

reading h, if the tank level is 3<br />

meter.<br />

Equating pressures on the XY line: U-tube<br />

P X = P Y<br />

∴P atm + 1000×(3 + ½h)×9.81 = P atm + 13600×h×9.81<br />

∴3 + ½h = 13.6h<br />

∴13.1h = 3<br />

∴h = 0.229 m<br />

= 229 mm<br />

Example 4-3<br />

A mercury (density 13600<br />

kg/m 3 ) well type manometer<br />

with A 2 /A 1 = 0.01 is used to<br />

measure the level of a liquid<br />

(density 1000 kg/m 3 ) in a 5<br />

meter high container, with the<br />

zero level of the manometer,<br />

exactly in line with the<br />

bottom of the tank. Calculate<br />

the manometer reading h if<br />

the tank level is 3 meter.<br />

5 m<br />

3 m<br />

We can approach this problem in three ways:<br />

a) Compare the pressures on the XY line, the<br />

same way as with the u tube manometer.<br />

b) Calculate P 1 and P 2 exactly on the well type<br />

manometer liquid meniscuses and utilise<br />

Equation 2-5 (Chapter 2).<br />

P ATM<br />

c) Ignore the level change (d) in the well (because it is so small), assume P 1 acts on<br />

the zero line and P 2 on the manometer liquid in the tube, and use Equation 2-5 to<br />

arrive at an approximate answer for h.<br />

Methods a) and b) will produce the exact value of h while method c) will give us an<br />

approximate value of h, although very nearly correct.<br />

X<br />

d<br />

P 1<br />

P 2 = P ATM<br />

h<br />

Zero line<br />

Y<br />

Well type manometer<br />

(Students must please take<br />

note of the orientation of<br />

the manometer. The well is<br />

always connected to the<br />

higher pressure.)


EIPINI Chapter 4: Level Measurement Page 4-7<br />

We will now use method a), b) and c) to solve this problem, but we will prefer the<br />

approximate but easier method c) for the remaining well type manometer problems<br />

and also for closed tanks involving well type manometers.<br />

a) Comparing pressures on the XY line:<br />

P ATM + 1000×3×9.81 + 1000×d×9.81 = P ATM + 13600×h×9.81 + 13600×d×9.81<br />

∴1000×3×9.81 + 1000×d×9.81 = 13600×h×9.81 + 13600×d×9.81<br />

∴29430 + 9810d = 133416h + 133416d<br />

∴29430 = 133416h + 123606d………………………………………………..(a)<br />

But from Equation (2), Chapter 2, Page 2-7: d = (A 2 /A 1 )h<br />

∴d = 0.01h……………………………………………………………………..(b)<br />

(b) in (a):<br />

29430 = 133416h + 123606×(0.01h)<br />

∴29430 = 133416h + 1236.06h<br />

∴134652.06h = 29430<br />

∴h = 0.21856 m<br />

b) P 1 = P ATM + 1000×3×9.81 + 1000×d×9.81<br />

= P ATM + 29430 + 9810×d = P ATM + 29430 + 9810×(0.01h)<br />

∴P 1 = P ATM + 29430 + 98.1h …………………………………….…………… (a)<br />

And P 2 = P ATM ………………………………………………………………… (b)<br />

From (a) and (b):<br />

P 1 – P 2 = 29430 + 98.1h ………………………………………………………. (c)<br />

But from Equation 2-5 (Chapter 2):<br />

P 1 – P 2 = ρhg(1 + A 2 /A 1 ) …………………………………...…………………. (d)<br />

(c) in (d):<br />

29430 + 98.1h = 13600×h×9.81(1 + 0.01)<br />

∴29430 + 98.1h = 133416×h×1.01<br />

∴29430 + 98.1h = 134750.16h<br />

∴134652.06h = 29430<br />

∴h = 0.21856 m<br />

c) P 1 = P ATM + 1000×3×9.81 (neglecting d)<br />

∴P 1 = P ATM + 29430 ………………………………………….…………….… (a)<br />

And P 2 = P ATM …………………………………………………….……...…… (b)<br />

From (a) and (b):<br />

P 1 – P 2 = 29430 ………………………………………………….……………. (c)<br />

And from Equation 2-5 (Chapter 2):<br />

P 1 – P 2 = ρhg(1 + A 2 /A 1 ) …………………………………...…………………. (d)<br />

(c) in (d):<br />

29430 = 13600×h×9.81(1 + 0.01)<br />

∴29430 = 134750.16h<br />

∴h = 0.2184 m


EIPINI Chapter 4: Level Measurement Page 4-8<br />

Example 4-4<br />

A DP transmitter must be calibrated to measure the level of a liquid in an open tank.<br />

The density of the liquid is 1000 kg/m 3 . The DP transmitter will be mounted one meter<br />

below the bottom of the tank. The tank is full when the height of the liquid in the tank<br />

is 5 meter and it is empty when there is only liquid in the high pressure line connected<br />

to the DP transmitter. Determine the necessary calibration specifications for an output<br />

signal of 4 to 20 mA.<br />

Empty:<br />

P 1 = P atm + ρhg<br />

= P atm + 1000×1×9.81<br />

= P atm + 9810<br />

∴P 1 – P 2 = 9810 Pa<br />

Full:<br />

P 1 = P atm + ρhg<br />

= P atm + 1000×6×9.81<br />

= P atm + 58860<br />

∴P 1 – P 2 = 58860 Pa<br />

Empty<br />

1 m<br />

P atm<br />

P atm<br />

P 1 P 2<br />

Full<br />

5 m<br />

1 m<br />

DP cell<br />

4 mA 20 mA<br />

P atm<br />

P 1 P 2<br />

DP cell<br />

Calibration specification: Output = 4 mA when input = 9810 Pa (empty condition)<br />

Output = 20 mA when input = 58860 Pa (full condition)<br />

Example 4-5<br />

If, in example 4-4, the DP cell is replaced by a<br />

u-tube manometer using mercury (density 13600<br />

kg/m 3 ), with its zero level 1 meter below the<br />

bottom of the container, calculate the full and<br />

empty readings, h empty and h full .<br />

Empty:<br />

Equating pressures on the XY line:<br />

P atm +1000×(1+½h)×9.81 = P atm +13600×h×9.81<br />

∴1 + ½h = 13.6h<br />

∴h empty = 0.07634 m = 76.34 mm.<br />

Full:<br />

Equating pressures on the XY line:<br />

P atm +1000×(6+½h)×9.81 = P atm +13600×h×9.81<br />

∴6 + ½h = 13.6h<br />

∴h full = 0.458 m. = 458 mm.<br />

Empty<br />

ZL<br />

Full<br />

5 m<br />

ZL<br />

1 m<br />

1 m<br />

P atm<br />

X<br />

P atm<br />

X<br />

P atm<br />

P atm<br />

h empty<br />

Y<br />

U tube<br />

P atm<br />

h full<br />

Y<br />

U tube


EIPINI Chapter 4: Level Measurement Page 4-9<br />

Example 4-6<br />

A DP transmitter, correctly calibrated to measure the level of<br />

a liquid in a 5 meter high open container, delivers an output<br />

pressure of 50 kPa. Calculate the liquid level in the tank.<br />

Liquid level H = [(50-20)/(100-20)]×5 = 1.875 m.<br />

Or alternatively: 50 = 16×H + 20 ⇒ H = 1.875 m<br />

Example 4-7<br />

A mercury (density 13600 kg/m 3 ) u tube manometer<br />

is used to measure the level of a liquid (density<br />

1000 kg/m 3 ) in a 5 meter high container, with the<br />

zero level of the manometer, 1 meter below the<br />

bottom of the tank. Calculate the level in the tank if<br />

the manometer reading is 300 mm (0.3 m).<br />

Comparing pressure on the XY line:<br />

P atm +1000×(H+1+0.15)×9.81=P atm +13600×0.3×9.81<br />

∴H + 1.15 = 13.6×0.3<br />

∴H = 4.08 – 1.15<br />

= 2.93 meter<br />

Example 4-8<br />

A mercury (density 13600 kg/m 3 ) well type<br />

manometer with A 2 /A 1 = 0.01, is used to measure<br />

the level of a liquid (density 1000 kg/m 3 ) in a<br />

5 meter high container, with the zero level of the<br />

manometer, 1 meter below the bottom of the tank.<br />

Calculate the level in the tank if the manometer<br />

reading is 200 mm (0.2 m).<br />

The input pressure, P 1 , on the well side of the<br />

manometer, with respect to the zero line, is:<br />

P 1 = P atm + 1000×(H + 1)×9.81<br />

and P 2 = P atm<br />

∴P 1 – P 2 = P atm + 1000×(H + 1)×9.81 – P atm<br />

= 1000×(H+1)×9.81 = 9810(H+1)<br />

Furthermore, from Equation 2-5 (Chapter 2):<br />

P 1 – P 2 = ρhg(1 + A 2 /A 1 ) = 13600×0.2×9.81×(1 + 0.01) = 2747×9.81 = 26948<br />

Therefore 9810(H+1) = 26948 ⇒ H+1 = 2.747<br />

∴H = 1.747 m.<br />

Note: If we did wish to include d, then d = (A 2 /A 1 )h = 0.01×0.2 = 0.002 m<br />

P 1 =P atm +1000×(H+1+d)×9.81=P atm +1000×(H+1+0.002)×9.81=9810(H+1.002)<br />

And P 1 – P 2 = 9810(H+1.002) = 26948 ⇒ H + 1.002 = 2.747<br />

∴H = 1.745 m<br />

H<br />

H<br />

ZL<br />

ZL<br />

100<br />

50<br />

20<br />

1 m<br />

d<br />

P o<br />

P o =16×Level+20<br />

0 H 5 Level<br />

1 m<br />

P atm<br />

X<br />

P atm<br />

P 1<br />

5 m<br />

5 m<br />

Well type<br />

manometer<br />

P atm<br />

0.3 m<br />

Y<br />

U tube<br />

P 2 =P atm<br />

0.2 m


4.4.2 Closed Containers<br />

EIPINI Chapter 4: Level Measurement Page 4-10<br />

For a closed container (Figure 4-8), the outer leg of the differential pressure<br />

measuring equipment, can not be open to atmosphere, as the tank may be<br />

pressurised and the tank pressure P t , could differ from atmospheric pressure. The<br />

outer leg is therefore brought back into the top of the vessel, to ensure that the<br />

pressure inside the tank and the outer leg pressure, are both equal to P t . Liquid<br />

vapour will typically condense in the outer leg and if an effort is made to keep this<br />

leg devoid of <strong>process</strong> fluid, it will result in what is called a dry leg system. If, on<br />

the other hand, the outer leg is allowed to be completely filled with the same<br />

<strong>process</strong> fluid as in the tank, a wet leg system will result. We will focus on wet leg<br />

closed tank systems.<br />

Differential pressure level measurement in a closed vessel (wet leg), is<br />

fundamentally different from level measurement in an open vessel. The high<br />

pressure input will always be connected to the outer leg while the low pressure<br />

input is obtained from the bottom of the tank. Another difference is that the<br />

differential pressure is zero when the tank is full, while the maximum differential<br />

pressure is obtained when the tank is empty. Students must ensure that they<br />

understand this critical difference between differential pressure measurement in an<br />

open vessel and a closed vessel, very clearly.<br />

Figure 4-8(a) illustrates a DP transmitter measuring the pressure difference between<br />

the outer leg and the tank. It may be connected in line with the bottom of the tank or<br />

below. In Figure 4-8 (b), a u tube manometer is used to measure the pressure<br />

difference. In the case where a manometers is used, it is clear from Figure 4-8 (b), that<br />

the zero line must be lower than the bottom of the tank.<br />

P t<br />

P t<br />

P t<br />

P t<br />

H<br />

L<br />

H<br />

L<br />

z<br />

LP<br />

HP<br />

DP transmitter<br />

z<br />

LP<br />

HP<br />

ZL<br />

h<br />

U tube<br />

manometer<br />

Figure 4-8 (a)<br />

Figure 4-8 (b)


EIPINI Chapter 4: Level Measurement Page 4-11<br />

Example 4-9<br />

A DP transmitter is used to measure the level of a liquid, with density 1000 kg/m 3 in a<br />

closed tank, that can store liquid to a maximum level of 5 meter. The DP transmitter is<br />

installed 1 meter below the bottom of the tank. Calculate the input differential pressure<br />

to the DP transmitter when the tank is empty and when it is full.<br />

P t<br />

P t<br />

P 1<br />

P t<br />

P t<br />

P 1<br />

5 m<br />

5 m<br />

P 2<br />

1 m 1 m<br />

Empty:<br />

DP transmitter Full: DP transmitter<br />

P 1 = P t + 1000×(5 + 1)×9.81 P 1 = P t + 1000×(5 + 1)×9.81<br />

P 2 = P t + 1000×1×9.81 P 2 = P t + 1000×(5 + 1)×9.81<br />

P 1 – P 2 = 49050 Pa = 49.05 kPa.<br />

P 1 – P 2 = 0 Pa.<br />

Example 4-10<br />

A u tube mercury (δ = 13.6) manometer is used to measure the level of a liquid (δ=1),<br />

in a closed tank, that can store liquid to a maximum level of 5 meter. The zero level of<br />

the manometer is 1 meter below the bottom of the tank. Calculate the manometer<br />

readings when the tank is empty and full.<br />

P 2<br />

P t<br />

P t<br />

5 m<br />

P t<br />

P t<br />

5 m<br />

u-tube<br />

1 m 1 m<br />

h empty ZL u-tube<br />

X<br />

Y<br />

X<br />

h full<br />

ZL<br />

Y<br />

Empty:<br />

Full:<br />

Equating pressures on the XY line: Equating pressures on the XY line:<br />

P t +1000×(1 - ½h)×9.81 + 13600×h×9.81 P t +1000×(5+1-½h)×9.81+13600×h×9.81<br />

= P t + 1000×(5 + 1 + ½h)×9.81 = P t + 1000×(5 + 1 + ½h)×9.81<br />

∴(1 - ½h) + 13.6h = 6 + ½h ∴(6 - ½h) + 13.6h = 6 + ½h<br />

∴13.6h – ½h - ½h = 6 – 1 ⇒ 12.6h = 5 ∴13.6h – ½h - ½h = 0 ⇒ 12.6h = 0<br />

∴h empty = 0.3968 m = 396.8 mm. ∴h full = 0 (as we knew all along)


Example 4-11<br />

The level of a liquid (density 1000<br />

kg/m 3 ) in a closed tank, 5 meter<br />

high, is measured with a mercury<br />

(density 13600 kg/m 3 ) well type<br />

manometer (A 2 /A 1 =0.01). The<br />

manometer is installed with its<br />

zero line 1 meter below the bottom<br />

of the tank. If the reading on the<br />

manometer is 0.2 m, calculate the<br />

level H of the fluid in the tank.<br />

Example 4-12<br />

The level of a liquid, with density ρ l , stored in a<br />

closed tank, L meters high, is measured with a u<br />

tube manometer, using a manometer liquid, with<br />

density ρ m (ρ m > ρ l ). The manometer is installed<br />

with its zero line a distance z meter below the<br />

bottom of the tank. When the fluid level in the<br />

container is H meter, the reading obtained from<br />

the manometer is h meter. Derive an equation<br />

that can be used to calculate the liquid level H, in<br />

terms of the manometer reading h and in terms of<br />

the constants ρ l , ρ m , L, and z.<br />

EIPINI Chapter 4: Level Measurement Page 4-12<br />

1 m P 2<br />

0.2 m<br />

P 1<br />

d=.002<br />

Zero line<br />

P 1 = P t + 1000×(5+1)×9.81 = P t + 58860<br />

X<br />

Y<br />

P 2 = P t + 1000×[H+(1-0.2)]×9.81 = P t + 9810×(H+0.8)<br />

∴P 1 – P 2 = 58860 – 9810(H+0.8) = 58860 – 9810H – 7848 = 51012 – 9810H…(1)<br />

And P 1 – P 2 = ρhg(1 + A 2 /A 1 ) = 13600×0.2×9.81×(1 + 0.01) = 26950 ………… (2)<br />

From (1) and (2): 51012 – 9810H = 26950 ⇒ 9810H = 51012 – 26950 = 24062<br />

∴H = 2.453 m.<br />

(The exact value of H may be obtained if we equate pressures on the XY line:<br />

P t +1000×[H+(1–0.2)]×9.81+13600×(0.2+0.002)×9.81=P t +1000×(5+1+0.002)×9.81<br />

∴[H+0.8] + 13.6×0.202 = 6.002 ⇒ H + 0.8 + 2.747 = 6.002 ⇒ H + 3.547 = 6.002<br />

∴H = 2.455 m.)<br />

Equating pressures at points X and Y:<br />

P t + ρ l ×(H + z - ½h)×g + ρ m ×h×g = P t + ρ l ×(L + z + ½h)×g<br />

∴ρ l ×(H + z - ½h)×g + ρ m ×h×g = ρ l ×(L + z + ½h)×g<br />

∴ρ l ×(H + z - ½h) + ρ m ×h = ρ l ×(L + z + ½h)<br />

∴ρ l H + ρ l z - ½ρ l h + ρ m h = ρ l L + ρ l z + ½ρ l h<br />

∴ρ l H = ρ l L + ρ l z + ½ρ l h - ρ l z + ½ρ l h - ρ m h = ρ l L + ρ l h - ρ m h<br />

ρ<br />

⎛ ρ<br />

∴H = L + h - m h = L + 1 h<br />

ρ<br />

ρ m ⎞<br />

⎜ ⎟<br />

⎜ − ⎟ (interestingly, z doesn’t play a role)<br />

l<br />

l<br />

⎝<br />

H<br />

⎠<br />

P t<br />

H<br />

ρ l<br />

5 m<br />

P t<br />

z<br />

X<br />

ρ m<br />

P t<br />

L<br />

Please note that<br />

the well (high<br />

pressure input)<br />

is connected to<br />

the outside tube<br />

h<br />

P t<br />

ZL<br />

Y


EIPINI Chapter 4: Level Measurement Page 4-13<br />

4.5 BUBBLE METERS<br />

A bubbler level meter (air purge system or gas flushing system), provides a simple<br />

and inexpensive but less accurate (±1-2%) level measurement system for virtually<br />

any liquid, and typically for corrosive or slurry-type applications. A constant flow<br />

of compressed air or an inert gas (usually nitrogen), is introduced through a dip<br />

tube. The pressure required to force bubbles through the tube at a constant rate<br />

(normally 60 bubbles per minute), is proportional to the hydrostatic pressure at the<br />

bottom end of the dip tube. A bubbler sight glass, facilitate with adjusting the<br />

bubble rate. The level is determined from the difference between the bubbler<br />

pressure (back pressure) and the pressure at the surface of the liquid, by means of a<br />

DP transmitter or manometer. Disadvantages of the bubbler system are 1) that the<br />

liquid’s density, influences measurement accuracy and 2) the need for compressed<br />

air or gas. The end of the dip tube may become plugged or clogged, and the dip<br />

tube is periodically purged (release of large amounts of air to clear the tube), to<br />

combat this situation. Figure 4-9 (a) and (b), depicts the measurement arrangement<br />

for an open tank and a closed tank respectively, using a u-tube manometer.<br />

Bubbler<br />

sight glass<br />

Pressure<br />

regulator<br />

Filter<br />

Air/gas<br />

supply<br />

Dip tube<br />

Figure 4-9 (a)<br />

Bubbler<br />

sight glass<br />

Pressure<br />

regulator<br />

Filter<br />

Air/gas<br />

supply<br />

Dip tube<br />

Figure 4-9 (b)


EIPINI Chapter 5: Temperature Measurement Page 5-1<br />

5. TEMPERATURE MEASUREMENT<br />

The purpose of this chapter is to introduce students to fundamental concepts related to<br />

temperature and to discuss important industrial methods used to measure temperature.<br />

5.1 INTRODUCTION<br />

Galileo Galilei (picture to the left), was born in the year<br />

1564 in Pisa, Italy. This brilliant scientist and astronomer is<br />

most famous for his early development of the telescope,<br />

being the first to see the moons of Jupiter and other celestial<br />

objects. Through his work in astronomy, Galileo supported<br />

the theory originated by Copernicus, that the earth moved<br />

around the Sun. Eventually, and under immense pressure<br />

from the Church, he publicly retracted his support for the<br />

Copernicus theory. Even so, Galileo was placed under house<br />

arrest and spent the last 10 years of his life in almost<br />

complete seclusion, having dared to offend the Church.<br />

Galileo was indeed an incredible man. He was the first to realise how the swinging<br />

of a pendulum, could be used to measure time and he is also credited to have<br />

developed the first device to indicate changes in temperature.<br />

This instrument did not measure temperature as such, and is<br />

therefore called a thermoscope. (Vincenzo Viviani states in<br />

his Vita di Galileo, that the thermoscope was conceived by<br />

Galileo in 1597.) Galileo’s thermoscope was based on the<br />

fascinating idea that air will expand when heated and contract<br />

when cooled. The instrument consisted of a glass bottle about<br />

the size of an egg, with a long glass neck. Air in the bottle<br />

was heated with the hands and the open end side of the tube<br />

immersed partially in a vessel containing water or wine. As<br />

the air in the tube cools, it contracts and the liquid is drawn<br />

into the tube. Once an equilibrium point is established, a rise<br />

Cold<br />

Hot<br />

in temperature increases the volume of the air, forcing fluid back down; a fall in<br />

temperature reduces the volume of the air, drawing more fluid into the tube.<br />

Early in the 1700’s, Gabriel Daniel Fahrenheit developed a<br />

more reliable temperature measuring device. A thin glass<br />

tube holds a volume of mercury in a bulb at the bottom.<br />

When heated, the mercury expands upward in the tube. The<br />

tube could be sealed and allowed for accurate, reproducible<br />

temperature readings. Because this device could be marked<br />

with a numerical scale, it was called a thermometer.<br />

Cold<br />

Hot


EIPINI Chapter 5: Temperature Measurement Page 5-2<br />

As the inventor of a reliable thermometer, Fahrenheit was afforded the luxury of<br />

developing a temperature scale to his liking. He defined a temperature scale on which<br />

0 was the coldest temperature he could reproduce in the lab (the temperature of an icesalt<br />

mixture) and 12 as his body temperature (as 12 inches in a foot). Fahrenheit soon<br />

discovered that 12 divisions between the two set points were not fine enough for good<br />

measurement so he doubled the divisions first to 24, then to 48 and finally to 96 (the<br />

average person’s body temperature is actually 98.6ºF). On the Fahrenheit temperature<br />

scale water freezes at 32º and boils at 212º. Indeed the unit, degrees, may arise from<br />

the fact that there are 180 increments between freezing and boiling of water.<br />

The Swedish astronomer, Anders Celsius, arbitrarily selected 100 to be the freezing<br />

point of water and 0 to be the boiling point of water. Later the end points were<br />

reversed and the centigrade scale was born (centigrade since there were 100<br />

increments between the freezing point and boiling point). In 1948 the name was<br />

officially changed from the centigrade scale to the Celsius scale.<br />

The Fahrenheit and Celsius scales are termed relative scales. An object can be colder<br />

than zero on the Fahrenheit scale or the Celsius scale and negative temperatures on<br />

both scales are common. Since temperature is related to the motion of atoms or<br />

molecules, zero should represent the absolute zero of temperature. Such a scale would<br />

be an absolute scale. Two absolute temperature scales do exist. The Kelvin scale<br />

(named after William Thomson, later Lord Kelvin) uses the same size increments as<br />

the Celsius scale, but has its zero at absolute zero temperature. The Rankine scale<br />

(named after William John Macquorn Rankine) is the absolute scale whose increments<br />

are the same size as those in the Fahrenheit scale.<br />

5.2 UNITS OF TEMPERATURE<br />

Definition of Temperature<br />

Temperature is defined as the degree of heat of a body. The SI unit for<br />

temperature is Kelvin (K).<br />

To set a temperature scale, two fixed temperature points must be defined. The distance<br />

between the two fixed points, is called the fundamental interval. The two fixed points<br />

used, are the following:<br />

Bottom fixed point: The temperature of ice (prepared from distilled water)<br />

mixed with distilled water, at standard atmospheric pressure of 760<br />

mm. mercury.<br />

Top fixed point: The temperature of distilled water that boils at standard<br />

atmospheric pressure of 760 mm. mercury.<br />

The Celsius scale<br />

The bottom fixed point is 0 degrees Celsius (0 °C).<br />

The top fixed point is 100 degrees Celsius (100 °C)<br />

The fundamental interval of the Celsius scale is thus divided into 100 even parts<br />

to express 1 °C.


The Fahrenheit scale<br />

EIPINI Chapter 5: Temperature Measurement Page 5-3<br />

The bottom fixed point is 32 degrees Fahrenheit (32 °F).<br />

The top fixed point is 212 degrees Fahrenheit (212 °F).<br />

The fundamental interval of the Fahrenheit scale is thus divided into 180 even<br />

parts to express 1 °F.<br />

The Kelvin scale<br />

Zero Kelvin (-273.15 °C) corresponds to the absolute zero of temperature. The<br />

melting point of ice is 273.15 Kelvin (K) and the boiling point of water is 373.15 K.<br />

A temperature change of 1 °C, corresponds to a temperature change of 1 K.<br />

The Rankine scale<br />

Zero Rankin (-459.67 °F) corresponds to the absolute zero of temperature. The<br />

melting point of ice is 491.67 Rankine (R) and the boiling point of water is 671.67 R.<br />

A temperature change of 1 °F, corresponds to a temperature change of 1 R.<br />

212 °F 672 R Water boils 100 °C 373 K<br />

68 °F 528 R Room temperature 20 °C 293 K<br />

32 °F 492 R Water freezes 0 °C 273 K<br />

-460 °F 0 R Absolute zero -273 °C 0 K<br />

Fahrenheit Rankine Celsius Kelvin<br />

Relationship between Fahrenheit and Celsius<br />

To determine the relationship between Fahrenheit and Celsius, we could draw a graph<br />

of F (temperature in °F) versus C (temperature in °C), for C in the range 0 ≤ C ≤ 100.<br />

Using the general equation for a<br />

straight line:<br />

y = m x + c,<br />

180<br />

F = C + 32 100<br />

∴F = 5<br />

9 C + 32 Equation 5-1<br />

We can now also express Celsius in<br />

terms of Fahrenheit.<br />

From Equation 5-1:<br />

9 C = F – 32<br />

5<br />

212<br />

F (°Fahrenheit)<br />

5<br />

∴C = (F – 32) Equation 5-2<br />

C<br />

0 100<br />

9 (°Celsius)<br />

32<br />

100<br />

180


EIPINI Chapter 5: Temperature Measurement Page 5-4<br />

Example 5-1<br />

Convert the absolute zero of temperature (–273.15 °C), to degrees Fahrenheit.<br />

F = (9/5)C + 32 = (9/5)×(-273.15) + 32 = -459.67 °F ≈ -460 °F<br />

Example 5-2<br />

Convert 86 °F into degrees Celsius, Rankine and Kelvin.<br />

C = (5/9)(F – 32) = (5/9)×(86 – 32) = 30 °C<br />

R = F + 460 = 86 + 460 = 546 R<br />

K = C + 273 = 30 + 273 = 303 K<br />

Example 5-3<br />

Determine the temperature in degrees Fahrenheit that would translate into the same<br />

value in degrees Celsius.<br />

In Equation 5-2 (or Equation 5-1), set C = F:<br />

∴F = (5/9)(F – 32) ⇒ 9F = 5F – 160 ⇒ 4F = -160 ⇒ F = -40 °F<br />

Example 5-4<br />

Convert 0 °F to degrees Celsius.<br />

C = (5/9)(F – 32) = (5/9)(0 – 32) = -17.78 °C.<br />

5.3 THE INTERNATIONAL TEMPERATURE SCALE<br />

This scale is based on a number of fixed temperature points and serves as calibration<br />

reference. All temperatures are measured at standard atmospheric pressure of 760 mm.<br />

mercury.<br />

• The oxygen point:<br />

• The ice point:<br />

• The boiling point:<br />

• The sulphur point:<br />

• The silver point:<br />

• The gold point:<br />

The boiling point of liquid oxygen: –182.97 °C.<br />

The melting point of pure ice: 0 °C.<br />

The boiling point of pure water: 100 °C.<br />

The boiling point of pure sulphur: 444.6 °C.<br />

The melting point of silver: 961.78 °C.<br />

The melting point of gold: 1064.18 °C.<br />

5.4 LIQUID IN GLASS THERMOMETERS<br />

The liquid in glass thermometer, is the most commonly used device to measure<br />

temperature and it is inexpensive to make and easy to use. The liquid in glass<br />

thermometer has a glass bulb attached to a sealed glass tube (also called the stem<br />

or capillary tube). A very thin opening, called a bore, exists from the bulb and<br />

extends down the centre of the tube. The bulb is typically filled with either<br />

mercury or red-coloured alcohol and is free to expand and rise up into the tube<br />

when the temperature increases, and to contract and move down the tube when<br />

the temperature decreases.


EIPINI Chapter 5: Temperature Measurement Page 5-5<br />

The background of the glass tube is covered with white enamel and the front of<br />

the glass tube forms a magnifying glass that enlarges the liquid column and<br />

facilitates with reading the temperature.<br />

In Figure 5-1 (a), an all glass thermometer is depicted, with its scale etched<br />

into the stem. Liquid in glass thermometers are fragile and for industrial use, the<br />

thermometer is mounted in a protective housing and the scale is engraved on a<br />

separate plate that is part of the protective case. An industrial thermometer is<br />

shown in Figure 5-1 (b).<br />

Protective case<br />

Scale<br />

(etched)<br />

Scale<br />

(on plate)<br />

Bore<br />

Lens front capillary<br />

Tube (stem)<br />

Bore<br />

Lens front capillary<br />

tube (stem)<br />

Liquid column<br />

Liquid column<br />

Socket<br />

Bulb<br />

Bulb chamber<br />

Figure 5-1 (a)<br />

Figure 5-1 (b)<br />

Bulb<br />

Heat conducting<br />

medium<br />

Liquids used in glass thermometers<br />

Liquid<br />

Temperature range (Celsius)<br />

Mercury -35 to +510<br />

Alcohol -80 to +70<br />

Pentane -200 to +30<br />

Toluene -80 to +100<br />

Creosote -5 to +200


EIPINI Chapter 5: Temperature Measurement Page 5-6<br />

5.5 MERCURY IN STEEL THERMOMETERS<br />

The mercury in steel thermometer<br />

system, allows for rugged construction<br />

and is used extensively in industrial<br />

applications. The thermometer consists<br />

of a steel bulb, a steel capillary tube<br />

and Bourdon tube, as shown in Figure<br />

5-2. An advantage of the mercury in<br />

steel thermometer is that measurements<br />

can be taken a distance away from the<br />

application, as the steel tube can be<br />

made fairly long and flexible . The<br />

whole system is completely filled with<br />

mercury under pressure and sealed off.<br />

When the temperature around the bulb<br />

increases, the mercury inside the bulb<br />

will expand. The effect of the mercury<br />

trying to increase its volume within a<br />

confined space, will be an increase in pressure, transmitted via the capillary tube,<br />

to the coiled Bourdon tube. The increase in mercury volume and pressure inside<br />

the coiled Bourdon tube, will result in the Bourdon tube starting to uncoil,<br />

proportional to the temperature. A pointer, linked to the free end of the Bourdon<br />

tube, will subsequently move over the scale to indicate the temperature.<br />

5.6 GAS FILLED THERMOMETERS<br />

Figure 5-2<br />

Pointer<br />

and scale<br />

Bourdon<br />

tube<br />

Steel tube<br />

(capillary)<br />

Mercury<br />

Steel bulb<br />

The thermometer consists of a steel<br />

bulb, a steel tube and Bourdon tube, as<br />

shown in Figure 5-3. The whole system<br />

is filled with a gas at a high pressure<br />

and sealed off. The gas commonly used<br />

is nitrogen which is an inert gas, with a<br />

high cubical expansion coefficient and<br />

which is readily available. It may be<br />

assumed that the volume of the gas<br />

remains nearly constant. For an ideal<br />

gas, the gas law applies:<br />

PV = nRT.<br />

It is clear from this equation that if V is<br />

constant, the pressure is proportional to<br />

the temperature T. Temperature is thus<br />

converted to pressure, which is<br />

detected by the Bourdon tube.<br />

Figure 5-3<br />

Pointer<br />

and scale<br />

Bourdon<br />

tube<br />

Steel tube<br />

(capillary)<br />

Gas<br />

(nitrogen)<br />

Steel bulb


EIPINI Chapter 5: Temperature Measurement Page 5-7<br />

5.7 VAPOUR PRESSURE THERMOMETERS<br />

The thermometer consists of a steel<br />

bulb, a steel tube and Bourdon tube, as<br />

shown in Figure 5-4. The bulb is<br />

partly filled with a volatile liquid,<br />

evacuated and sealed off. With rising<br />

temperature, the average velocity of<br />

the molecules in the liquid will<br />

increase. As a result, more liquid<br />

molecules will acquire enough energy<br />

to escape from the surface of the liquid<br />

and saturate the evacuated area. The<br />

result will be an increase in vapour<br />

pressure in the capillary tube and<br />

Bourdon tube. The Bourdon tube,<br />

sensitive to the changes in pressure,<br />

will record the temperature via the<br />

pointer that is linked to the moving<br />

end of the Bourdon tube.<br />

5.8 BI-METAL THERMOMETERS<br />

Bimetal strips consists of two different<br />

metals welded together to form a cantilever<br />

as shown in Figure 5-5 (a). When heated,<br />

both metals expand, but the bottom strip<br />

(assuming that the lower strip has the higher<br />

expansion coefficient), expands more than<br />

the top metal. The result is that the bottom<br />

of the strip becomes longer than the top, and<br />

the cantilever curls upwards as shown in<br />

Figure 5-5 (b).<br />

Figure 5-4<br />

Pointer and<br />

scale<br />

Bourdon<br />

tube<br />

Steel tube<br />

(capillary)<br />

Vapour<br />

Steel bulb<br />

Volatile<br />

liquid<br />

Pointer<br />

and scale<br />

Socket<br />

Bearing<br />

Figure 5-5 (a)<br />

Figure 5-5 (b)<br />

Shaft<br />

Guide<br />

A bi-metal thermometer uses a bi-metal strip, shaped in<br />

a helix or spiral form, as shown in Figure 5-6. The one<br />

end is fixed and the other end is free to rotate as the helix<br />

curls in or out with changing temperature. A shaft and<br />

pointer is linked to the rotating helix, to indicate the<br />

temperature. The stem is filled with silicone fluid, to<br />

provide damping and thermal conductivity between the<br />

stem and bi-metal strip.<br />

Stem<br />

Helical<br />

bi-metal<br />

element<br />

Figure 5-6


EIPINI Chapter 5: Temperature Measurement Page 5-8<br />

5.9 RESISTANCE TEMPERATURE DETECTORS (RTD’s)<br />

5.9.1 Temperature coefficient of resistance (TCR)<br />

Resistance thermometers are based on the principle that the resistance of a metal<br />

increases with temperature. The temperature coefficient of resistance (TCR) for<br />

resistance thermometers (denoted by α o ), is normally defined as the average resistance<br />

change per °C over the range 0 °C to 100 °C, divided by the resistance of the<br />

thermometer, R o , at 0 °C.<br />

α o =<br />

R 100 − R 0<br />

R 0 × 100<br />

Equation 5-3<br />

where,<br />

R 0 = resistance of wire at 0 °C (ohm), and<br />

R 100 = resistance of wire at 100 °C (ohm),<br />

As a first approximation, the relationship<br />

between resistance and temperature, may<br />

then be expressed as (see Figure 5-7):<br />

R t (Ω)<br />

R o<br />

Note: Starting with a straight line<br />

in standard form, y = mx + c:<br />

R t = mt + c where c = R o<br />

and m = (R 100 - R o )/100 = R o α o .<br />

∴R t = R o α o t + R o = R o (1 + α o t)<br />

R t = R o (1+α o t)<br />

t (°C)<br />

R 100<br />

Figure 5-7<br />

R t = R o (1 + α o t), Equation 5-4<br />

where:<br />

R t = resistance of wire at<br />

temperature t (ohm),<br />

R o = resistance of wire at 0 °C (ohm), and<br />

α o = temperature coefficient of resistance (TCR) at 0 °C (per °C)<br />

Example 5-5<br />

A platinum thermometer measures 100 Ω at 0 °C and 139.1 Ω at 100 °C.<br />

a) Calculate the TCR for platinum.<br />

b) Use Equation 5-4 and calculate the resistance of the thermometer at 50 °C.<br />

c) Use Equation 5-4 and calculate the temperature when the resistance is 110 Ω.<br />

R<br />

a) From Equation 5-3: α o = 100 − R 0 139.1-100<br />

= = 0.00391 /°C.<br />

R 0 × 100 100×<br />

100<br />

b) From Equation 5-4: R 50 = R o (1 + α o t) = 100(1 + 0.00391×50) = 119.55 Ω<br />

c) From Equation 5-4: R t = R o (1 + α o t) ⇒ 110 = 100(1 + 0.00391t)<br />

∴1 + 0.00391t = 1.1 ⇒ 0.00391t = 0.1 ⇒ t = 25.58 °C.<br />

5.9.2 Resistance curves<br />

In general, the relationship between resistance and temperature, is not linear and<br />

Equation 5-4 does not describe the resistance / temperature behaviour of metals,<br />

adequately. In Figure 5-8, the resistance curves for nickel, copper and platinum (with<br />

R o = 100 Ω) are shown and it is immediately clear that higher order equations must be<br />

used to describe the resistance / temperature relationship. The resistance / temperature<br />

characteristic for a platinum thermometer, for example, can be described by the<br />

equation:<br />

0<br />

100


EIPINI Chapter 5: Temperature Measurement Page 5-9<br />

⎪<br />

⎧<br />

⎨<br />

⎪⎩<br />

Ro[1+<br />

At + Bt + Ct −<br />

< < °<br />

=<br />

3 (t 100)] - 200 t 0 C<br />

R 2<br />

t<br />

Ro[1+<br />

At + Bt 2 ,<br />

]<br />

0 ≤ t < 850 ° C<br />

where:<br />

R t is the resistance at temperature t<br />

R o is the ice point resistance and<br />

A, B and C are coefficients describing the particular thermometer.<br />

800<br />

600<br />

400<br />

200<br />

-200<br />

R (ohm)<br />

Nickel<br />

Copper<br />

Platinum<br />

0 200 400 600 800<br />

Figure 5-8<br />

t (°C)<br />

An excerpt from the resistance table for a Minco<br />

platinum thermometer (<strong>code</strong> PB), is given in Table 5-1.<br />

Example 5-6<br />

In the temperature range 0 °C ≤ t < 850 °C, the<br />

resistance / temperature relation of a platinum<br />

thermometer with ice point resistance of 100 Ω, is<br />

described by the second order expression:<br />

Partial temperature/<br />

resistance table for<br />

platinum (R o =100Ω)<br />

Temperature<br />

(°C)<br />

Resistance<br />

(Ω)<br />

-200 17.26<br />

-100 59.64<br />

-80 67.83<br />

-60 75.96<br />

-40 84.03<br />

-20 92.04<br />

0 100.00<br />

20 107.92<br />

40 115.78<br />

60 123.60<br />

80 131.38<br />

100 139.11<br />

200 177.04<br />

300 213.81<br />

400 249.41<br />

500 283.84<br />

600 317.09<br />

700 349.18<br />

R t = 100×[1 + 0.0039692×t – (5.8495×10 -7 )×t 2 ].<br />

Table 5-1<br />

a) Calculate the resistance of the thermometer at 50 °C.<br />

b) Calculate the temperature when a resistance of 110 Ω is measured.<br />

a) R 50 = 100×(1 + 0.0039692×50 – 5.8495×10 -7 ×50 2 )<br />

= 100×(1 + 0.19846 - 0.001462375) = 119.7 Ω.<br />

b) R t = 100[1 + (3.9692×10 -3 )×t – (5.8495×10 -7 )t 2 ].<br />

∴110 = 100[1 + (3.9692×10 -3 )×t – (5.8495×10 -7 )t 2 ]<br />

∴110 = 100 + 0.39692t – 5.8495×10 -5 t 2<br />

∴(5.8495×10 -5 )t 2 + (– 0.39692)t + 10 = 0<br />

∴t =<br />

=<br />

0.39692 ±<br />

( 0.39692 )<br />

2×<br />

( 5.8495×<br />

10<br />

0.39692 ± 0.39396<br />

−6<br />

116.99×<br />

10<br />

2<br />

− 4×<br />

( 5.8495×<br />

10<br />

−5<br />

)<br />

−5<br />

) × 10<br />

= 25.301 °C (ignoring 6760 °C.)<br />

[ax 2 +bx+c=0]<br />

[x=<br />

− b ±<br />

b<br />

2<br />

− 4ac<br />

2a<br />

]


EIPINI Chapter 5: Temperature Measurement Page 5-10<br />

5.9.3 Thermometer construction<br />

Depending upon their intended<br />

application, resistance thermometers are<br />

available in different shapes and forms.<br />

A typical structure is shown in Figure<br />

5-9. The resistance winding is located in<br />

the lower end of the protecting tube or<br />

stem, which is sealed. The upper part of<br />

the stem terminates in the terminal<br />

housing (head) with the resistor winding<br />

leads. The resistance winding must be in<br />

good thermal contact with the stem, for<br />

fast heat transfer from the medium to the<br />

winding while electrical isolation must<br />

be ensured. The fundamental design<br />

problem with resistance thermometers is<br />

to achieve high electrical insulation and<br />

minimum thermal insulation.<br />

Terminal<br />

cap<br />

Connector<br />

conduit<br />

Socket<br />

Resistor bulb<br />

(resistance winding<br />

or thin film)<br />

Figure 5-9<br />

Stem<br />

Leads<br />

5.9.4 Measuring temperature with resistance thermometers<br />

The simplest way to measure the<br />

resistance of a RTD, is to inject a<br />

constant current into the thermometer<br />

and to measure the voltage that develops<br />

across the thermometer. A Wheatstone<br />

bridge circuit, shown in Figure 5-10, is<br />

however generally used to detect the<br />

changes in resistance of a resistance<br />

thermometer. The values of the fixed<br />

resistors, R 1 ,R 2 and R 3 , are very<br />

accurately known, while RT represents<br />

the resistance thermometer with leads a<br />

and b. The bridge is said to be in null<br />

balance, when the voltage across points<br />

A and B is zero. This occurs when RT = R 3 ×<br />

R<br />

1<br />

R<br />

2<br />

, causing VAC = V BC , resulting in<br />

the reading on M, to become zero. The zero condition would correspond to the zero<br />

point or set point of the resistance thermometer output. As the temperature<br />

increases, the resistance RT, of the resistance thermometer will increase, causing<br />

the bridge to become unbalanced, and meter M to show a reading. The meter M<br />

may be calibrated in temperature units or V AB may be converted into a standard 4 to<br />

20 mA or 1 to 5 V signal. The current flowing through the thermometer must be<br />

kept as low as possible (< 1 mA) to minimise errors caused by I 2 R losses and<br />

associated temperature rise in the thermometer itself.<br />

E<br />

B<br />

R 2<br />

R 3<br />

M<br />

C<br />

Figure 5-10<br />

a<br />

R 1<br />

b<br />

RT<br />

A


EIPINI Chapter 5: Temperature Measurement Page 5-11<br />

Example 5-7<br />

The Wheatstone bridge in Figure 5-10, is supplied from a 10 V battery, and used to<br />

measure temperature with a platinum resistance RT. The value of each of the fixed<br />

resistors R 1 , R 2 and R 3 , is 100 Ω (in a practical circuit, R 1 and R 2 would be chosen<br />

much higher to improve bridge linearity and minimize errors caused by I 2 R losses in<br />

the thermometer). The meter M measures the thermometer output voltage V, which<br />

is given by: V = V AB = V AC - V BC . At 0 °C, the resistance RT, of the platinum RTD,<br />

is 100 Ω, while at 100 °C, the resistance RT increases to 139.1 Ω. Calculate the<br />

output voltage V, of the thermometer, when the temperature is 0 °C and when the<br />

temperature is 100 °C. Assume that the measuring device draws negligible current<br />

from the circuit.<br />

0 °C:<br />

I 2<br />

I 1<br />

R<br />

I 1 = 10/(100 + 100) = 50.00 mA.<br />

2 R 1<br />

∴V AC = RT×I 1 = 100×(50×10 -3 )<br />

10V 100Ω V 100Ω<br />

= 5 volt<br />

B<br />

A<br />

I 2 = 10/(100 + 100) = 50.00 mA.<br />

R 3<br />

V<br />

∴V BC = R 3 ×I 2 = 100×(50×10 -3 AB<br />

RT<br />

V BC V AC<br />

)<br />

100Ω 100Ω<br />

= 5 volt<br />

∴V = V AB = V AC – V BC (using Kirchoff’s<br />

C<br />

0 °C<br />

C<br />

= 5 – 5 = 0 volt voltage law: V AB -V AC +V BC =0)<br />

We of course expected this result, as the bridge is indeed balanced at 0 °C.<br />

100 °C:<br />

I 1 =10/(100+139.1)=41.82 mA.<br />

∴V AC = RT×I 1 = 139.1×(41.82×10 -3 )<br />

= 5.817 volt<br />

I 2 = 10/(100 + 100) = 50.00 mA.<br />

∴V BC = R 3 ×I 2 = 100×(50×10 -3 )<br />

= 5 volt<br />

∴V = V AB = V AC – V BC<br />

= 5.817 – 5.000 = 0.817 volt<br />

Example 5-8<br />

The Wheatstone bridge in Figure 5-10, is supplied from a 10 V battery, and used to<br />

measure temperature with a platinum resistance RT. The value of each of the fixed<br />

resistors R 1 , R 2 and R 3 , is 100 Ω. The meter M measures the thermometer output<br />

voltage V, which is given by: V = V AB = V AC - V BC .<br />

a) Derive an expression for the resistance RT of the platinum element, in terms of<br />

the thermometer output voltage V.<br />

b) Derive an expression for the temperature t in terms of RT, if in the temperature<br />

range 0 °C ≤ t < 850 °C, the relationship between the resistance RT of the<br />

thermometer and the temperature t, is given by:<br />

RT = 100[1 + (3.9692×10 -3 )×t – (5.8495×10 -7 )t 2 ].<br />

c) Calculate the measured temperature when the bridge output voltage is 0, 0.2, 0.4,<br />

0.6, 0.8, and 1.0 volt.<br />

10V<br />

I 2<br />

V BC<br />

B<br />

C<br />

R 2<br />

100Ω<br />

R 3<br />

100Ω<br />

V<br />

100 °C<br />

R 1<br />

100Ω<br />

RT<br />

139.1Ω<br />

I 1<br />

A<br />

C<br />

V AC


EIPINI Chapter 5: Temperature Measurement Page 5-12<br />

10<br />

a) I 1 = ampere<br />

RT + 100<br />

I 2<br />

I 1<br />

10×<br />

RT<br />

R 2 R 1<br />

∴V AC = I 1 ×RT = volt<br />

RT + 100<br />

10V 100Ω V 100Ω<br />

10<br />

and I 2 = = 50 milliamp.<br />

B<br />

A<br />

100 + 100<br />

R 3<br />

∴V BC = I 2 ×R 3 = (50×10 -3 )×100<br />

V BC<br />

100Ω RT<br />

= 5 volt<br />

∴V = V AB = V AC - V BC<br />

C<br />

C<br />

10×<br />

RT 10×<br />

RT<br />

∴V = - 5 ⇒ = V + 5 ⇒ 10×RT = (RT + 100)×(V + 5)<br />

RT + 100 RT + 100<br />

∴10×RT = V×RT + 100×V + 5×RT + 100×5<br />

∴10×RT – 5×RT - V×RT = 100×V + 100×5<br />

∴(10 – 5 – V)×RT = 100×(V + 5) ⇒ (5 – V)×RT = 100×(V + 5)<br />

100(V + 5)<br />

∴RT =<br />

…………………………….………………. Equation (a)<br />

5 - V<br />

b) RT = 100[1 + (3.9692×10 -3 )×t – (5.8495×10 -7 )t 2 ]<br />

∴RT = 100 + 0.39692×t – (5.8495×10 -5 )×t 2<br />

∴(5.8495×10 -5 )×t 2 - 0.39692×t + RT – 100 = 0<br />

5.8495×<br />

10<br />

∴<br />

-5<br />

5.8495×<br />

10 -5 ×t 2 0.39692<br />

1<br />

-<br />

5.8495×<br />

10 -5 ×t +<br />

5.8495×<br />

10 -5 ×(RT-100) = 0<br />

∴ t 2 - 6786t + [17095×(RT-100)] = 0<br />

[x 2 +bx+c=0]<br />

2<br />

6786 − (6786) − 4×<br />

17095×<br />

(RT −100)<br />

− b ± b<br />

2<br />

− 4c<br />

∴t =<br />

[x= ]<br />

2<br />

2<br />

6786 − 46.05×<br />

10 6 − 68380(RT −100)<br />

∴t =<br />

…….………….. Equation (b)<br />

2<br />

(Note: From Example 5-6 we already know that we can ignore the second<br />

solution associated with the positive square root.)<br />

c) V = 0: From Equation (a): RT = 100(0 + 5)/(5 - 0) = 100 Ω<br />

From Equation (b): t = {6786 - √[46.05×10 6 - 68380(100 - 100)]}/2<br />

= [6786 - 6786]/2 = 0 °C<br />

V = 0.2: From Equation (a): RT = 100(0.2 + 5)/(5 - 0.2) = 108.33 Ω<br />

From Equation (b): t = {6786 - √[46.05×10 6 -68380(108.33-100)]}/2<br />

= [6786 - √45.48×10 6 ]/2 = 21.05 °C<br />

V = 0.4: RT = 100(0.4 + 5)/(5 - 0.4) = 117.39 Ω<br />

t = {6786 - √[46.05×10 6 - 68380(117.39 - 100)]}/2 = 44.09 °C<br />

V = 0.6: RT = 100(0.6 + 5)/(5 – 0.6) = 127.27 Ω<br />

t = {6786 - √[46.05×10 6 - 68380(127.27 - 100)]}/2 = 69.40 °C<br />

V = 0.8: RT = 100(0.8 + 5)/(5 - 0.8) = 138.1 Ω<br />

t = {6786 - √[46.05×10 6 - 68380(138.1 - 100)]}/2 = 97.37 °C<br />

V AC


EIPINI Chapter 5: Temperature Measurement Page 5-13<br />

V = 1.0:<br />

RT = 100(1 + 5)/(5 - 1) = 150 Ω<br />

t = {6786 - √[46.05×10 6 - 68380(150 - 100)]}/2 = 128.4 °C<br />

The results from Example<br />

5-8, highlight again that the<br />

t (°C)<br />

relationship between the 140<br />

thermometer output and the<br />

measured temperature, is not<br />

120<br />

linear (a graph of our 100<br />

calculated temperatures versus<br />

thermometer output, is shown<br />

to the right). The non-linear<br />

behaviour arises from the nonlinear<br />

relationship between RT<br />

and V in Equation (a) and<br />

from the non-linear<br />

80<br />

60<br />

40<br />

20<br />

V (volt)<br />

relationship between the<br />

temperature t and RT in<br />

Equation (b).<br />

0<br />

0.2 0.4 0.6 0.8 1.0<br />

In Example 5-8, temperature and resistance were related by a simple quadratic<br />

equation. When converting resistance to temperature, thermometer manufacturers may<br />

in certain cases, be required to find the root of a third or fourth order polynomial. In<br />

the case of nickel, individual manufacturers have developed different step-wise<br />

approximations to calculate temperature from resistance. Fortunately today, many<br />

instruments use micro<strong>process</strong>ors and manufacturers are using resistance / temperature<br />

lookup tables stored in read only memory, with linear interpolation for measurements<br />

falling between table entries.<br />

5.9.5 Ambient temperature compensation<br />

If it is necessary to perform the temperature measurement, some distance away<br />

from where the resistance thermometer bulb is installed, ambient temperature can<br />

have a detrimental effect on the integrity of the measurement. The reason is that the<br />

leads that connect the thermometer to the instrument, will have resistance of their<br />

own, and what is more, this resistance will change with changing ambient<br />

temperature. The error introduced by lead resistance, can of course be minimised,<br />

if leads with very low resistance is used.<br />

An ingenious approach to the lead resistance problem, is the three wire method,<br />

illustrated in Figure 5-11. The three wire method attempts to cancel the effect of<br />

the lead resistance, by introducing the same resistance in both branches (wires c<br />

and a) of the Wheatstone bridge. The three wire method is used extensively for<br />

industrial RTD’s. Three leads a, b and c, connect to the resistance thermometer and<br />

the resistance of wires a and c must be matched. This method assumes that the<br />

meter is a high impedance device, and that essentially no current is flowing in the b<br />

wire, which only acts as a voltage sense lead.


EIPINI Chapter 5: Temperature Measurement Page 5-14<br />

R 1<br />

E<br />

A<br />

R 2<br />

V<br />

B<br />

c<br />

b<br />

R lead<br />

R lead<br />

a<br />

R lead<br />

RT<br />

R 3<br />

C<br />

Figure 5-11<br />

The four wire method, depicted in Figure 5-12, alleviates many problems<br />

associated with the Wheatstone bridge. Instead of using a Wheatstone bridge<br />

configuration, a current source is employed to supply a constant current I, to the<br />

thermometer, through wires a and d. A high impedance voltmeter M, measures the<br />

voltage developed across the thermometer, via wires b and c. The measured<br />

voltage is directly proportional to the resistance of the thermometer, so only the<br />

conversion from resistance to temperature is necessary. Wires b and c only act as<br />

voltage sense leads, and with M a high impedance meter, virtually no current flows<br />

in wires b and c, and therefore no voltage drop in these leads and thus no lead<br />

resistance error in the measurement. The disadvantage of the four wire method is<br />

that we need one more wire than with the three wire method, but it is a small price<br />

to pay if we are at all concerned with the accuracy of the temperature<br />

measurement. The four wire method is however gaining in popularity in industry.<br />

a<br />

R lead<br />

b<br />

R lead<br />

I<br />

M<br />

c<br />

R lead<br />

RT<br />

Figure 5-12<br />

d<br />

R lead<br />

Example 5-9<br />

A Wheatstone bridge, with fixed resistors R 1 , R 2 and R 3 , of 100 Ω and supplied from a<br />

10 V battery, is used to measure a temperature of 100 °C with a platinum resistance<br />

RT, that has a resistance of 139.1 Ω at 100 °C . The thermometer is connected to the<br />

bridge with leadwires that have a resistance R lead of 10 Ω at 20 °C and the thermometer<br />

is correctly calibrated for an ambient temperature of 20 °C. Assume that the lead<br />

resistance changes to 11 Ω when the ambient temperature rises to 30 °C.


EIPINI Chapter 5: Temperature Measurement Page 5-15<br />

Calculate the temperature error introduced at an ambient temperature of 30 °C, when<br />

using a) the two wire method and b) the three wire method<br />

a) Measuring with two leads<br />

Calibrated bridge voltage at<br />

Bridge output voltage at<br />

20 °C ambient temperature 30 °C ambient temperature<br />

I 2<br />

100Ω I 1<br />

100Ω<br />

I 2<br />

100Ω I 1<br />

100Ω<br />

10V<br />

Process<br />

temberature<br />

100°C<br />

I 1 = 10/(100+10+139.1+10) = 0.0386 A I 1 = 10/(100+11+139.1+11) = 0.0383 A<br />

I 2 = 10/(100+100) = 0.05 A<br />

I 2 = 10/(100+100) = 0.05 A<br />

V = (10+139.1+10)×0.0386 - 100×0.05 V = (11+139.1+11)×0.0383 - 100×0.05<br />

= 6.141 – 5 = 1.141 V (calibrated 100°C) = 6.17 – 5 = 1.17 V (105°C)<br />

Therefore, if the bridge was correctly calibrated for an ambient temperature of 20 °C,<br />

an error of 0.029 V will occur when the ambient temperature rises to 30 °C. So instead<br />

of reading 100 °C, it would measure approximately 105 °C, an error of 5 °C or 5%.<br />

b) Measuring with three leads<br />

Calibrated bridge voltage at<br />

Bridge output voltage at<br />

20 °C ambient temperature 30 °C ambient temperature<br />

10V<br />

100Ω<br />

I 2<br />

I 1<br />

100Ω<br />

100Ω<br />

V<br />

V<br />

b<br />

a<br />

100Ω<br />

c<br />

b<br />

a<br />

10Ω<br />

139.1Ω<br />

RT<br />

10Ω<br />

10Ω<br />

10Ω<br />

139.1Ω<br />

RT<br />

10Ω<br />

Process<br />

temberature<br />

100°C<br />

I 2<br />

I 1<br />

100Ω<br />

100Ω<br />

100Ω<br />

Process<br />

temberature<br />

100°C<br />

Process<br />

temberature<br />

100°C<br />

I 1 = 10/(100+10+139.1+10) = 0.0386 A I 1 = 10/(100+11+139.1+11) = 0.0383 A<br />

I 2 = 10/(100+100) = 0.05 A<br />

I 2 = 10/(100+100) = 0.05 A<br />

V = (139.1 + 10)×0.0386 - 100×0.05 V = (139.1 + 11)×0.0383 - 100×0.05<br />

= 5.755 – 5 = 0.755 V (calibrated 100°C) = 5.749 – 5 = 0.749 V (99.13°C)<br />

Therefore, measuring with the three wire method, an error of only 0.006 V will<br />

occur. If the thermometer was correctly calibrated to read 100 °C at an ambient<br />

temperature of 20 °C, then if the ambient temperature rises to 30 °C, the reading of<br />

0.749 V would correspond to a value of 138.76 Ω for RT with no rise in ambient<br />

temperature, or in terms of temperature, 99.13 °C, an error of 0.87 °C or 0.87%. This<br />

is clearly a remarkable improvement over the two wire method.<br />

10V<br />

10V<br />

100Ω<br />

V<br />

V<br />

b<br />

a<br />

c<br />

b<br />

a<br />

11Ω<br />

139.1Ω<br />

RT<br />

11Ω<br />

11Ω<br />

11Ω<br />

139.1Ω<br />

RT<br />

11Ω


EIPINI Chapter 5: Temperature Measurement Page 5-16<br />

5.10 THERMOCOUPLES<br />

5.10.1 The Seebeck effect<br />

If two dissimilar metals are joined together to form a closed loop, and if one<br />

junction is kept at a different temperature from the other, an electromotive force is<br />

generated and electric current will flow in the closed loop.<br />

Metal A<br />

Metal B<br />

Seebeck emf<br />

Metal B<br />

This very important discovery in the field of thermometry, was made by T.J. Seebeck<br />

in the year 1821. Experiments by Seebeck and others, have shown that the generated<br />

emf (called the Seebeck voltage in his honor), is relative, in a predictable manner, to<br />

the difference in temperature between the two junctions. So, if the temperature of one<br />

junction is kept at a known value, the temperature of the other junction can be<br />

determined by the amount of voltage produced. This discovery resulted in the<br />

temperature sensor that we know today as the thermocouple.<br />

The Seebeck voltage is made up of two components: the Peltier voltage generated at<br />

the junctions, plus the Thomson voltage generated in the wires by the temperature<br />

gradient. The Peltier voltage is proportional to the temperature of each junction while<br />

the Thomson voltage is proportional to the square of the temperature difference<br />

between the two junctions. It is the Thomson voltage that accounts for most of the<br />

observed voltage and non-linearity in thermocouple response.<br />

5.10.2 Thermocouple laws<br />

The following empirically derived thermocouple laws, are useful to understand,<br />

diagnose and utilise thermocouples.<br />

a) Law of homogeneous circuits<br />

If two thermocouple junctions are at T 1 and T 2 , then the thermal emf generated is<br />

independent and unaffected by any temperature distribution along the wires<br />

T 3<br />

T 1<br />

T 4<br />

T 2<br />

Figure 5-13<br />

In Figure 5-13, a thermocouple is shown with junction temperatures at T 1 and T 2 .<br />

Along the thermocouple wires, the temperature is T 3 and T 4 . The thermocouple emf is,<br />

however, still a function of only the temperature gradient T 2 – T 1 .


EIPINI Chapter 5: Temperature Measurement Page 5-17<br />

b) Law of intermediate metals<br />

The law of intermediate metals states that a third metal may be inserted into a<br />

thermocouple system without affecting the emf generated, if, and only if, the<br />

junctions with the third metal are kept at the same temperature.<br />

Metal A<br />

T 1<br />

T 2<br />

Metal B<br />

Metal C<br />

Metal B<br />

T 3<br />

Figure 5-14<br />

When thermocouples are used, it is usually necessary to introduce additional metals<br />

into the circuit This happens when an instrument is used to measure the emf, and when<br />

the junction is soldered or welded. It would seem that the introduction of other metals<br />

would modify the emf developed by the thermocouple and destroy its calibration.<br />

However, the law of intermediate metals states that the introduction of a third metal<br />

into the circuit will have no effect upon the emf generated so long as the junctions of<br />

the third metal are at the same temperature, as shown in Figure 5-14.<br />

c) Law of intermediate temperatures<br />

The law of intermediate temperatures states that the sum of the emf developed by a<br />

thermocouple with its junctions at temperatures T 1 and T 2 , and with its junctions at<br />

temperatures T 2 and T 3 , will be the same as the emf developed if the thermocouple<br />

junctions are at temperatures T 1 and T 3 .<br />

T 1<br />

T 2<br />

T 2 T 3 T 1<br />

T 3<br />

emf (T 2 -T 1 ) emf (T 3 -T 2 ) emf (T 3 -T 1 )<br />

Figure 5-15<br />

This law, illustrated in Figure 5-15, is useful in practice because it helps in giving a<br />

suitable correction in case a reference junction temperature other than 0 °C is<br />

employed. For example, if a thermocouple is calibrated for a reference junction<br />

temperature of 0 °C and used with a junction temperature of 20 °C, then the correction<br />

required for the observation would be the emf produced by the thermocouple between<br />

0 °C and 20 °C.<br />

5.10.3 Thermocouple types<br />

Any two dissimilar metals can in theory be made into a thermocouple. However,<br />

certain metals have been selected over time that make ideal thermocouples for various<br />

applications. These metals have been chosen for their emf output and their ability to<br />

operate under various conditions. There are several types of these “standard”<br />

thermocouples in use today. Some are listed Table 5-2.


EIPINI Chapter 5: Temperature Measurement Page 5-18<br />

Type Positive Isolation Negative Isolation Outer Temperature<br />

element colour element colour isolation range (°C)<br />

K Chromel Yellow Alumel Red Yellow -200 to 1200<br />

E Chromel Purple Constantan Red Purple -200 to 800<br />

J Iron White Constantan Red Black -200 to 750<br />

T Copper Blue Constantan Red Blue -200 to 350<br />

S<br />

90% Platinum<br />

10% Rhodium<br />

Black Platinum Red Green 0 to 1500<br />

Table 5-2<br />

Types K, E, J and T, are the 'general purpose' thermocouples. Type K is the most<br />

popular thermocouple in use today while type T is used for lower temperature<br />

applications. The conventional type J thermocouple, even with its unfavourable iron<br />

lead, is still popular, mainly because of its widespread use in older instruments. Type<br />

E is the most sensitive of the standard thermocouples (68μV/°C). Noble metal type S,<br />

has low sensitivity(10μV/°C), is expensive and used in the higher temperature range.<br />

Thermocouple alloys referred to in Table 5-2, are chromel (chrome and nickel), alumel<br />

(aluminium and nickel) and constantan (copper and nickel).<br />

Different colour <strong>code</strong>s have been adopted for thermocouple wire and product<br />

identification and the colour <strong>code</strong>s listed in Table 5-2, are used in the United States.<br />

5.10.4 Thermocouple construction<br />

Thermocouple thermometers are available in different shapes and in different<br />

coverings. Three basic types of construction can be recognised:<br />

Wire construction<br />

The most basic construction of the thermocouple<br />

is the two dissimilar metals joined (welded)<br />

together to form the measuring junction. In this<br />

form, the exposed junction offers good response<br />

times, but may suffer from environmental<br />

damage. A bare thermocouple element is shown<br />

in Figure 5-16.<br />

Sheathed construction<br />

To improve mechanical strength, mineral<br />

insulated thermocouples were developed. The<br />

thermocouple wires are embedded in compressed<br />

mineral oxide powder and enclosed in a metal<br />

sheath, usually stainless steel or inconel (nickelchromium-iron<br />

alloy). A completely insulated<br />

thermocouple is shown in Figure 5-17 (a).<br />

Bare thermocouple junction<br />

Magnesium<br />

oxide<br />

Figure 5-16<br />

Insulated junction<br />

Thermocouple<br />

wires<br />

Figure 5-17 (a)<br />

Inconel<br />

sheath


EIPINI Chapter 5: Temperature Measurement Page 5-19<br />

To improve response time, the junction may be grounded as shown in Figure<br />

5-17 (b), or the thermocouple tip could be exposed as shown in Figure 5-17 (c).<br />

Grounded junction<br />

Figure 5-17 (b)<br />

Protecting tube construction<br />

Protecting tubes (or thermowells)<br />

are used to shield thermocouple<br />

sensing elements against<br />

mechanical damage, thermal shock<br />

and corrosive or contaminating<br />

atmospheres. Various types of<br />

constructions are available. A<br />

typical construction method,<br />

similar to that of RTD’s, is shown<br />

in Figure 5-18. Cast iron protection<br />

tubes are used for example, in<br />

molten aluminum, magnesium and<br />

zinc applications while ceramic<br />

tubes are used in industries such as<br />

iron and steel, glass, cement and<br />

lime <strong>process</strong>ing.<br />

Exposed junction<br />

Figure 5-17 (c)<br />

Cover<br />

Conduit<br />

entrance<br />

Socket<br />

Metal or ceramic<br />

protecting tube<br />

Sheathed<br />

thermocouple<br />

Figure 5-18<br />

5.10.5 Measuring temperature with thermocouples<br />

The voltage generated by a thermocouple is a function of the temperature difference<br />

between the measurement and reference junctions. Traditionally the reference junction<br />

was held at 0 °C by an ice bath, as shown in Figure 5-19. The thermocouple emf is<br />

measured with a high impedance voltmeter.<br />

Hot junction<br />

v<br />

Cold junction<br />

Ice bath<br />

Figure 5-19


EIPINI Chapter 5: Temperature Measurement Page 5-20<br />

The relationship between thermocouple voltage and temperature is unfortunately not<br />

linear, and it is necessary to use thermocouple temperature conversion tables to find<br />

temperature from the measured voltage. An extract from the voltage / temperature<br />

table for a type K thermocouple (0 °C reference), is given in Table 5-3.<br />

Temperature (°C) versus emf (μV) for type K thermocouple with 0 °C reference.<br />

deg C 0 1 2 3 4 5 6 7 8 9<br />

0 0 39 79 119 158 198 238 277 317 357<br />

10 397 437 477 517 557 597 637 677 718 758<br />

20 798 838 879 919 960 1000 1041 1081 1122 1163<br />

30 1203 1244 1285 1326 1366 1407 1448 1489 1530 1571<br />

40 1612 1653 1694 1735 1776 1817 1858 1899 1941 1982<br />

50 2023 2064 2106 2147 2188 2230 2271 2312 2354 2395<br />

60 2436 2478 2519 2561 2602 2644 2685 2727 2768 2810<br />

70 2851 2893 2934 2976 3017 3059 3100 3142 3184 3225<br />

80 3267 3308 3350 3391 3433 3474 3516 3557 3599 3640<br />

90 3682 3723 3765 3806 3848 3889 3931 3972 4013 4055<br />

100 4096 4138 4179 4220 4262 4303 4344 4385 4427 4468<br />

Table 5-3<br />

We could store these look-up table values in a computer and use the table to convert<br />

between emf and temperature. A more viable approach used by manufacturers<br />

however, is to approximate the table values using a power series polynomial (see<br />

Example 5-14) and allow the instrument’s micro<strong>process</strong>or or <strong>process</strong> computer, to<br />

calculate temperature from emf or the emf from temperature (inverse polynomial).<br />

The ice bath cold junction is not considered practical anymore. Instead the terminals<br />

connecting the thermocouple to the measuring device, are now assumed to play the<br />

role of the reference junction or 'cold junction', as it is still called today. The reference<br />

junction temperature may now be kept, for example, at room temperature, where the<br />

junction temperature is measured with an auxiliary temperature sensor, such as a RTD.<br />

According to the law of intermediate temperatures, the thermocouple voltage that<br />

corresponds to the cold junction temperature, may be added to the measured<br />

thermocouple voltage. The true temperature of the hot junction, with respect to 0 °C,<br />

can then be determined from this augmented voltage.<br />

Example 5-10<br />

Calculate the average sensitivity (μV/°C) of a type K thermocouple in the temperature<br />

range 0 °C to 100 °C.<br />

From Table 5-3 the change in emf developed by a type K thermocouple from 0 °C to<br />

100 °C, is 4096 μV. The average sensitivity is therefore 4096/100 = 40.96 μV/°C.


EIPINI Chapter 5: Temperature Measurement Page 5-21<br />

Example 5-11<br />

The cold junction of a type K thermocouple is kept at 0 °C. Use Table 5-3 to<br />

determine the temperature if the measured voltage is a) 798 μV and b) 2602 μV.<br />

a) 20 °C b) 64 °C<br />

Example 5-12<br />

The relationship between emf and temperature for a certain (imaginary) thermocouple,<br />

is described by the relation: v = t 2 , where v is the generated thermocouple emf in<br />

microvolt (μV), and t the temperature difference in °C, between the hot junction and<br />

0 °C. If the thermocouple emf reading is 3000 microvolt and the temperature of the<br />

cold junction is 25 °C, calculate the temperature of the hot junction.<br />

Emf corresponding to (25 – 0) °C = 25 2 = 625 μV<br />

Total emf (T - 0) = 3000 + 625 = 3625 μV<br />

According to the law of intermediate temperatures:<br />

Hot junction temperature T = √v = √3625 = 60.21 °C<br />

(T is NOT = √3000 + 25 = 54.77 + 25 = 79.77 °C)<br />

Example 5-13<br />

An unknown temperature is measured with a type K thermocouple. A thermocouple<br />

voltage of 2602 μV is measured. If the cold junction temperature is 20 °C, calculate<br />

the <strong>process</strong> temperature, measured by the hot junction side of the thermocouple.<br />

From Table 5-3, the cold junction (20°C) emf is<br />

798 μV. According to the law of intermediate<br />

temperatures, the correction voltage of 798 μV<br />

should be added to the measured voltage of<br />

2602 μV, to obtain 3400 μV. The corrected voltage<br />

represents the thermocouple emf that would be<br />

obtained, if the reference junction was kept at 0 °C.<br />

Again from Table 5-3, the temperature that<br />

corresponds to 3400 μV, is somewhere between<br />

83 °C and 84 °C. To find the correct temperature,<br />

we must use linear extrapolation between these<br />

two values. The difference between 3433 μV<br />

(84 °C) and 3391 μV (83 °C) is 42 μV, while<br />

3400 μV is 9 μV more than 3391 μV (83 °C).<br />

Therefore the temperature we are looking for is<br />

83 °C plus (9/42) °C, which is 83.2143 °C.<br />

The way NOT to calculate the hot junction<br />

temperature, is to look up the measured voltage<br />

(2602 μV) as 64 °C and then to add the cold<br />

junction temperature of 20 °C, to obtain 84 °C.<br />

This is NOT CORRECT.<br />

T<br />

0°C 25°C 25°C<br />

X 1 9 = ∴X = × 1<br />

9 42 42<br />

625μV 3000μV<br />

0°C T<br />

3625μV<br />

0°C 20°C 20°C T<br />

798μV<br />

2602μV<br />

0°C T<br />

3400μV<br />

84°C<br />

1°C<br />

<br />

X<br />

83°C<br />

9μV<br />

3391μV<br />

42μV<br />

3400μV 3433μV


EIPINI Chapter 5: Temperature Measurement Page 5-22<br />

Example 5-14<br />

The following equations are provided for a type K thermocouple, to calculate<br />

temperature (t in °C) from emf (v in μV) in the temperature range 0 °C to 500 °C:<br />

t = (2.508355×10 -2 )v + (7.860106×10 -8 )v 2 - (2.503131×10 -10 )v 3<br />

+ (8.315270×10 -14 )v 4 – (1.228034×10 -17 )v 5 + (9.804036×10 -22 )v 6<br />

- (4.413030×10 -26 )v 7 + (1.057734×10 -30 )v 8 – (1.052755×10 -35 )v 9 ,<br />

Equation (a)<br />

and to calculate emf (v in μV) from temperature (t in °C) in the temperature range<br />

0 °C to 1372 °C:<br />

v = -17.600413686 + 38.921204975t + (1.8558770032×10 -2 )t 2<br />

- (9.9457592874×10 -5 )t 3 + (3.1840945719×10 -7 )t 4<br />

- (5.6072844889×10 -10 )t 5 + (5.6075059059×10 -13 )t 6<br />

- (3.2020720003×10 -16 )t 7 + (9.7151147152×10 -20 )t 8<br />

− 4<br />

2<br />

– (1.2104721275×10 -23 )t 9 + 118.5976e<br />

−1.183432×<br />

10 (t −126.9686)<br />

Equation (b)<br />

a) Use Equation (a) to calculate the temperature when the emf is 2602 μV.<br />

b) Use Equation (b) to calculate the emf when the temperature is 20 °C.<br />

a) t = (2.508355×10 -2 )×2602 + (7.860106×10 -8 )×2602 2 - (2.503131×10 -10 )×2602 3<br />

+ (8.315270×10 -14 )×2602 4 – (1.228034×10 -17 )×2602 5 + (9.804036×10 -22 )×2602 6<br />

- (4.413030×10 -26 )×2602 7 + (1.057734×10 -30 )×2602 8 – (1.052755×10 -35 )×2602 9<br />

= 65.26740 + 0.5321609 – 4.409664 + 3.811584 – 1.464694 + 0.3042627<br />

- 0.03563592 + 0.002222464 – 0.05755631<br />

= 63.95 °C<br />

b) v = -17.600413686 + 38.921204975×20 + (1.8558770032×10 -2 )×20 2<br />

- (9.9457592874×10 -5 )×20 3 + (3.1840945719×10 -7 )×20 4<br />

- (5.6072844889×10 -10 )×20 5 + (5.6075059059×10 -13 )×20 6<br />

- (3.2020720003×10 -16 )×20 7 + (9.7151147152×10 -20 )×20 8<br />

− 4<br />

2<br />

– (1.2104721275×10 -23 )×20 9 + 118.5976e<br />

−1.183432×<br />

10 (20 −126.9686)<br />

= -17.600414 + 778.4241 + 7.423508 – 0.7956607 + 0.05094551<br />

- 0.001794331 + 35.88804×10 -6 – 409.8652×10 -9 + 2.487069×10 -9<br />

- 6.197617×10 -12 + 30.61899<br />

= 798.1 microvolt<br />

(Compare these answers to those of Example 5-11)<br />

5.10.6 Compensating leads<br />

Thermocouple thermometers are normally installed some distance away from the<br />

voltmeter or computer that measures the emf generated by the thermocouple. For<br />

this purpose, cheaper and lower grade thermocouple wires, called extension wire or<br />

compensating leads, are used to connect the thermocouple to the measuring device<br />

at the reference junction. Compensating leads have the same thermoelectric


EIPINI Chapter 5: Temperature Measurement Page 5-23<br />

properties as the thermocouple and do not introduce a significant error into the<br />

temperature measurement. Compensating leads must be matched to the<br />

thermocouple and for each type of thermocouple, corresponding extension leads<br />

are available.<br />

5.11 THERMISTORS<br />

A thermistor is similar to a resistance thermometer, but a semiconductor material is<br />

used instead of a metal. A distinct advantage of thermistors over resistance<br />

thermometers, is that their temperature coefficient of resistance is approximately<br />

ten times higher than that of a resistance thermometer, causing thermistors to be<br />

much more sensitive temperature detectors. The resistance change with<br />

temperature is however very nonlinear, and unlike RTD’s, most thermistors have a<br />

negative temperature coefficient of resistance, that is, the resistance of a thermistor<br />

decreases with increasing temperature.<br />

In Figure 5-20, the resistance curve of a typical thermistor is shown. The following<br />

equation, called the Steinhart and Hart equation, is used to describe the resistance /<br />

temperature relation for thermistors:<br />

1<br />

= A + Bln(R) + ClnR 3 ,<br />

T<br />

where T is absolute temperature in Kelvin, R is the thermistor resistance at<br />

temperature T and A, B and C are coefficients that describe a given thermistor.<br />

Resistance R<br />

Figure 5-20<br />

Temperature T<br />

The resistivity of thermistors is also much larger than that of RTD’s, and<br />

thermistors can therefore be made very small which means they will respond<br />

quickly to temperature changes.<br />

Thermistors cannot be used to measure high temperatures, compared to RTDs. In<br />

fact, the maximum temperature of operation is sometimes only 100 or 200 o C. The<br />

high resistivity as well as high temperature coefficient of resistance of the<br />

thermistor, makes it feasible to use the simpler two wire technique when measuring<br />

thermistor resistance.


EIPINI Chapter 5: Temperature Measurement Page 5-24<br />

For example, a common thermistor value is 5000 ohms at 25°C. With a typical<br />

temperature coefficient of resistance of 0.04/°C, a measurement lead resistance of<br />

10 Ω, produces only 0.05 °C error. This error is a factor of 500 times less than the<br />

equivalent RTD error. Thermistors have not gained nearly the popularity of RTDs or<br />

even thermocouples in industry due to their limited span as well as other<br />

disadvantages. Since thermistors are semiconductor devices, they are quite susceptible<br />

to permanent decalibration when exposed to high temperatures. In addition,<br />

thermistors are quite fragile and great care must be taken to mount them so that they<br />

are not exposed to shock or vibration.<br />

5.12 RTD, Thermocouple or Thermistor <br />

Resistance temperature detectors (RTDs)<br />

An RTD sensing element consists of a wire coil or deposited film of pure metal. The<br />

element’s resistance increases with temperature in a known and repeatable manner.<br />

RTDs exhibit excellent accuracy over a wide temperature range and represent the<br />

fastest growing segment among industrial temperature sensors. Their advantages<br />

include:<br />

• Temperature range: Models cover temperatures from -260 to 850°C.<br />

• Repeatability and stability: Industrial RTDs typically drift less than 0.1°C/year.<br />

• Sensitivity: The voltage drop across an RTD provides a much larger output than a<br />

thermocouple.<br />

• Linearity: Platinum and copper RTDs produce a more linear response than<br />

thermocouples or thermistors. RTD non-linearities can be corrected through<br />

proper design of resistive bridge networks.<br />

• Low system cost: RTDs use ordinary copper extension leads and require no cold<br />

junction compensation.<br />

Thermocouples<br />

A thermocouple consists of two wires of dissimilar metals welded together into a<br />

junction. At the other end of the signal wires, usually as part of the input instrument, is<br />

another junction called the reference junction. Heating the sensing junction generates a<br />

thermoelectric potential (emf) proportional to the temperature difference between the<br />

two junctions. This millivolt-level emf, when compensated for the known temperature<br />

of the reference junction, indicates the temperature at the sensing tip. Published<br />

millivolt tables assume the reference junction is at 0°C. Thermocouples are simple and<br />

familiar. Designing them into systems, however, is complicated by the need for special<br />

extension wires and reference junction compensation. Thermocouple advantages<br />

include:<br />

• Extremely high temperature capability: Thermocouples with precious metal<br />

junctions may be rated as high as 1800°C.<br />

• Ruggedness: The inherent simplicity of thermocouples makes them resistant to<br />

shock and vibration.<br />

• Small size/fast response: A fine-wire thermocouple junction takes up little space<br />

and has low mass, making it suitable for point sensing and fast response.


EIPINI Chapter 5: Temperature Measurement Page 5-25<br />

Thermistors<br />

A thermistor is a resistive device composed of metal oxides formed into a bead and<br />

encapsulated in epoxy or glass. A typical thermistor shows a large negative<br />

temperature coefficient. Resistance drops dramatically and non-linearly with<br />

temperature. Sensitivity is many times that of RTDs but useful temperature range is<br />

limited. There are wide variations of performance and price between thermistors from<br />

different sources. Typical benefits are:<br />

• Low sensor cost: Basic thermistors are quite inexpensive. However, models with<br />

tighter interchangeability or extended temperature ranges often cost more than<br />

RTDs.<br />

• High sensitivity: A thermistor may change resistance by tens of ohms per degree<br />

temperature change, versus a fraction of an ohm for RTDs.<br />

• Point sensing: A thermistor bead can be made the size of a pin head for small area<br />

sensing.<br />

RTD<br />

Sensor<br />

type<br />

Thermocouple<br />

Thermistor<br />

Temperature<br />

range<br />

-260 to<br />

850°C<br />

-270 to<br />

1800°C<br />

-80 to<br />

150°C<br />

(typical)<br />

Sensor<br />

cost<br />

System<br />

cost<br />

Stability Sensitivity Linearity Specify for:<br />

Moderate Moderate Best Moderate Best<br />

Low High Low Low Moderate<br />

Low Moderate Moderate Best Poor<br />

General purpose<br />

sensing<br />

Highest accuracy<br />

Highest<br />

temperatures<br />

Best sensitivity<br />

Narrow ranges<br />

(e.g. medical)<br />

Point sensing


EIPINI Chapter 6: Process Control Page 6-1<br />

6. PROCESS CONTROL<br />

The purpose of this chapter is to introduce students to industrial <strong>process</strong> control<br />

methods used in industry to ensure that manufactured products meet predetermined<br />

quality requirements. This is accomplished by continuously monitoring the production<br />

<strong>process</strong> and automatically correcting or minimizing any deviations from the required<br />

specifications, that may be detected during the manufacturing <strong>process</strong>.<br />

6.1 INTRODUCTION<br />

The history of automatic control goes back many centuries. Water level control may<br />

already be identified in water clocks used in the middle ages while mechanical clocks,<br />

making their appearance in the 1200’s, used an escapement mechanism that may be<br />

described in terms of feedback control. With the advent of the steam engine, it was<br />

evident that some sort of speed control was needed. In what is generally considered a<br />

major event in the history of feedback control, James Watt, completed<br />

the design of the centrifugal flyball governor for regulating the<br />

speed of the rotary steam engine, in 1788. This device<br />

employed two pivoted rotating flyballs which were flung<br />

outward by centrifugal force. As the speed of rotation increased,<br />

the flyweights swung further out, operating a steam flow throttling<br />

valve which slowed the engine down. Thus, a constant speed was<br />

achieved automatically. From this point onwards, many new developments followed<br />

to improve control systems as well as mathematical research to understand physical<br />

<strong>process</strong>es. A major invention, however, was the introduction of proportional, integral<br />

and derivative (PID) control, which was formulated in 1922 by Nicholas Minorsky<br />

(1885-1970). Observing the way in which a helmsman steered a ship and compensate<br />

for the disturbances from the ocean, motivated this threefold control strategy.<br />

Although advanced tools, such as model predictive controllers, are available today,<br />

the Proportional, Integral, Derivative (PID) control strategy, is still the most widely<br />

used in modern industry, controlling more than 95% of closed loop industrial<br />

<strong>process</strong>es.<br />

6.1.1 A SIMPLE WATER LEVEL CONTROL<br />

In Figure 6-1, an operator has the simple task of<br />

keeping the water level in a tank at a prescribed<br />

value, for instance half full. If the water level<br />

should drop below the desired level due to<br />

increased water outflow, the operator must<br />

increase the water inflow by opening the inflow<br />

valve more until the desired level is restored.<br />

Similarly, if the water level should increase due<br />

to a decrease in outflow, the operator must<br />

decrease the inflow by closing the inflow valve<br />

more, until the required level is restored.<br />

Inflow<br />

valve<br />

Water inlet<br />

Inflow<br />

Figure 6-1<br />

Outflow<br />

Outflow<br />

valve


EIPINI Chapter 6: Process Control Page 6-2<br />

If the water level reaches the desired value (also called the set point), the operator<br />

must try to keep the inflow equal to the outflow. This manual control example<br />

illustrates essentially how we would expect an automatic control system to behave.<br />

In Figure 6-2, an<br />

arrangement for<br />

automatic control<br />

of the water level<br />

in the container,<br />

is shown. A<br />

level measuring<br />

instrument senses<br />

the water level in<br />

the container, and<br />

Inflow<br />

valve<br />

feeds this value back to the controller. The basic function of the controller is to<br />

calculate the difference between the desired water level and the measured water level,<br />

and to determine a sensible control action, that will minimize this difference. Once the<br />

controller has calculated an appropriate control action, this value is transmitted as a<br />

pneumatic signal to the inflow valve, which is adjusted accordingly. A well-designed<br />

controller should respond quickly to changes in the water level and always steer the<br />

level in the water container back to the required level, in a disciplined way.<br />

6.1.2 CONTROL TERMINOLOGY<br />

The water container system in Figure<br />

6-2 could be represented as an<br />

input/output system or <strong>process</strong>, as<br />

shown in Figure 6-3. The quantity that<br />

we want to regulate is the water level<br />

in the container, which is called the<br />

controlled variable. The controlled<br />

Level<br />

controller<br />

Water inlet<br />

Water inflow<br />

(Input)<br />

Level<br />

detector<br />

Water outflow<br />

Figure 6-2<br />

Water level<br />

control<br />

Process<br />

(plant)<br />

Figure 6-3<br />

Inflow<br />

Outflow<br />

Outflow<br />

valve<br />

Water level<br />

(Output)<br />

variable is also called the output variable, the <strong>process</strong> variable or the measured<br />

variable. The rate of water inflow into the container (a system input) with which we<br />

may change or manipulate the water level in the container, is called the manipulated<br />

variable. The manipulated variable is also called the controller output or control<br />

variable. A change in the rate at which the water flows out of the container (a system<br />

input), will cause the water level in the container to change or to be disturbed, and is<br />

called a disturbance variable. (Note: there could be other disturbance variables that<br />

may influence the controlled variable, for instance a leak in the container.)<br />

Three values are associated with the controlled variable. Firstly, the measured value,<br />

which is the current water level in the container, determined by the level sensor.<br />

Secondly, the desired value (normally locally programmed into the controller), which<br />

is the water level that we need or require, for example 50% full. The desired value is<br />

also called the set point or reference value. Thirdly the error value which is the<br />

difference between the desired value and the measured value of the water level.


EIPINI Chapter 6: Process Control Page 6-3<br />

1. CONTROLLED VARIABLE: Process output variable that is maintained<br />

between specified limits.<br />

2. MEASURED VALUE: Actual value of the controlled variable, as determined<br />

by the <strong>instrumentation</strong>.<br />

3. DESIRED VALUE: Required value of the controlled variable (set point).<br />

4. ERROR VALUE: The difference between the desired value and the measured<br />

value.<br />

5. MANIPULATED VARIABLE: Process input variable that is adjusted, to steer<br />

the controlled variable towards the desired value.<br />

6. DISTURBANCE VARIABLE: Process input variable that can cause the<br />

controlled variable to deviate from the desired value.<br />

Example 6-1<br />

A control system for a heat exchanger is shown in Figure 6-4. Cold water enters the<br />

container at point 1 and is heated by steam, entering the steam line 3, at point 2. The<br />

heated water leaves the system at point 7. The temperature of the hot water is<br />

measured by a thermometer at point 6 and presented to the temperature controller,<br />

block 5. The controller operates a steam valve 4, such that when the hot water is below<br />

the required temperature, the steam flow rate through the steam line is increased,<br />

allowing more heat transfer to the cold water, and when the hot water is above the<br />

required temperature, the steam flow is reduced. Identify: a) the controlled variable, b)<br />

the measured value, c) the manipulated variable and d) some disturbance variables.<br />

2<br />

Steam<br />

supply<br />

4<br />

Steam<br />

control<br />

valve<br />

Cold water<br />

1<br />

3<br />

TIC<br />

Steam outlet<br />

6<br />

5<br />

Figure 6-4<br />

Heat<br />

exchanger<br />

7<br />

Hot water<br />

a) Controlled variable: The temperature of the hot water delivered at 7.<br />

b) Measured value: Temperature indicated by the temperature detector 6.<br />

c) Manipulated variable: The steam flow rate through the steam line 3, adjusted by<br />

valve 4.<br />

d) Disturbance variables:<br />

i) Change in hot water demand (this is an energy change at the output 7 and is<br />

known as an output or demand load disturbance).<br />

ii) Change in steam supply pressure (this is an energy change at the input 2, and<br />

is known as an input or supply load disturbance).<br />

iii) More heat loss through the heat exchanger walls, because of colder ambient<br />

conditions.


EIPINI Chapter 6: Process Control Page 6-4<br />

6.1.3 OPEN LOOP AND CLOSED LOOP SYSTEMS<br />

Automatic control systems may operate under open or closed loop control. The input<br />

of an open loop system is not influenced by its output. This is in contrast with a closed<br />

loop system that monitors its output and then manipulates the input, to ensure that the<br />

output meets set conditions. An example of an open loop system is a bullet fired from<br />

a rifle and an example of a closed loop systems is a cruiser missile, fired from a ship,<br />

continuously monitoring and adjusting its trajectory.<br />

Open loop control strategies may perform well in situations where the behaviour of<br />

the system is very well defined and modelled. For example a room temperature<br />

controller set to a certain on/off heater schedule, will produce the desired room<br />

temperature. However, it is difficult to suggest an open loop control strategy for the<br />

water level control in Figure 6-2, as the factors that influence the water level are<br />

difficult to predict, and a closed loop control strategy, as indicated in Figure 6-2, is<br />

clearly the most sensible solution.<br />

Open loop system: The input to the system is not determined by the output.<br />

Closed loop system: The input to the system is determined by the output.<br />

6.1.4 FEEDBACK AND FEEDFORWARD CONTROL<br />

Feedback closed loop control systems measure changes in the controlled variable<br />

directly and feed them back as input variables to the controller. Feedforward control<br />

systems on the other hand, do not measure changes in the controlled variable but<br />

rather other variables, termed intermediate variables, which are indicative of the<br />

controlled variable.<br />

The interrelationship between the elements of a general closed loop feedback control<br />

system, is illustrated with the block diagram in Figure 6-5. The water level control<br />

system in Figure 6-2 is an example of a closed loop feedback control system. The<br />

water level provided by the level sensor, is fed back to the error detector which<br />

calculates the error. The controller uses the error value to compute a suitable<br />

command for the inflow valve, thereby regulating the water level close to the desired<br />

value.<br />

Disturbance<br />

variables<br />

Desired<br />

value<br />

Comparator<br />

Measured<br />

value<br />

Error<br />

value<br />

Control<br />

unit<br />

Manipulated<br />

variable<br />

Sensor<br />

Process<br />

Controlled<br />

variable<br />

Output<br />

Figure 6-5 Block diagram of a feedback control system


EIPINI Chapter 6: Process Control Page 6-5<br />

Feedforward control systems, on the other hand, will measure disturbance variables<br />

that directly influence the controlled variable. It is conceivable, for example, to<br />

measure the outflow in the water level control system in Figure 6-2, and to base the<br />

inflow control on the value of this disturbance variable. Because the source of the<br />

level disturbance is monitored and acted upon, a change in water level is anticipated<br />

and corrected before it even occurs.<br />

Whereas feedback systems must wait for a deviation to occur before corrective<br />

action is taken, the principle advantage of feedforward control systems is that they act<br />

before deviations occur. It is clear that feedforward control alone would not be<br />

adequate for the water level control system, as the water level would eventually drift<br />

away from set point because of measuring and modelling errors. In general,<br />

feedforward control is used to complement feedback control and to enhance system<br />

performance.<br />

Feedback control: Measure the controlled variable to determine the control strategy.<br />

Feedforward control: Measure disturbance variables to determine the control<br />

strategy.<br />

6.1.5 DIRECT ACTING AND REVERSE ACTING CONTROL<br />

A control valve that opens more when the pressure input is increased (air to open) is<br />

called a reverse acting valve. For the water level control system depicted in Figure 6-2,<br />

we have assumed that a reverse acting control valve is used. This means that when the<br />

water level drops below set point, the controller must increase its output to the control<br />

valve to increase the water inflow. This type of control action is called reverse acting,<br />

because when the water level decreases, the controller output must increase. If a direct<br />

acting (air to close) control valve was used, the control action would be direct acting,<br />

as the controller output would decrease when the water level decreases.<br />

A heating system would typically be reverse acting, as a decrease in temperature<br />

would demand an increased controller output to the heating element. A cooling<br />

system, on the other hand, would typically be direct acting because an increase in<br />

temperature would require an increased output to the cooling element.<br />

Direct acting control: A control arrangement in which the controller output<br />

increases if the measured value rises above the set point.<br />

Reverse acting control: A control arrangement in which the controller output<br />

increases if the measured value drops below the set point.<br />

6.1.6 PROCESS TIME LAGS<br />

Most <strong>process</strong>es exhibit a time delay between its input and output. For example, if the<br />

heating element of a temperature control system is switched on, the temperature will<br />

not immediately rise to the maximum value. The inlet flow valve in the level control<br />

system in Figure 6-2 is another example. When it receives a signal to open 100%, the<br />

flow rate will not immediately increase to 100% due to delays inherent in the<br />

operation of the valve.


EIPINI Chapter 6: Process Control Page 6-6<br />

Figure 6-6 shows the typical response to a step input of a <strong>process</strong> such as a control<br />

valve receiving a command to open from 50 % to 60 %, resulting in the increase in<br />

inflow from 50 % to 60 %.<br />

Input<br />

60 %<br />

50 %<br />

Output<br />

60 %<br />

50 %<br />

t d<br />

t f<br />

Time<br />

Dead time and first order lag<br />

Time<br />

Figure 6-6<br />

The response curve in Figure 6-6 illustrates two kinds of time delays that may be<br />

identified. Firstly a dead time t d , when the system does not respond at all, and<br />

secondly a first order delay t f , when the system output changes to its new final value<br />

in an exponential fashion.<br />

The dead time is associated with a delay in the transportation of material or<br />

information. For example, it may take 20 minutes to send a signal to the Rover on<br />

Mars. During this time, the vehicle can not respond to this control command, because<br />

the control information is still being transported to the vehicle. The dead time is also<br />

known as a distance-velocity lag, and the reason for this is illuminated in<br />

Example 6-2. Another term used for the dead time, is transportation lag.<br />

The first order lag is associated with a delay in the transfer of energy. For example,<br />

when you step on the accelerator of a car, the car will take time to convert the<br />

chemical energy in the fuel, into mechanical kinetic energy, and the speed increase<br />

will be exponential in nature. The first order lag is also known as a transfer lag,<br />

exponential lag or resistance-capacitance lag.<br />

Dead time: Delay due to the time it takes information or material to be transported<br />

from one point to another.<br />

First order lag: Delay due to the time it takes energy to be transferred from one<br />

point or form to another.<br />

Example 6-2<br />

In the system depicted Figure 6-7, a conveyor belt is used to transport material that<br />

was deposited by a feeder, to the delivery station. A request for more material is given<br />

to the feeder. Calculate the time it will take the sensor at the delivery site to detect a<br />

reaction to the increased demand for material. The velocity of the conveyor belt is v<br />

meter per second and the distance between the feeder and delivery point is d meter.


EIPINI Chapter 6: Process Control Page 6-7<br />

Feeder<br />

v<br />

Delivery<br />

station<br />

Figure 6-7<br />

Sensor<br />

d<br />

Any increase in material delivered, must travel a distance d with a velocity v, before it<br />

reaches the delivery point, where the sensor will detect the increase. The time it takes<br />

the material to be transported to the sensor is t = distance/velocity = d/v seconds. The<br />

dead time lag in this case is called a distance-velocity lag and it is a problem not easily<br />

handled by a PID controller.<br />

6.2 CONTROL SCHEMAS<br />

We will now discuss some of the common<br />

techniques (on-off and PID control modes)<br />

used to control closed loop feedback<br />

systems. For this purpose we will use the<br />

water level control system in Figure 6-2 as<br />

basis. Central to our system is the water<br />

container shown in Figure 6-8. We will use<br />

a container with height 5 m and base area<br />

of 1 m 2 . The level of the water in the<br />

container is H, with a maximum value of<br />

5 m. The rate of water outflow is Q OUT ,<br />

with a maximum value of 0.01 m 3 /sec. The<br />

rate of water inflow is Q IN . In order to have<br />

any control over the water level in the tank whatsoever, the maximum inflow should<br />

be more or equal to the maximum outflow. We will assume that they are equal, that is,<br />

when the demand is a maximum of 0.01 m 3 /sec. and the inflow is a maximum of<br />

0.01 m 3 /sec, the water level will remain constant.<br />

To simplify matters, we will use percentage values for all variables. The variable M<br />

will be used to represent the water level, expressed as a percentage value. The water<br />

level will be measured by an electronic DP transmitter, delivering 4 mA when the tank<br />

is empty (H = 0 m or M = 0 %) and 20 mA when the tank is full (H = 5 m or<br />

M = 100 %). We will also assume that the level detector output M will immediately<br />

(with no time delay) reflect the actual water level H.<br />

The variable C will be used to represent the controller output, expressed as a<br />

percentage value. The controller will transmit 20 kPa to the control valve to close the<br />

valve completely (C = 0 %) and 100 kPa to open the valve fully (C = 100 %). We will<br />

also assume that a linear relationship (with no time delay) exists between the<br />

controller output and the water inflow. This simplification means that we can use the<br />

same variable C for the controller output and the inflow (manipulated variable).<br />

H<br />

Q IN<br />

Figure 6-8<br />

Maximum inflow:<br />

0.01 m 3 /sec.<br />

1 m 2<br />

5m<br />

Q OUT<br />

Maximum<br />

level:<br />

5 meter<br />

Maximum outflow:<br />

0.01 m 3 /sec.


EIPINI Chapter 6: Process Control Page 6-8<br />

Therefore, if the controller output C is 0 %, the water inflow Q IN will immediately<br />

cease (Q IN = 0 m 3 /sec.) and when C is 100 %, the water inflow will immediately rise<br />

to its maximum value of Q IN = 0.01 m 3 /sec.<br />

The outflow demand or load Q OUT will be represented by the percentage variable L.<br />

L will be 0 % when the outflow Q OUT = 0 m 3 /sec and L will equal 100 % when the<br />

outflow is the maximum value of Q OUT = 0.01 m 3 /sec.<br />

The complete water level control system is shown in Figure 6-9.<br />

Inflow C (0% – 100%)<br />

100%<br />

Water inflow C<br />

Controller<br />

Output C<br />

100%<br />

Inflow<br />

valve<br />

C<br />

Level<br />

controller<br />

M<br />

Level<br />

detector<br />

Water level M<br />

(0% – 100%)<br />

Water supply<br />

Figure 6-9<br />

Water level<br />

control system<br />

Outflow L (0 – 100%)<br />

Before we can use our water tank, we must find out how it works. This means that<br />

we need to understand how the inflow and outflow influence the water level in the<br />

container. Clearly when C = L, the water level M will stay constant. When C > L, the<br />

water level will rise and when C < L, the water level will drop. It is therefore easy to<br />

Δ M<br />

argue that the rate at which the water level changes, , is proportional to C – L. In<br />

Δt<br />

the example given at the end of this chapter in Appendix 6-1, the exact relationship<br />

between M, C and L, was found to be:<br />

dM = 0.002×(C – L) percent per second.<br />

dt<br />

Finally, the error value will play a fundamental part in the operation of on-off and<br />

PID control. The percentage error value E is defined as:<br />

E = S – M ………….…………..………..…………………… Equation 6-1<br />

The set point is denoted by S and is expressed as a percentage of the maximum<br />

measured value. For example, a set point value S = 50 %, will imply a required water<br />

level of 2.5 m or half filled tank for the water level control system.<br />

6.2.1 ON-OFF OR TWO-POSITION CONTROL<br />

With on-off control mode (also called bang-bang control), the controller output C, can<br />

take on only two values, on or off. Referring to the level control in Figure 6-9, either


EIPINI Chapter 6: Process Control Page 6-9<br />

the inflow is switched on fully (100%) when the water level drops below a certain<br />

level, say 40%, or it is closed completely (0%), when the water level rises above a<br />

certain level, say 60%. With the aid of Equation 6-1, we could express the control law<br />

for the water level on-off control as:<br />

0% if E ≤ -10% (or M ≥ 60%)<br />

C = unchanged if -10% < E < 10% (or 60% > M > 40%)<br />

100% if E ≥ 10% (or M ≤ 40%)<br />

A possible response of the water level to on-off control as well as the controller<br />

action, is shown in Figure 6-10. On average, the water level remains 50%.<br />

Water level (M)<br />

60%<br />

40%<br />

Water inflow (C)<br />

100%<br />

Time<br />

0%<br />

Time<br />

Figure 6-10 Typical on/off control performance<br />

On-off control will also work well for a room temperature control – the heater is<br />

fully on, until the room temperature increases to a prescribed value, after which the<br />

heater is switched off, and so on. Not all systems are suited to on-off control. Imagine<br />

a speed cruise control in a car, set to 120 km/h, using on-off control. 118 km/h: full<br />

acceleration, 122 km/h: no throttle. Surely not very comfortable driving.<br />

A disadvantage of on-off control is the wear on the final control element, such as the<br />

control valve in the present example that continually moves from one extreme position<br />

to the other.<br />

Example 6-3<br />

A water container is 5 meter high and has a base area of 1 m 2 . The maximum water<br />

inflow rate, as well as the maximum water outflow rate, is 0.01 m 3 /sec. On-off control<br />

is used to regulate the water level in the tank, between 40% and 60% of its maximum<br />

level (5 meter). Assume that the rate of water outflow, is regulated and constant at<br />

60% of the maximum rate of outflow. Calculate:<br />

a) The time period that the inflow valve is closed<br />

b) The time period that the inflow valve is open.<br />

Solution:<br />

a) Water level when the inflow valve must close: 60% of 5 m = 0.6×5 = 3 meter.<br />

Water level when the inflow valve must open: 40% of 5 m = 0.4×5 = 2 meter.<br />

Constant demand from tank: 60% of 0.01 = 0.6×0.01 = 0.006 meter 3 /second.


EIPINI Chapter 6: Process Control Page 6-10<br />

The inflow valve closes when the<br />

water level reaches the 3 m mark, and<br />

stays closed until the water level<br />

drops to the 2 m mark. Total volume<br />

of water that must leave the container<br />

before the 2 m mark is reached:<br />

V = Ah = 1×1 = 1 m 3 .<br />

Rate of outflow Q, is 0.006 m 3 /s.<br />

But volume = flow rate×time<br />

∴V = Qt [(m 3 /sec)×(sec)]<br />

∴t closed = V/Q = 1/0.006<br />

= 166.7 sec = 2.778 min<br />

b) The inflow valve opens when the<br />

water level reaches the 2 m mark and<br />

stays open until the water level rises<br />

to the 3 m mark. Total volume of<br />

water that must enter the container, is<br />

1 m 3 . The resultant rate at which<br />

water flows into the tank is the rate of<br />

inflow, minus the outflow rate.<br />

Q = 0.01–0.006 = 0.004 m 3 /sec.<br />

∴t open = V/Q = 1/0.004<br />

= 250 sec = 4.167 min.<br />

5m<br />

Max inflow (100%) = 0.01 m 3 /s<br />

Min level=40%=2 m<br />

Controlled and<br />

1 m 2 constant outflow<br />

Outflow=60% of max (0.01) = 0.006 m 3 /s<br />

M<br />

60%<br />

40%<br />

C<br />

100%<br />

0%<br />

2.778<br />

min<br />

Max level=60%=3 m<br />

4.167<br />

min<br />

t<br />

t<br />

On-off control: A control strategy in which the controller output switches the final<br />

control element fully on or off to keep the controlled variable near set point.<br />

6.2.2 PROPORTIONAL CONTROL<br />

The proportional control strategy provides more smooth control than on-off control.<br />

The control effort is proportional to the magnitude of the error, which makes sense if<br />

one thinks about a person driving a car and trying to obey a 120 km/h speed limit. The<br />

more the speed falls below 120 km/h during an uphill climb, the more he will step on<br />

the accelerator, and when the speed tends to go above 120 km/h as he goes downhill,<br />

he will start to release the accelerator. When reaching 120 km/h on a straight road, he<br />

will keep the accelerator in just the right position to maintain his target speed of<br />

120 km/h.<br />

This is very important to understand, the driver does not release the accelerator<br />

completely when the target is reached. For proportional control, this control effort that<br />

is maintained when the error is zero, is called the bias. We will denote the bias by the<br />

percentage variable R. When designing a proportional controller, a fixed bias R must<br />

be chosen and this choice depends on the typical load on the system. For our water<br />

level control example, we will assume that the typical outflow demand is 50%, and a<br />

suitable choice for the bias, is R = 50%.


EIPINI Chapter 6: Process Control Page 6-11<br />

A simple and logical proportional control law for the water level system can now be<br />

formulated. The inflow equals 100% when the tank is empty, the inflow equals 50%<br />

(bias) when the tank is half full (equal to the desired value) and the inflow equals 0%<br />

when the tank is full (overflowing). This control strategy is depicted in Figure 6-11.<br />

Water inflow C (%)<br />

100 %<br />

50 %<br />

R (Bias)<br />

Figure 6-11<br />

Basic<br />

proportional<br />

control law<br />

This is obviously a reverse acting control strategy, because the controller output<br />

increases when the measured value decreases. A negative slope of the C-E line would<br />

indicate a direct acting control strategy. It is now a simple matter to write down an<br />

expression for this proportional control law, as the graph of C versus E, is a straight<br />

line of the form: y = mx + c. From Figure 6-11: C = E + 50. The slope of this line is<br />

clearly 1, because the controller output C, changes by 100% when the error E, changes<br />

by 100%. The slope of the line is called the proportional gain K P , of the controller.<br />

For the controller characteristic in Figure 6-11, the gain is therefore K P = 1.<br />

If the controller gain, K P , is changed, the slope of the straight line, describing the<br />

relationship between controller output C, and error value E, will also change. Graphs<br />

of C versus E, for three different controller gains, K P = 2, K P = 1 and K P = 0.5, are<br />

shown in Figure 6-12. (Note: Although a gain of 0.5, prescribes a theoretical error of<br />

100% before the output reaches 100 %, such an error could of course not occur in our<br />

water level control system.)<br />

100%<br />

-50 %<br />

Tank full,<br />

M = 100%<br />

C (%)<br />

100%<br />

0 % 50 %<br />

Tank half full,<br />

M = 50%<br />

C (%)<br />

Tank empty,<br />

M = 0%<br />

100%<br />

75%<br />

C (%)<br />

Error (%)<br />

(E=50-M)<br />

R=50%<br />

R=50%<br />

R=50%<br />

E(%)<br />

E(%)<br />

25%<br />

E(%)<br />

-25% 0% 25% -50% 0% 50% -100% -50% 0% 50% 100%<br />

K P = 2 K P = 1 K P = 0.5<br />

Figure 6-12 Proportional control with different proportional gains<br />

An alternative representation of the relationship between C and E, for K P = 2, K P = 1<br />

and K P = 0.5, is shown in Figure 6-13. The movement of the left hand side of a


EIPINI Chapter 6: Process Control Page 6-12<br />

Figure 6-13 Proportional band<br />

C E<br />

E<br />

0% 50%<br />

25%<br />

0<br />

-25%<br />

50%<br />

100%<br />

0<br />

-50%<br />

C<br />

0%<br />

50%<br />

100%<br />

0<br />

E<br />

100%<br />

C<br />

0%<br />

50%<br />

100%<br />

K P = 2<br />

K<br />

-100%<br />

P = 1 K P = 0.5<br />

pivoted beam, represents changes in the error E, while the right hand side, sweeps out<br />

corresponding values of the controller output C. For K P = 2, an error range of 50%<br />

(from -25% to 25%), sweeps out the complete output range of 100%. For K P = 1, an<br />

error range of 100% results in an output change of 100%. For K P = 0.5, the error value<br />

must change 200% to cover the whole output range. The error range that results in the<br />

total change in output range, is called the proportional band, of the controller, and is<br />

denoted by the percentage variable PB.<br />

For a high gain (narrowband) system (K P =2 in Figure 6-13 for example), a small<br />

error would cause a large reaction from the controller in his effort to correct the error.<br />

A low gain (wideband) system (K P = 0.5 in Figure 6-13 for example), would act more<br />

gently, because large errors, will cause a milder controller reaction. The relationship<br />

between proportional band PB (in percent) and proportional gain K P , is given by:<br />

PB =<br />

100 percent. …………...……………………….. Equation 6-2<br />

K<br />

P<br />

We can now obtain a general expression for the output C, of a proportional<br />

controller, in terms of the proportional band PB, the set point S, the measured value M<br />

and the bias R. In general, for any gain K P and bias R, the relationship between the<br />

controller output C, and the <strong>process</strong> error E, is a straight line of the form, y = mx + c,<br />

with the slope of the line equal to the controller gain K P and the y intersect, equal to<br />

the controller bias R.<br />

C (%)<br />

∴C = K P E + R ………………..………...….… (1)<br />

From Equation 6-1: E = S – M …...............….. (2)<br />

R (%)<br />

100 Slope=K P<br />

And from Equation 6-2: K P = ………..…. (3)<br />

PB<br />

(2) and (3) in (1):<br />

E (%)<br />

100<br />

C = K P E + R = (S – M) + R ……………………..…..…. Equation 6-3<br />

PB<br />

100<br />

In a typical application, S = 50 and R = 50, therefore C = (50 – M) + 50. PB


EIPINI Chapter 6: Process Control Page 6-13<br />

Example 6-4<br />

A <strong>process</strong> is controlled by a proportional controller. The controller is programmed for<br />

a positive gain (reverse acting controller) and proportional band of 80 %, a set point of<br />

50 % and a bias of 50 %. The system error is calculated from the difference between<br />

set point and the measured value i.e. E = 50 – M.<br />

a) Calculate the proportional gain of the controller.<br />

b) If the measured value of the <strong>process</strong> is indicated as 36 %, calculate the output of<br />

the controller at that instant.<br />

c) If the measured value of the <strong>process</strong> is indicated as 65 %, calculate the output of<br />

the controller at that instant.<br />

d) For the given settings, give the proportional control law (C as a function of E).<br />

e) Draw a graph of the controller control law (C versus E).<br />

a) From Equation 6-2: d) C = K P E + Bias<br />

K P = 100/PB = 100/80 = 1.25. ∴ C = 1.25E + 50<br />

b) From Equation 6-3: e)<br />

100<br />

C<br />

C = ×(S – M) + R<br />

100%<br />

ΔC<br />

PB<br />

Slope =<br />

ΔE<br />

100<br />

= ×(50 – 36) + 50 = 67.5%.<br />

100<br />

80<br />

=<br />

c) From Equation 6-3:<br />

50%<br />

80<br />

= 1.25<br />

100<br />

C = ×(S – M) + R<br />

PB<br />

E<br />

100<br />

= ×(50 – 65) + 50 = 31.25%.<br />

80<br />

-40% 0% 40%<br />

Example 6-5 (students must <strong>study</strong> this important example very carefully)<br />

The water level in a container is controlled by a proportional controller with K P = 1<br />

and bias R = 50%. At a given moment, the outflow demand L, is 50% and the inflow<br />

C, is 50%. The water level in the tank is also stable at its set point of 50% (the error is<br />

zero, therefore the inflow is 50% which is equal to the outflow). The outflow demand<br />

suddenly increases to L = 60%. Determine the new stable water level in the tank.<br />

Solution: The moment<br />

the outflow increases, the<br />

water level will start to<br />

drop, and it will continue<br />

to drop until the inflow C,<br />

also increases to 60%.<br />

The inflow is driven by<br />

the error value according<br />

to C = K P E + R = E + 50.<br />

C=50%<br />

L=50%<br />

S=50%<br />

M=50%<br />

C=60%<br />

S=50%<br />

L=60%<br />

M=40%<br />

Offset<br />

(E=10%)<br />

Therefore the controller output will reach 60% when E = 10%. This means, from<br />

E = 50–M, that the water level M, must drop to 40%. Although trying to keep the<br />

water level close to 50%, this example clearly illustrates that a proportional controller<br />

is not able to force the measured value back to set point, when the demand is


EIPINI Chapter 6: Process Control Page 6-14<br />

different from the bias value. Decreasing the proportional band (increasing K P ), will<br />

however keep the level closer to set point. For example, if K P = 2, the controller will<br />

need an error of 5 % to increase its output to 60 %. The water level will thus drop to<br />

only 45%.<br />

The difference between the set point and the measured value that may occur in<br />

proportional control, is called the offset, and it is a serious disadvantage of<br />

proportional only control. It is also important to note that we could have restored the<br />

level to 50 %, by changing the bias value of the controller to 60 %, if the outflow<br />

remained constant at 60 %.<br />

Example 6-6<br />

The water level in a container is controlled by a proportional controller and control<br />

valve. The controller gain is 1 and the bias is 50 %. Assume that the water inflow<br />

delivered by the valve is equal (no time delay) to the control signal C, and that the<br />

level detector instantly reflects the level M, of the water in the container. The<br />

relationship between the water level M, and the inflow C and outflow L, is given by:<br />

dM = 0.002×(C – L).<br />

dt<br />

a) Sketch the installation.<br />

b) Draw a block diagram of the system.<br />

c) The system is initially in the state: C = 50%, L = 50%, M = 50% and S = 50%.<br />

When time equals zero, the outflow demand L is suddenly increased to 60%.<br />

Determine a time expression for the water level M. (This will demonstrate the<br />

system’s response to a disturbance and is also called regulatory control action).<br />

a)<br />

S=50<br />

Inflow = C<br />

Inflow<br />

valve<br />

C<br />

Level<br />

controller<br />

M<br />

Level<br />

detector<br />

M<br />

b)<br />

S=50<br />

Water supply<br />

E=50-M<br />

K P =1 & R=50<br />

Control unit<br />

C=E+50<br />

C<br />

Control<br />

valve<br />

C<br />

L<br />

Outflow demand = L<br />

Water container<br />

dM =0.002(C–L)<br />

dt<br />

M<br />

M<br />

Controller<br />

Sensor<br />

c) The output of the proportional controller is given by:<br />

C = K P E + R = K P ×(S – M) + R = 1×(50 – M) + 50 = 100 – M.<br />

dM<br />

and = 0.002(C – L) = 0.002[(100-M) – 60] = 0.08 – 0.002M<br />

dt


EIPINI Chapter 6: Process Control Page 6-15<br />

dM<br />

∴ +0.002M = 0.08 and with M(0) = 50 gives the solution M = 40 + 10e<br />

-t/500<br />

dt<br />

A graph of the response of the water level M to the disturbance, is shown below.<br />

Demand L<br />

60 %<br />

50 %<br />

Time (sec.)<br />

Water level M<br />

50 %<br />

40 %<br />

t=0<br />

M = 40 + 10e -t/500<br />

43.7%<br />

t=500 sec.<br />

Offset<br />

Time (sec.)<br />

The behaviour of the water level may be compared with our findings in Example 6-5.<br />

Proportional control: A control strategy in which the controller output is<br />

proportional to the magnitude of the error. {C = K P E + R = (100/PB)×(S-M) + R}<br />

Proportional gain: Ratio of controller output change to error value change ΔC/ΔE.<br />

Proportional band: The error range that causes 100 % change in controller output.<br />

Offset: The steady state difference between the set point and the measured value.<br />

6.2.3 PROPORTIONAL AND INTEGRAL CONTROL<br />

Examples 6-5 and 6-6 illustrated very clearly the<br />

L<br />

basic weakness of proportional only control. When<br />

60%<br />

a disturbance causes an error, proportional control<br />

uses that same error to combat the disturbance. By<br />

50%<br />

its very nature, proportional control cannot M<br />

eliminate the offset caused by a disturbance. In the<br />

50%<br />

past, control engineers discovered that the offset<br />

40%<br />

could be eliminated by slowly changing the bias<br />

value up or down until the measured value is equal E<br />

to the desired value. This adjustment was called<br />

10%<br />

resetting the controller. In Example 6-5, for<br />

instance, shifting the bias to 60 % would reset the 0%<br />

level back to 50 %. This manual reset is in essence<br />

the same as automatic reset or integral reset. The C<br />

offset problem is highlighted again in Figure 6-14. 60%<br />

At instant A, the demand increases from 50% to 50%<br />

60%. With proportional only control, the level M<br />

drops to 40 %, the error E rises to 10 % and the<br />

inflow C, driven by the error value, rises to 60 %.<br />

Time<br />

Time<br />

Time<br />

A B Time<br />

Figure 6-14


EIPINI Chapter 6: Process Control Page 6-16<br />

The graphs from point B onwards, suggest a possible procedure with which we<br />

could eliminate the offset (perhaps someone gave us a hosepipe with running water<br />

and the task to restore the water level to 50 %). Firstly, we must keep the total inflow<br />

more than 60 % so that the level starts returning to 50 %. While we are doing this, we<br />

must remember that the error is getting smaller and we will get less and less assistance<br />

from the proportional controller. Secondly, after the level is restored to 50 %, the<br />

inflow from the proportional controller will only be 50 % and we need to supply the<br />

extra 10 % to equal the demand of 60 % so that the level remains at 50 %.<br />

Integral action gives an output proportional to the time integral of the error. This<br />

action can accomplish the two tasks mentioned above, namely, restoring the measured<br />

value to set point and providing the extra bias to match the demand. Integral control is<br />

rarely used alone but proportional control together with integral control, or PI control,<br />

is widely used in industry because it provides the important practical advantage of<br />

eliminating the offset. The controller output C, of a PI controller, can be expressed as:<br />

C = K P E + K I Edt + R, ……………….……………………… Equation 6-4<br />

∫<br />

where K P is the proportional gain, K I is defined as the integral gain and R is the bias.<br />

(Note: For PI control mode the bias term is optional and may be disabled.)<br />

Comparison of Integral Control with Proportional Control<br />

In Figure 6-15 the load change behaviour of the water level control system with<br />

proportional plus integral control, is compared with the behaviour of the same system<br />

with only proportional control (refer to Example 6-6). For this illustration, S = 50 %,<br />

K P = 1 and R = 50 % for both controllers. The integral gain K I of the PI controller,<br />

was chosen as K I = 0.0005, which is small and produces a very sluggish but easilly<br />

understandable response. The integral component is denoted by I = 0.0005<br />

∫ Edt .<br />

The fundamental difference between P control and PI control is that with P control,<br />

the error persists while with PI control, the error vanishes. With P control it is indeed<br />

the error that sustains the final increase in inflow of 10 % to satisfy the increased<br />

demand of 60 %. With PI control it is the area under the error curve, accumulated by<br />

the integral action I = K I∫ Edt , that sustains the final increase in inflow of 10 % to<br />

satisfy the increased demand of 60 %. In a sense, the final scenario resembles a P only<br />

controller that shifted its bias automatically from 50 % to 60 %. In contrast with P<br />

control, the magnitude of the error has no effect on the inflow in the end, because the<br />

error is erased by the integral action. It is however clear from Figure 6-15, that in the<br />

early stage after the disturbance, the error and ascociated proportional action is<br />

playing a critical part in assisting the integral action to provide the extra inflow needed<br />

to replace the water lost during the initial drop in water level and to restore the water<br />

level back to 50 %. For the integral action to carry this burden alone, would typically<br />

result in the water level oscillating around the set point and may very well be one of<br />

the reasons why integral action is rarely used without proportional action.


EIPINI Chapter 6: Process Control Page 6-17<br />

Proportional Control<br />

Demand L<br />

60%<br />

50%<br />

Water level M<br />

50%<br />

40%<br />

Error E=50-M<br />

10%<br />

0%<br />

Time<br />

Time<br />

Time<br />

Proportional plus Integral Control<br />

Demand L<br />

60%<br />

50%<br />

Water level M<br />

50%<br />

40%<br />

Error E=50-M<br />

10%<br />

0%<br />

I=0.0005<br />

∫ Edt<br />

Time<br />

Time<br />

Time<br />

10%<br />

0%<br />

Time<br />

Bias R<br />

50%<br />

Controller output<br />

C = E + R<br />

60%<br />

50%<br />

Time<br />

Time<br />

Bias R<br />

50%<br />

Controller output<br />

C = E + I + R<br />

60%<br />

50%<br />

Time<br />

Time<br />

Figure 6-15 P and PI control comparison<br />

Repeat time and reset time<br />

Many commercial PI controllers do not directly allow for the adjustment of the<br />

integral gain K I , but rather the adjustment of the repeat time or reset time. Imagine an<br />

error signal that suddenly changes from 0 to 10 % and remains constant after that, as<br />

shown in Figure 6-16.<br />

Error E<br />

Area = 10×T R<br />

10%<br />

0%<br />

Time<br />

PI controller output<br />

C=K P E+K I<br />

∫ Edt<br />

Integral effect = K I<br />

∫ Edt = K I∫ 10dt = 10K IT R<br />

Proportional effect = K P ×E = K P ×10<br />

50%<br />

Time<br />

T<br />

Figure 6-16<br />

R<br />

(Repeat or reset time)<br />

Reset time<br />

If this step in error signal is presented to the input of a PI controller, that initially


EIPINI Chapter 6: Process Control Page 6-18<br />

had an output of 50 %, there would be a sudden increase in controller output because<br />

the proportional term K P E tracks the error instantaneously. Furthermore, as the<br />

integral term, K I ∫ Edt accumulates the area under the error curve, it wil gradually add<br />

to the output signal, in a linear fashion. There will come a time when the contribution<br />

of the integral term, equals the proportional contribution. The time it takes for this to<br />

happen is called the repeat time, as it indicates the time it takes the integral action to<br />

‘repeat’ the action of the proportional action.<br />

It is immediately clear from Figure 6-16 that the integral and proportional effects are<br />

equal when 10K I T R = 10K P or K I T R = K P . The repeat/reset time is therefore given by:<br />

T R =<br />

51<br />

K<br />

P<br />

K<br />

I<br />

or K I =<br />

K<br />

P ………..……….………..………….. Equation 6-5<br />

T<br />

R<br />

Controller tuning<br />

In a practical situation, different values for K P and K I would be chosen for the water<br />

level control system so that the water level would be corrected much faster after a<br />

disturbance. Choosing suitable values for K P and K I is called tuning of the controller.<br />

Guidelines that help with controller tuning do exist and even self tuning systems are<br />

available. One such rule of thumb suggests the following settings for a liquid level<br />

control: PB < 100 % and reset time of 10 minutes. If we should choose a PB of 80 %<br />

(K P = 1.25), then a reset time of 10 minutes (600 sec.) would translate into an integral<br />

gain of K I = 0.002083 (from Equation 6-5: K I = K P /T R = 1.25/600). Using these values<br />

for K P and K I , the plots of the water level M and controller output C, shown in<br />

Figure 6-17, were obtained with Matlab. The water level M drops to 45% and returns<br />

to setpoint within 25 minutes.<br />

M(%)<br />

C(%)<br />

70<br />

50<br />

49<br />

48<br />

65<br />

60<br />

47<br />

46<br />

55<br />

45<br />

0 500 1000 1500 2000 2500 3000 3500 4000<br />

50<br />

0 500 1000 1500 2000 2500 3000 3500 4000<br />

t (sec)<br />

t (sec)<br />

a) Water level Figure 6-17 b) Controller output<br />

System response to demand change from 50% to 60% with PI control (K P =1.25, K I =0.002083)<br />

Integral control: A control strategy in which the controller output is proportional<br />

to the integral of the error.<br />

{For PI control: C = K P E + K I<br />

∫ Edt + R = K ⎛<br />

⎞<br />

⎜ 1 ⎟<br />

P E + Edt + R}<br />

⎜ T ∫ ⎟<br />

⎝ R ⎠<br />

Reset or repeat time: Time taken for the integral control action to equal the<br />

proportional control action under the influence of a constant error. {T R = K P /K I }


EIPINI Chapter 6: Process Control Page 6-19<br />

6.2.4 PROPORTIONAL AND DERIVATIVE CONTROL<br />

Derivative or differential control action (also called rate control or pre-act control)<br />

gives an output which is proportional to the derivative of the error. Because derivative<br />

action depends on the slope of the error curve, it has no effect if the error is constant.<br />

Derivative control action therefore only adds to the controller output when the error<br />

changes, that is when the set point or the measured value changes. Derivative control<br />

can thus not be used alone because the controller would not combat the error if it is not<br />

changing, even when the error is large. Derivative control action is usually found in<br />

combination with proportional control to form a PD controller. The output of a PD<br />

controller can be expressed as:<br />

dE<br />

C = K P E + K D + R, …………….…………….…………… Equation 6-6<br />

dt<br />

where K D is defined as the derivative gain. The basic effect of derivative control is to<br />

oppose any change in the error value (if the water level M, in our example, would<br />

drop, dM/dt would be negative and dE/dt = d/dt(S-M) would be positive and the<br />

inflow C would immediately increase). As such, derivative action will always have a<br />

stabilizing effect on the system by combating aggressive and unstable behaviour.<br />

Derivative control reacts to changes in the error and starts with corrective action much<br />

earlier than the proportional (and integral) action. Derivative action was in fact named<br />

‘pre-act’ because of its ability to ‘anticipate’ the future movement of the measured<br />

variable and the direction in which the <strong>process</strong> is heading. A PD controller, however,<br />

does not have the capability to eliminate the offset and will also add to the wear on the<br />

final control element such as a control valve.<br />

Rate time<br />

Commercial PD controllers normally do not directly allow for the adjustment of the<br />

derivative gain K D , but rather the adjustment of the rate time. Imagine a linear error<br />

signal that gradually changes from 0 to 10 % in T D seconds, as shown in Figure 6-18.<br />

Error E<br />

Figure 6-18<br />

10%<br />

Rate time<br />

10<br />

0%<br />

Time<br />

T D<br />

PD controller output<br />

Proportional effect = K<br />

dE<br />

P ×E = K P ×10<br />

C=K P E+K D<br />

dt<br />

dE 10<br />

Derivative effect = K D × = KD ×<br />

50%<br />

dt T<br />

D<br />

Time<br />

T D<br />

If this ramp error signal is presented to the input of a PD controller, that initially had<br />

an output of 50 %, there would be a sudden increase in controller output because the<br />

dE<br />

derivative term, K D , immediately takes on a value proportional to the slope of the<br />

dt<br />

error curve. Furthermore, as the proportional term, K P E, follows the error value, it will<br />

increasingly add to the output signal, in a linear fashion. The time, T D , that it takes the


EIPINI Chapter 6: Process Control Page 6-20<br />

contribution of the proportional term to equal the derivative contribution is called the<br />

rate time, and it gives an indication of how much faster the derivative action is<br />

compared to the proportional action.<br />

It is immediately clear from Figure 6-18 that 10×K I T R = 10×K P = 10×K D /T D or<br />

K P T D = K D . The rate time is therefore given by:<br />

K<br />

T D =<br />

D<br />

or K D = K P T D ……..……….…..………………….. Equation 6-7<br />

K<br />

P<br />

Derivative control: A control strategy in which the controller output is<br />

proportional to the derivative of the error.<br />

dE ⎛ dE ⎞<br />

{For PD control: C = K P E + K D + R = KP ⎜E<br />

+ T ⎟ + R }<br />

dt<br />

D dt ⎠<br />

Rate time: Time taken for proportional action to equal derivative action under the<br />

influence of a linearly changing error. {T D = K D /K P }<br />

6.2.5 PROPORTIONAL, INTEGRAL AND DERIVATIVE CONTROL<br />

A proportional, integral and derivative controller (PID controller), gives an output that<br />

is determined by proportional action, integral action and derivative action.<br />

Mathematically the output of a PID controller can be expressed as:<br />

C = K P E + K I<br />

∫ Edt + dE KD + R. …………….……………… Equation 6-8<br />

dt<br />

Controller tuning<br />

Adding integral and derivative action to a controller increases the complexity of the<br />

system. Choosing suitable P, I and D gains for a PID controller (tuning the controller),<br />

becomes complicated. Even for the very simple water level control system in Figure<br />

6-9, analysis of the complete PID system is difficult. We will however conclude with<br />

one example of PID control of the water level control system. The controller gains<br />

chosen were K P = 5, K I = 0.08 (reset time = 1.04 min.) and K D = 500 (rate time = 1.7<br />

min.). The plots in Figure 6-19, were obtained with Matlab. The response curves in<br />

Figure 6-19, show the water level (although oscillating), not even dropping below<br />

49% and returning back to setpoint in less than seven minutes.<br />

M(%)<br />

50.6<br />

50.4<br />

50.2<br />

C(%)<br />

65<br />

⎝<br />

50<br />

49.8<br />

60<br />

49.6<br />

49.4<br />

49.2<br />

0 500 1000 1500 2000 2500 3000 3500 4000<br />

a) Water level<br />

Figure 6-19<br />

t (sec)<br />

55<br />

0 500 1000 1500 2000 2500 3000 3500 4000<br />

b) Controller output<br />

System response to demand change from 50 % to 60 %<br />

with PID control (K P = 5, K I = 0.08 and K D = 500)<br />

t (sec)


6.3 PID CONTROLLERS<br />

6.3.1 Pneumatic PID controller<br />

One may justifiably ask<br />

whether pneumatic controllers<br />

still have a place in a digital<br />

age. The answer is yes and<br />

some of the reasons that can be<br />

put forward are familiarity,<br />

simplicity, a large installed<br />

base, no fear of electricity and<br />

above all, they are inherently<br />

explosion-proof.<br />

A typical front panel of a<br />

pneumatic controller is shown<br />

in Figure 6-20. Depending on<br />

the manufacturer, controllers<br />

will have slightly different<br />

configurations of pushbuttons,<br />

meters and switches.<br />

EIPINI Chapter 6: Process Control Page 6-21<br />

Manual<br />

pushbutton<br />

(decrease<br />

output)<br />

Manual<br />

button with<br />

indicating<br />

light<br />

MV<br />

0<br />

CLOSE<br />

MANUAL<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0 SP<br />

OUTPUT<br />

50<br />

100<br />

OPEN<br />

AUTOMATIC<br />

Set point<br />

thumbwheel<br />

Manual<br />

pushbutton<br />

(increase<br />

output)<br />

Automatic<br />

button with<br />

indicating<br />

light<br />

Figure 6-20<br />

Pneumatic implementation of the PID function is achieved in a variety of ways using<br />

diaphragms and valves in different mechanical designs. One possible arrangement<br />

revolves around the force balance principle utilizing a flapper-nozzle and bellows<br />

assembly. The pressure variable that represents the measured value is normally<br />

connected to the force bar (beam) by means of a bourdon tube or a bellows element.<br />

The set point value, the bias value and the proportional band settings are adjusted in<br />

several ways by different manufacturers using various mechanical or pneumatic<br />

techniques.<br />

i) Proportional control<br />

The controller arrangement for proportional control is shown in Figure 6-21. For the<br />

purpose of this illustration, we will assume that all variables are applied to the force<br />

balance system by means of bellows elements and that the pivot is positioned at a<br />

distance ‘a’ units from the S/M bellows and at a distance ‘b’ units from the C/R<br />

bellows. The position of the proportional bellows in Figure 6-21 will result in reverse<br />

acting control because a decrease in measured value will cause the flapper to move<br />

towards the nozzle and balance will be restored at a higher controller output pressure.<br />

Given that the flapper and nozzle will always maintain equilibrium, the controller<br />

operation can very easily be described by equating moments around the pivot point,<br />

assuming that all the bellows elements have a cross-sectional area of one unit.<br />

a<br />

a×S + b×R = a×M + b×C ⇒ bC = a(S – M) + bR ⇒ C = (S – M) + R<br />

b<br />

Using E = S-M and K P = (a/b), it follows that C = K P E + R.<br />

The device therefore provides a proportional control output with the proportional<br />

gain determined by the position of the pivot point.


EIPINI Chapter 6: Process Control Page 6-22<br />

Set point S<br />

(20-100 kPa)<br />

50% = 60 kPa<br />

Pilot<br />

relay<br />

Restriction<br />

Controller<br />

output C<br />

Air<br />

supply<br />

S<br />

Beam<br />

Proportional<br />

bellows<br />

C<br />

Measured<br />

value M<br />

(20-100 kPa)<br />

M<br />

a<br />

Pivot<br />

b<br />

Reset<br />

bellows<br />

R<br />

Bias value R<br />

(20-100 kPa)<br />

Flapper and<br />

nozzle<br />

Figure 6-21<br />

Pneumatic<br />

proportional<br />

controller<br />

ii) Proportional plus integral control<br />

With proportional control, the fixed bias value R that is applied to the reset bellows, is<br />

called ‘manual reset’ because one way to reset the controlled variable back to set point<br />

in a proportional only control system, is to manually change the bias value. In order to<br />

change the bias value automatically with the aim to reset the controlled variable<br />

automatically to set point, the reset bellows must now obtain its pressure from the<br />

controller output pressure C. To achieve automatic reset the output pressure must be<br />

applied via an adjustable needle valve restriction, as shown in Figure 6-22.<br />

Set point S<br />

(20-100 kPa)<br />

50% = 60 kPa<br />

C<br />

Pilot<br />

relay<br />

Restriction<br />

Controller<br />

output C<br />

Air<br />

supply<br />

Measured<br />

value M<br />

(20-100 kPa)<br />

S<br />

M<br />

Beam<br />

a<br />

C<br />

Pivot<br />

Reset time<br />

adjustment<br />

(r)<br />

Proportional<br />

bellows<br />

b<br />

Reset<br />

bellows<br />

Needle<br />

valve<br />

R<br />

C<br />

R<br />

Flapper and<br />

nozzle<br />

Automatic reset R<br />

Figure 6-22<br />

Pneumatic<br />

proportional<br />

plus<br />

integral<br />

controller


EIPINI Chapter 6: Process Control Page 6-23<br />

The restriction will prevent the pressure R in the reset bellows to immediately<br />

follow the output pressure C and causes a delay in the pressure build up (or decay)<br />

in the reset bellows which will depend on the needle valve setting. This delay in<br />

reset pressure will cause integral action to be added to the controller and it can be<br />

a ra<br />

shown that the controller output is C = (S – M) + (S-<br />

M)dt . Using E = S–M,<br />

b b<br />

K P = b<br />

a and KI = b<br />

ra , this equation then becomes C = KP E + K I<br />

∫ Edt .<br />

The controller thus provides a proportional plus integral output. The dynamic<br />

integral term effectively replaces the fixed bias with an ‘automatic bias’. Commercial<br />

controllers do however allow a fixed bias to be enabled for PI control, if so required.<br />

iii) Proportional plus derivative control<br />

Derivative action may be added to the proportional controller in Figure 6-21, by<br />

inserting a variable restriction in the pressure line that feeds the proportional bellows,<br />

as shown in Figure 6-23. The proportional bellows now receives a pressure P d that is a<br />

delayed version of the output pressure C.<br />

∫<br />

P d<br />

Needle valve<br />

C<br />

Set point S<br />

(20-100 kPa)<br />

50% = 60 kPa<br />

Rate time<br />

adjustment<br />

(d)<br />

Pilot<br />

relay<br />

Restriction<br />

Controller<br />

output C<br />

Air<br />

supply<br />

S<br />

Beam<br />

Proportional<br />

bellows<br />

P D<br />

Measured<br />

value M<br />

(20-100 kPa)<br />

M<br />

a<br />

Pivot<br />

b<br />

Reset<br />

bellows<br />

R<br />

Bias value R<br />

(20-100 kPa)<br />

Flapper and<br />

nozzle<br />

Figure 6-23<br />

Pneumatic<br />

proportional<br />

plus<br />

derivative<br />

controller<br />

Again it can be shown that for the configuration in Figure 6-23, the controller output<br />

a a d(S- M)<br />

a a<br />

is given by C = (S – M) + + R. Using E = S - M, K P = and KD = ,<br />

b bd dt<br />

b bd<br />

dE<br />

this equation then becomes C = K P E + K D + R. dt<br />

It is now clear that inserting a delay in the proportional pressure line, causes<br />

derivative action to be added to the proportional action.<br />

Adding delay actions simultaneously in both the reset and the feedback line, will<br />

result in a complete PID controller and such a structure is shown in Appendix 6-2.


EIPINI Chapter 6: Process Control Page 6-24<br />

6.3.2 Electronic PID controller<br />

Figure 6-24 shows an electronic implementation of the basic PID function. The first<br />

operational amplifier generates the error signal from the reference S and the measured<br />

value M. The error signal is fed to the P, I and D blocks to create the three PID<br />

functions which are added together to form the final PID control function C.<br />

R PF<br />

Proportional<br />

action<br />

E<br />

R PI<br />

Op-amp<br />

V P<br />

100k<br />

Gnd<br />

R<br />

V P = - PF E<br />

R<br />

PI<br />

= -K P E<br />

Error<br />

M<br />

S<br />

100k<br />

100k<br />

100k<br />

E<br />

Op-amp<br />

E=S-M<br />

100k<br />

Gnd<br />

E<br />

R I<br />

C I<br />

Op-amp<br />

Gnd<br />

Integral<br />

action<br />

V I<br />

100k<br />

100k<br />

Op-amp<br />

1<br />

V I = -<br />

R C ∫ Edt<br />

I I<br />

= -K I<br />

∫ Edt Gnd<br />

Adder<br />

C<br />

C = - V P – V I – V D<br />

E<br />

C D<br />

R D<br />

Op-amp<br />

Derivative<br />

action<br />

V D<br />

100k<br />

= K P E<br />

P<br />

+ K I<br />

∫ Edt<br />

I<br />

dE<br />

+ K D<br />

dt<br />

D<br />

Gnd<br />

dE<br />

V D = -R D C D<br />

dt<br />

dE<br />

= -K D<br />

dt<br />

Figure 6-24<br />

Electronic PID<br />

controller


EIPINI Chapter 6: Process Control Page 6-25<br />

6.3.3 Digital PID controller<br />

Digital or software implementation of the PID control algorithm uses a digital<br />

computer to calculate the control action as shown in Figure 6-25. Computers are<br />

discrete machines and not capable of calculating the control action on a continuous<br />

basis as the analog structure in Figure 6-24. It will instead try to approximate Equation<br />

6-8, using numerical methods. Therefore the measured value M must be sampled with<br />

sufficiently small sampling period T s and digitised with an analog to digital converter<br />

to deliver the sampled values M 0 , M 1 , M 2 , M 3 , M 4 , etc. to the computer at time instants<br />

0, 1, 2, 3, 4. At each time instant, the proportional, integral and derivative component<br />

must be calculated and added together to form the sequence of corresponding<br />

controller outputs C 0 , C 1 , C 2 , C 3 , C 4 , …..<br />

M<br />

T S<br />

M 0<br />

Analog to digital<br />

converter<br />

M 1 M 2<br />

Computer<br />

Figure 6-25<br />

Digital to analog<br />

converter<br />

C C 3<br />

2<br />

M 3 M C 1 4<br />

C 0<br />

For instance, referring to Figure 6-26, at time instant 4, just after M 4 was sampled,<br />

the error E 4 may be obtained from 50–M 4 , given that the set point is 50%. From E 4 the<br />

proportional action can be calculated as P 4 = K P ×E 4 . If we assume that the controller<br />

started its operation at time equals zero, then a very simple method to calculate the<br />

integral component I 4 is to approximate the area A, under the error curve with<br />

4<br />

rectangles to obtain A=<br />

∫ × T<br />

S Edt ≈ T s ×E 0 + T s ×E 1 + T s ×E 2 + T s ×E 3 . The integral<br />

0<br />

action I 4 may then be expressed as I 4 = K I ×(T s ×E 0 + T s ×E 1 + T s ×E 2 + T s ×E 3 ) or<br />

I 4 =K I ×T s ×(E 0 + E 1 +E 2 +E 3 ). The derivative component D 4 , depends on the slope of the<br />

error curve at time instant 4, and may be approximated with D 4 = K D<br />

M<br />

S=50%<br />

E<br />

E 2<br />

E 3<br />

E<br />

T S<br />

E − E<br />

4 3<br />

T<br />

s<br />

C<br />

C 4<br />

E 0<br />

E 1 E 1<br />

E<br />

M 2<br />

0<br />

E 3 E 4 E<br />

M 0<br />

1<br />

M 2 M 3 M 4<br />

0 1 2 3 4 Time 0 1 2 3 4<br />

T s T s T s T s T s<br />

T s<br />

Figure 6-26<br />

E 3 ×T s E 4<br />

E 2 ×T s<br />

E 3<br />

E 1 ×T s<br />

E 0 ×T s<br />

T s T s<br />

Time<br />

C 4 = K P E 4 + K I ×T S ×(E 0 + E 1 + E 2 + E 3 ) +<br />

P<br />

I<br />

0 1 2 3 4 Time<br />

T s T s<br />

T s<br />

E − E<br />

K D × 4 3<br />

T<br />

s<br />

D<br />

T s


EIPINI Chapter 6: Process Control Page 6-26<br />

6.4 CONTROL VALVES<br />

Control valves are used to regulate the flow rate of a medium and<br />

serve as the correcting element in many control systems.<br />

A pneumatic control valve consists of a diaphragm actuator and a<br />

valve body with plug and seat through which the fluid is<br />

regulated. The principle of operation of a control valve is very<br />

simple. A diaphragm is bolted to a dished metal head to form a<br />

pressure tight compartment. A pneumatic signal is applied to this<br />

compartment to move the diaphragm which is opposed by a<br />

spring. Attached to the diaphragm is the valve stem and plug so<br />

that any movement of the diaphragm, results in a corresponding<br />

movement of the plug, thereby controlling the flow. A typical<br />

reverse acting control valve assisted by a valve positioner (section<br />

6.4.4) to enhance valve operation, is illustrated in Figure 6-28.<br />

6.4.1 Actuators<br />

Actuators may be categorized<br />

as ‘direct acting’ or ‘reverse<br />

acting’ and some configurations<br />

are indicated in Figure 6-27. In<br />

a reverse acting actuator an<br />

increase in the pneumatic<br />

pressure applied to the<br />

diaphragm lifts the valve stem<br />

(in a normally seated valve this<br />

will open the valve and is<br />

called ‘air to open’). In a directacting<br />

actuator, an increase in<br />

the pneumatic pressure applied<br />

to the diaphragm extends the<br />

valve stem (for a normally<br />

seated valve this will close the<br />

valve and is called ‘air to<br />

close’). The choice of valve<br />

action is dictated by safety<br />

considerations. In one case it<br />

may be desirable to have the<br />

valve fail fully open when the<br />

pneumatic supply fails. In<br />

another application it may be<br />

considered better if the valve<br />

fails fully shut.<br />

Reverse acting<br />

Direct acting<br />

Figure 6-27


EIPINI Chapter 6: Process Control Page 6-27<br />

Spring tension adjust<br />

(Valve closed at 20 kPa)<br />

Spring<br />

Diaphragm<br />

plate<br />

Diaphragm<br />

Actuator<br />

(Motor)<br />

Air<br />

supply<br />

Output<br />

Open<br />

½<br />

Close<br />

Stem connector<br />

(stroke adjust<br />

20 – 100 kPa)<br />

Travel<br />

indicator<br />

Stem<br />

Valve<br />

positioner<br />

Connector<br />

arm<br />

Yoke<br />

Instrument<br />

signal<br />

(20-100 kPa)<br />

Gland and<br />

packing<br />

Bonnet (yoke<br />

hold down) nut<br />

Bonnet<br />

Gasket<br />

Plug<br />

Seat<br />

Valve<br />

body<br />

Figure 6-28<br />

Reverse acting (air to open) pneumatic control valve


EIPINI Chapter 6: Process Control Page 6-28<br />

6.4.2 Valve types<br />

The essential function of a valve is to start, throttle and stop a fluid flow. Throttling,<br />

as shown in Figure 6-29, means regulating or controlling the rate of fluid flow. Globe<br />

valves, gate valves, needle valves, pinch valves and diaphragm<br />

valves use a linear stem movement to operate. A gate valve is<br />

illustrated in Figure 6-30 (a) as an example of linear stem<br />

movement. Ball valves, plug valves and butterfly valves function<br />

with a 90° rotational movement and the operation of a ball valve is<br />

indicated in Figure 6-30 (b). The valve components inside the<br />

valve body are collectively referred to as the trimming.<br />

Closed Throttling Open<br />

Figure 6-29<br />

Ball valve<br />

Gate valve<br />

Figure 6-30 (a)<br />

Linear valve movement<br />

Figure 6-30 (b)<br />

Rotational valve movement


EIPINI Chapter 6: Process Control Page 6-29<br />

1. Globe valves<br />

Globe valves are the most widely used in industry for flow control in both on/off and<br />

throttling service. They typically have rounded bodies from<br />

which their name is derived. They are linear motion valves<br />

with a tapered plug or disk attached to the stem that closes onto<br />

a seating surface to act as flow control element, as shown in<br />

Figure 6-31. The liquid must make two 90° turns when passing<br />

through the valve and because of that, the pressure drop in the<br />

globe valve is significant, even when fully open. The high Figure 6-31<br />

pressure drop is the main disadvantage of the globe valve.<br />

Globe valve<br />

2. Gate valves<br />

Gate valves (also known as knife valves or slide valves) are the most common valve<br />

used for on/off service but could be used for throttling. They are linear motion valves<br />

and a flat disk or wedge slides into the flow stream<br />

to act as flow control element, as shown in Figure<br />

6-32. The direction of fluid flow is not changed by<br />

the valve. When fully open, the wedge completely<br />

clears the flow path creating minimum pressure drop.<br />

Gate valves are advantageous in applications<br />

involving slurries, as their ‘gates’ can cut right<br />

through the slurry. They are also used in applications<br />

that involve viscous liquids such as heavy oils, light<br />

grease, varnish, molasses, honey and cream.<br />

3. Needle valves<br />

A needle valve, shown in Figure 6-33, is used to make relatively<br />

fine adjustments to the fluid flow and can be used for on/off and<br />

throttling service. They are linear motion valves with a tapered<br />

‘needle like’ cone shaped plug that acts as the flow control<br />

element. The fluid enters from below into the seat which forms<br />

part of the flow path. The needle plug permits the flow opening<br />

in this channel to be increased or decreased very gradually.<br />

Needle valves are widely used for steam, air, gas, water or other<br />

non-viscous liquids. Disadvantages of the needle valve are a<br />

large pressure loss and possible clogging of the flow orifice.<br />

4. Pinch valves<br />

The relatively inexpensive pinch valve, illustrated<br />

In Figure 6-34, is the simplest of all the valve<br />

designs and may be used for on/off and throttling<br />

service. Pinch valves are linear motion valves<br />

that use a flexible tube or sleeve to effect valve<br />

closure. Pinch valves are ideally suited for the<br />

handling of slurries, liquids with large amounts<br />

of suspended solids and corrosive chemicals.<br />

Parallel<br />

disk<br />

Wedge<br />

Knife<br />

Figure 6-32<br />

Gate valve<br />

Figure 6-33<br />

Needle valve<br />

Figure 6-34<br />

Pinch valve


EIPINI Chapter 6: Process Control Page 6-30<br />

5. Diaphragm valves<br />

A diaphragm valve, illustrated in Figure 6-35, is related to pinch<br />

valves and one of the oldest types of valve known. Leather<br />

diaphragm valves were used by Greeks and Romans to control<br />

the temperature of their hot baths. Diaphragm valves use linear<br />

stem movement for on/off and throttling and are excellent for<br />

controlling fluid flow containing suspended solids. A stud moves<br />

a flexible and resilient diaphragm into the flow thereby acting as<br />

flow control element. The diaphragm valve is used primarily for<br />

handling viscous and corrosive fluids as well as slurries.<br />

6. Plug valves<br />

Plug valves, illustrated<br />

in Figure 6-36, are also<br />

called plug cock or stop<br />

cock valves and they<br />

also date back to ancient<br />

Close Throttle Open<br />

times, when used by Romans in plumbing systems. Today they<br />

remain one of the most widely used valves for both on/off and<br />

throttling services. A plug valve is a rotary moving valve that uses a cylindrical or<br />

tapered plug as the flow control element. The opening through the plug may be<br />

rectangular or round. Flow is regulated from closed to open during a 90° turn. If the<br />

opening is the same size or larger than the pipe’s inside diameter, it is referred to as a<br />

full port otherwise as standard round port. Full port plug valves offer little resistance<br />

to flow when fully open, resulting in small pressure loss. A disadvantage off plug<br />

valves is its poor throttling characteristics.<br />

7. Ball valves<br />

Ball valves, illustrated in Figure 6-37, are related to plug valves<br />

and are used in situations where tight shut-off is required. Ball<br />

valves are rotational motion valves and are used for on/off and<br />

throttling service. The flow control element is a sphere with a<br />

round opening rotating in a spherical seat. Flow is regulated<br />

from closed to open during a 90° turn. Full port ball valves<br />

create a minimum pressure loss when fully open. A disadvantage<br />

of the ball valve is its poor throttling characteristics.<br />

8. Butterfly valves<br />

A butterfly valve, illustrated in Figure 6-38, is a rotary<br />

movement valve used for on/off flow control and especially<br />

in throttling applications. The flow control element in<br />

butterfly valves is a circular disk with its pivot axis at right<br />

angles to the direction of flow. A 90° turn of the axis moves<br />

the valve from closed to open. Butterfly valves are well<br />

suited to handle large flows of liquids or gasses as well as<br />

slurries and liquids with large amounts of suspended solids.<br />

Figure 6-35<br />

Diaphragm valve<br />

Figure 6-36<br />

Plug valve<br />

Figure 6-37<br />

Ball valve<br />

Figure 6-38<br />

Butterfly valve


EIPINI Chapter 6: Process Control Page 6-31<br />

6.4.3 Valve characteristics<br />

In essence a valve simply functions as a restriction in a flow line,<br />

with flow areas A 1 and A 2 and pressure differential p 1 - p 2 , as shown<br />

in Figure 6-39. As such, the flow equation (Equation 3-8(e),<br />

Chapter 3) will in principle describe the flow rate of a fluid through<br />

a valve. The form of this equation used by valve suppliers is:<br />

q<br />

p 1<br />

p2<br />

A 1<br />

A 2<br />

Figure 6-39<br />

q = C<br />

Δp<br />

G<br />

Equation 6-9<br />

where C is called the valve flow coefficient, q the flow rate of the liquid through the<br />

valve, Δp = p 1 - p 2 the pressure difference across the valve and G the specific gravity<br />

(relative density) of the fluid. For historical reasons, the flow rate q, in Equation 6-9, is<br />

measured in gallons per minute (gpm) and the pressure difference Δp, in pounds per<br />

square inch (psi). When we compare Equations 3-8(e) and 6-9, it becomes clear that C<br />

is fundamentally determined by the effective flow areas A 1 and A 2 of the valve. The<br />

value of C will therefore change from zero (when the valve is fully closed) to a<br />

maximum value (when the valve is fully opened).<br />

Assume that a valve is opened a certain amount x (with x denoting the fractional<br />

valve opening or valve travel between 0 and 1). The flow rate q, of the liquid (we will<br />

assume water with G = 1 in Equation 6-9), flowing through the valve, is now allowed<br />

to increase until the pressure Δp across the valve, reaches a maximum of say 4 psi. We<br />

may conceivably expect the results depicted in Figure 6-40, shown for x = 0 (valve<br />

fully closed), x = 0.25 (valve one quarter open), x = 0.5 (valve half open), x = 0.75<br />

(valve three quarters open) and x = 1 (valve fully open).<br />

q<br />

0% open 25% open 50% open 75% open 100% open<br />

x = 0 x = ¼ x = ½ x = ¾ x = 1<br />

q q q q<br />

Δp<br />

Δp<br />

Δp<br />

Δp<br />

Δp<br />

q=0×<br />

Δ p<br />

q=10×<br />

Δ p<br />

q=20×<br />

Δp<br />

q=30×<br />

Δ p<br />

q=40×<br />

Δp<br />

(C=0)<br />

(C=10)<br />

(C=20)<br />

(C=30)<br />

(C=40)<br />

q (gpm)<br />

q (gpm)<br />

q (gpm)<br />

q (gpm)<br />

q (gpm)<br />

80<br />

60<br />

40<br />

20<br />

0<br />

4<br />

Δp<br />

psi<br />

0<br />

4<br />

Δp<br />

psi<br />

0<br />

4<br />

Δp<br />

psi<br />

0<br />

4<br />

Δp<br />

psi<br />

0<br />

4<br />

Δp<br />

psi<br />

Figure 6-40


EIPINI Chapter 6: Process Control Page 6-32<br />

The special value of C that corresponds to the valve fully open, is called the<br />

characteristic flow coefficient of the valve and denoted by C V . The characteristic flow<br />

coefficient C V , is an extremely important parameter of a specific valve, and is used<br />

extensively when choosing the correct valve for a certain application. For the valve<br />

characteristics in Figure 6-40, for example, C V = 40.<br />

The way q varies when the pressure is kept constant (normally the rated pressure<br />

drop across the valve for maximum flow) and the way C changes as the valve opening<br />

x changes, may be available for a particular valve, provided by the manufacturer in<br />

tabulated format. For example, the graph of q (with Δp = 4 psi) as a function of x and<br />

the graph of C as a function of x, are shown in Figure 6-41 (a) and (b) respectively, for<br />

the flow characteristics of our hypothetical valve, depicted in Figure 6-40.<br />

Figure 6-41(a)<br />

Figure 6-41(b)<br />

q (gpm)<br />

C<br />

80<br />

40<br />

60<br />

40<br />

30<br />

20<br />

Valve<br />

Valve<br />

20<br />

10<br />

opening<br />

opening<br />

x<br />

x<br />

0<br />

0<br />

¼ ½ ¾ 1<br />

¼ ½ ¾ 1<br />

(0%) (25%) (50%) (75%) (100%)<br />

(0%) (25%) (50%) (75%) (100%)<br />

The different values of C obtained as the valve travels through its full range (or<br />

stroke as it is also called) from closed to open, is generally expressed in the following<br />

format, by valve manufacturers:<br />

C = C V ×f(x) Equation 6-10<br />

where f(x) is called the inherent valve characteristic of a particular valve. For<br />

example, from the graph of C, in Figure 6-41(b), we could express C as:<br />

C = 40×x,<br />

and we conclude therefore from Equation 6-10, that C V = 40 and f(x) = x.<br />

The function f(x) varies from 0 (valve closed with x equal to 0) to 1 (valve fully<br />

open with x equal to 1). The inherent valve characteristic f(x), is intimately linked to<br />

the flow rate q through a valve, as we can see if we replace C with C V f(x) (as per<br />

Equation 6-10), to rewrite the valve Equation 6-9 in the form:<br />

q = C V f(x)<br />

ΔP<br />

G<br />

Equation 6-11<br />

It is indeed clear from Equation 6-11 that if we keep Δp constant (for a given liquid, G<br />

is constant as well), q and f(x) will have exactly the same shape.


EIPINI Chapter 6: Process Control Page 6-33<br />

A valve for which f(x) = x, is called a linear valve because the flow rate will change<br />

linearly with valve opening x. Depending however on the valve design, f(x) may also<br />

reflect quick opening or equal percentage (slow opening) characteristics, as shown in<br />

Figure 6-42.<br />

100%<br />

Graphs also<br />

represent<br />

flow rate q<br />

(% of max.<br />

flow) if Δp<br />

is constant.<br />

0%<br />

q<br />

f(x)<br />

1<br />

Quick<br />

opening<br />

Linear<br />

Figure 6-42<br />

Inherent valve<br />

characteristics<br />

(Ratio of<br />

Equal<br />

valve travel<br />

percentage x to maximum<br />

0<br />

valve travel)<br />

0 1<br />

(0% open) (100% open)<br />

Quick<br />

opening<br />

Linear<br />

Equal<br />

percentage<br />

The flow behaviour of a valve is dictated by the manner in which the flow areas A 1<br />

and A 2 changes with valve position and therefore by the style and design of the valve<br />

trimming and in particular the design of the valve seat and closure member (plug). The<br />

quick opening flow characteristic provides for maximum change in flow rate at low<br />

valve travels with a nearly linear relationship. Additional increases in valve travel<br />

gives sharply reduced changes in flow rate. The linear flow characteristic curve allows<br />

the flow rate to be directly proportional to the valve travel (Δq/Δx equals a constant) or<br />

in terms of the inherent valve characteristic, f(x) = x. An equal percentage valve starts<br />

initially with a slow increase in flow rate with valve position which dramatically<br />

increases as the valve opens more.<br />

The term equal percentage for a slow opening characteristic curve may at first be<br />

confused with the description of a linear characteristic curve. However, for an equal<br />

percentage valve, Δq/Δx at any stage, is proportional to the flow rate q at that moment.<br />

This is in contrast with a linear characteristic for which Δq/Δx is constant. That Δq/Δx<br />

is proportional to q, may be rephrased as Δq/q is proportional to Δx. This means the<br />

percentage change Δq with respect to the current flow rate q (that is (Δq/q)×100), is<br />

equal at every valve travel position x for the same change in valve travel Δx, hence the<br />

term ‘equal percentage’. The inherent valve characteristic for an equal percentage<br />

valve is exponential in nature and is normally given by valve manufacturers in the<br />

form f(x) = R x-1 , where R is a constant for the valve. The exponential behaviour of an<br />

equal percentage valve is explored further in Example 6-7.<br />

Example 6-7<br />

An equal percentage valve delivers 8 gpm of water when the valve is 50% open<br />

(x=0.5). When the valve is 60% open (x=0.6), the flow rate increases to 16 gpm.


EIPINI Chapter 6: Process Control Page 6-34<br />

Estimate the flow rate through the valve when it is 70% open (x=0.7). Assume that the<br />

pressure drop across the valve remains constant.<br />

Answer: When the valve opens from<br />

50% to 60%, the flow rate changes<br />

q (gpm.)<br />

from 8 gpm to 16 gpm. This means<br />

that Δq is 8 gpm. when Δx is 0.1 and<br />

=%<br />

q = 8. Therefore Δq/q = 8/8 = 1 and<br />

the percentage change is 100%<br />

<br />

(Δq/q×100) at x = 0,5. For an equal<br />

Δq=<br />

percentage valve, the percentage 16<br />

change in flow rate when the valve 8<br />

Δq=8<br />

x<br />

opens from 50% to 60%, (Δx = 0.1)<br />

must be equal to the percentage<br />

change in flow rate when the valve<br />

0<br />

0 .5 .6 .7 1<br />

opens from 60% to 70% (the same<br />

Δx=.1 Δx=.1<br />

Δx of 0.1). Therefore Δq/q at x = 0.6 must also be 1 (or 100%) for Δx = 0,1. But q is<br />

16 gpm. when the valve is 60% open and we conclude that Δq should be 16 gpm.<br />

when the valve opening changes from 60% to 70%. The flow rate through the valve is<br />

therefore approximately 32 gpm. at 70% valve opening.<br />

Typical application of quick opening, linear and equal percentage valves<br />

i) Quick opening valve:<br />

a) Frequent on-off service.<br />

b) Used for systems where ‘instant’ large flow is needed (safety or cooling<br />

water systems).<br />

ii) Linear valve:<br />

a) Liquid level and flow control loops.<br />

b) Used in systems where the pressure drop across the valve is expected to<br />

remain fairly constant.<br />

iii) Equal percentage valve (most commonly used valve):<br />

a) Temperature and pressure control loops.<br />

b) Used in systems where large changes in pressure drop across the valve are<br />

expected.<br />

Installed flow characteristic<br />

When valves are installed with pumps, piping and fittings, and other <strong>process</strong><br />

equipment, the pressure drop across the valve will vary as the valve travel changes.<br />

When the actual flow in a system is plotted against valve opening, the curve is called<br />

the installed flow characteristic and it will differ from the inherent valve characteristic<br />

which assumed constant pressure drop across the valve. When in service, a linear<br />

valve will in general resemble a quick opening valve while an equal percentage valve<br />

will in general resemble a linear valve.


EIPINI Chapter 6: Process Control Page 6-35<br />

Valve sizing<br />

The characteristic flow coefficient C V of a valve and the inherent valve characteristic<br />

f(x), completely describe the flow-pressure profile of the valve and play an important<br />

role during the valve selection or valve sizing <strong>process</strong>. Although the characteristic<br />

flow coefficient (C V ) is the most fundamental and important parameter influencing the<br />

user when selecting a valve, it is not the only consideration. Conditions such as piping<br />

particulars, fluid type, laminar or turbulent flow, may be included as additional factors<br />

in Equation 6-9. Nevertheless, to select an appropriate valve for a certain<br />

implementation, the user must specify the maximum flow rate required and the<br />

pressure drop expected with the valve fully open. With this information, the necessary<br />

characteristic valve flow coefficient C V , can be calculated from Equation 6-9 by<br />

setting C = C V , q = Q Rated (rated maximum flow rate with valve fully opened) and<br />

Δp = ΔP Full (pressure across fully opened valve at maximum rated flow rate):<br />

C V =<br />

ΔP<br />

Q Full ………………...…………………. Equation 6-12<br />

Rated G<br />

Example 6-8<br />

A system is pumping water from one tank to another through a piping system with<br />

total pressure drop of 150 psi. The maximum design flow rate is 150 gpm. Calculate<br />

the characteristic flow coefficient required for a control valve to be used in this system.<br />

Solution: The usual rule of thumb is that a valve should be designed to use 10-15% of<br />

the total pressure drop or 10 psi, whichever is greater, when fully open. For this<br />

system 10% of the total pressure drop is 15 psi, which is what we must use. From<br />

Equation 6-12 (remember that specific gravity G is the same as relative density or in<br />

other words, G = δ):<br />

C V = Q Rated / ΔP Full<br />

/G = 150/√(15/1) = 38.72 ≈ 39<br />

(Note: C V is normally given as a dimensionless number but it really is a dimensional<br />

quantity with units [gallon-inch/minute-poundforce ½ ] from Equation 6-12.)<br />

Example 6-9<br />

The user in Example 6-8 has a choice between a valve with C V = 35 which would be<br />

too small and the next one with C V = 45 which could be too large. So he decides to<br />

check what the valve travel positions will be for the larger valve if he wants to control<br />

the water flow between a maximum flow rate of 150 gpm and a minimum flow rate of<br />

40 gpm. He expects a pressure drop of 15 psi across the valve at full flow and a rise in<br />

pressure drop to 25 psi, when the flow rate drops to 40 gpm. The valve he intends to<br />

purchase is a linear valve with inherent valve characteristic, f(x) = x.<br />

Solution: From Equation 6-11 (q = C V f(x) ΔP/G ),<br />

with q max = 150 gpm. and Δp qmax = 15 psi: 150 = 45×x × 15/1<br />

∴x = 150/(45×√15) = 0.8607 (86%)<br />

with q min = 40 gpm. and Δp qmin = 25 psi: 40 = 45×x × 25/1<br />

∴x = 40/(45×√25) = 0.1778 (18%)<br />

This seems to be good enough, as the aim should be to keep valve movement between<br />

20% and 80% of maximum travel.


EIPINI Chapter 6: Process Control Page 6-36<br />

6.4.4 Valve positioners<br />

A 20-100 kPa controller signal travelling perhaps a long distance to a control valve to<br />

exercise the required control action, may possibly not be powerful enough to quickly<br />

move a large valve to its new position. It may therefore be advantageous to allow the<br />

controller signal to control the air supply locally at the valve rather than to operate the<br />

valve by itself. The device that may be added to enhance the controller signal to a<br />

valve, is called a valve positioner. A valve positioner is essentially a pneumatic<br />

controller that senses the valve stem position, compares it to the incoming controller<br />

signal and adjusts the pressure to the actuator until the stem position corresponds to<br />

the controller signal. Positioners can be used to overcome stem friction, high fluid<br />

pressure, viscous or dirty fluids, or to improve slow system dynamic response.<br />

Valve actuator<br />

Elastic<br />

Forcebalance<br />

beam<br />

Pivot<br />

Cam<br />

Instrument<br />

bellows<br />

Flapper and<br />

nozzle<br />

Valve stem<br />

Controller signal<br />

(Instrument signal)<br />

20-100 kPa<br />

Pilot<br />

relay<br />

Restriction<br />

Figure 6-43 Valve positioner<br />

Actuator<br />

output<br />

Air supply<br />

Figure 6-43 shows a typical pneumatic valve positioner, assisting a reverse acting<br />

valve. The operation revolves around a force balance flapper and nozzle<br />

arrangement. The valve stem position is communicated to the force balance beam<br />

by means of a lever and cam. If the controller signal increases, with the intention to<br />

open the valve more, the instrument bellows will push the flapper (a flexible plate)<br />

towards the nozzle. This will increase the valve actuator pressure and the valve<br />

stem will start to move upwards. This will turn the cam which will subsequently<br />

push the flapper away from the nozzle, until balance is reached so that the valve<br />

remains in the new more opened position. If the controller signal decreases, the<br />

flapper will move away from the nozzle and the actuator pressure will be reduced.<br />

The valve will begin to close and the valve stem will begin to move downwards,<br />

relieving cam pressure on the beam that kept the flapper away from the nozzle. The<br />

flapper will tend to move towards the nozzle thereby restoring balance with the<br />

valve more closed than before.


EIPINI Chapter 6: Process Control Page 6-37<br />

APPENDIX 6-1<br />

A water container is 5 meter high and has a base area of 1 meter 2 . The rate, measured<br />

in cubic meter per second, at which water flows into the container at any instant, is<br />

Q IN . The rate, measured in cubic meter per second, at which water flows out of the<br />

container at any instant, is Q OUT . The level, measured in meter, of the water in the<br />

container at any instant, is H. The maximum value of Q IN and Q OUT , is 0.01 cubic<br />

meter per second, and the maximum value of H is 5 meter.<br />

a) Express H in terms of a percentage M, of the maximum level.<br />

b) Express Q IN in terms of a percentage C, of the maximum inflow.<br />

c) Express Q OUT in terms of a percentage L, of the maximum outflow.<br />

d) Calculate the volume V, of the water in the container, in terms of M.<br />

e) Calculate the resultant rate Q, at which water accumulates into the tank or drains<br />

out of the container. Express your answer in terms of C and L.<br />

Δ V<br />

f) Calculate the rate , at which the volume changes, in terms of C and L.<br />

Δt<br />

Δ M<br />

g) Calculate the rate , at which the water level changes, in terms of C and L.<br />

Δt<br />

Solution:<br />

0.01m 3 C (inflow in %)<br />

/s<br />

Q<br />

M Max in (inflow in m 3 /s)<br />

a) H = ×5 = 0.05×M meter …...……… (1)<br />

100<br />

C<br />

5m max<br />

b) Q in = ×0.01 = 0.0001×C m 3 /s ….….. (2) M (%)<br />

100<br />

H (m)<br />

L<br />

c) Q out = ×0.01 = 0.0001×L m 3 /s ….…. (3)<br />

100<br />

L (outflow in %) 0.01m 3 /s<br />

Q<br />

d) Volume V, of water in tank:<br />

out (outflow in m 3 /s) max<br />

V = Height×Area = H×1<br />

Q=0.0001×(C-L)<br />

From (1): V = 0.05×M×1<br />

C<br />

∴V = 0.05×M cubic meter ……..…..….. (4)<br />

e) Resultant rate Q = Q in - Q out<br />

M V=0.05M<br />

From (2) and (3):<br />

∴Q = 0.0001×C–0.0001×L<br />

∴Q = 0.0001(C – L) m 3 /s ….....……....... (5)<br />

Area = 1 m 2 L<br />

f) The rate at which the water volume in the container increases or decreases, is the<br />

same as the rate at which water is accumulated in or drained from the tank.<br />

ΔV<br />

ΔV<br />

∴ = Q. Therefore from (5): = 0.0001(C - L) cubic meter per second.<br />

Δ t Δt<br />

∴ΔV = 0.0001(C - L)×Δt meter 3 ….….... (6)<br />

g) From (4): V = 0.05×M<br />

∴ΔV = 0.05×ΔM meter 3 ……….…......... (7)<br />

Substituting ΔV from (7) into (6): 0.05×ΔM = 0.0001(C – L)×Δt<br />

ΔM<br />

0.0001<br />

dM Δ M<br />

∴ = ×(C – L), or, using the proper notation for ,<br />

Δ t 0.05<br />

dt Δt<br />

dM = 0.002×(C – L) percent per second.<br />

dt


EIPINI Chapter 6: Process Control Page 6-38<br />

APPENDIX 6-2<br />

C<br />

Set point S<br />

(20-100 kPa)<br />

50% = 60 kPa<br />

P d<br />

P d<br />

Needle valve<br />

Rate time<br />

adjustment<br />

(d)<br />

Pilot<br />

relay<br />

Restriction<br />

Controller<br />

output C<br />

Air<br />

supply<br />

S<br />

Beam<br />

Proportional<br />

bellows<br />

P d<br />

M<br />

a<br />

Pivot<br />

b<br />

Reset<br />

bellows<br />

R<br />

Flapper and<br />

nozzle<br />

Measured<br />

value M<br />

(20-100 kPa)<br />

Reset time<br />

adjustment<br />

(r)<br />

Needle<br />

valve<br />

Automatic reset R<br />

Pneumatic PID controller<br />

P d<br />

R<br />

a(r + d) ar<br />

C = (S-M) +<br />

bd<br />

b<br />

Using E = S-M, K P =<br />

a(r + d)<br />

bd<br />

C = K P E + K I<br />

∫ Edt + dE<br />

KD<br />

dt<br />

∫<br />

a<br />

(S-<br />

M)dt (S-M) + bd<br />

d(S- M)<br />

dt<br />

ar a<br />

, K I = and KD = :<br />

b bd

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!