You are on page 1of 155

Ministry of Higher Education Pyramids Higher Institute for

Engineering and Technology, 6th of October City, Giza.

Measurements and Instrumentations 1

Dr. Heba Emara

2nd Term 1st Year

2023-2024
‫رؤية ورسالة المعهد‬
‫يسعى معهد الأهرامات العالي للهندسة والتكنولوجيا لأن يصبح احد‬
‫المعاهـد الرائدة والمصنفة عالمياً في مجال التعليم الهندسي والتكنولوجي‬
‫رؤية‬
‫الذي يوازن بين احتياجات التنمية المحلية والمتغيرات والتحديات العالمية‬ ‫المعهد‬

‫يلتزم معهد الأهرامات العالي للهندسة والتكنولوجيا بإعداد خريجين ذوى‬


‫كفاءة عالية ‪،‬ملتزمين بأخلاقيات المهنة ومؤهلين لتلبية احتياجات ومتطلبات‬
‫سوق العمل المحلى ‪،‬قادرين على مواكبة التطورات الهندسية والتكنولوجيا‬ ‫رسالة‬
‫الحديثة على المشاركة الفعالة والبناءة في تنمية وتطوير المجتمع‪ .‬كما يؤكد‬
‫المعهد التزامه نحو تفعيل وتطوير البحوث العلمية والتطبيقية لتقديم أفضل‬
‫المعهد‬
‫الخدمات للمجتمع ‪،‬والارتقاء بالعملية التعليمية‬

‫رؤية ورسالة القسم‬


‫يأمل برنامج هندسة الإلكترونيات والاتصالات بمعهد الأهرامات العالي‬
‫للهندسة والتكنولوجيا أن يكون أحد البرامج الرائدة محليا ٕواقليميا وعالميا في‬
‫مجال هندسة الإلكترونيات والاتصالات‪ ،‬وأن يقدم تعليما متميزا على مستوى‬
‫عالي يفي بالاحتياجات المحلية وبمتطلبات التطورات والتحديات العالمية‪.‬‬
‫يتوج هذا الطموح بتخريج مهندس قادر على التعامل مع التطبيقات‬ ‫رؤية القسم‬
‫الهندسية المتقدمة وفهمها والتعرف عليها والتي تنصهر فيها العلوم‬
‫والمعارف الموزعة بين هندسة الإلكترونيات والاتصالات والحاسبات‬
‫والشبكات والأقمار الصناعية والأنظمة المدمجة والتحكم‪.‬‬
‫يلتزم برنامج هندسة الإلكترونيات والاتصالات في معهد الأهرامات العالي‬
‫للهندسة والتكنولوجيا بالسادس من أكتوبر بإمداد الطالب بتعليم متميز‬
‫ينمي مستواه الفكري ويطوره كشخص بالغ لديه معايير أخلاقية عالية‪ ،‬قادرا‬
‫على المشاركة الكاملة كقائد في مجتمعه‪ .‬كما يتعهد البرنامج بإعداد‬
‫مهندسين محترفين ذوي مهارات وخبرات عالية ومدربين على مواكبة‬
‫التطورات الحديثة في تكنولوجيا الإلكترونيات والاتصالات والمعلومات‬ ‫رسالة‬
‫الدولية والمحلية الواسعة الانتشار بالمؤسسات المختلفة ‪ ،‬ومؤهلين‬
‫لمواكبة التطور السريع والمستمر في مجال النظم والاتصالات الذكية‬
‫القسم‬
‫بعناصرها (لإلكترونيات والاتصالات والحاسبات والشبكات والأقمار الصناعية‬
‫‪ ،‬الاستشعار عن بعد ‪ ،‬والأنظمة المدمجة والتحكم) ‪ ،‬قادرين على مواصلة‬
‫التعلم والالتحاق ببرامج تدريبية متخصصة وبرامج الدراسات العليا المختلفة‬
‫بما يسهم في تنمية البيئة وخدمة المجتمع‪.‬‬
CONTENTS

Chapter 1 Introduction to Instrument types and Its


performances 6
1. Introduction 6
1.1 Active and passive instruments 6
1.2 Null-type and deflection-type instruments 9
1.3 Analogue and digital instruments 11
1.4 Indicating instruments and instruments with a signal
Output 13
1.5 Smart and non-smart instruments 15
2 Static characteristics of instruments 15
2.1 Accuracy and inaccuracy (measurement uncertainty) 16
2.2 Precision/repeatability/reproducibility 17
2.3 Tolerance 20
2.4 Range or span 21
2.5 Linearity 21
2.6 Sensitivity of measurement 22
2.7 Threshold 23
2.8 Resolution 24

3
2.9 Sensitivity to disturbance 25
2.10 Hysteresis effects 29
2.11 Dead space 31
3 Dynamic characteristics of instruments 31
3.1 Zero order instrument 34
3.2 First order instrument 34
3.3 Second order instrument 40
4 Summary 43

Chapter 2 Measurements and Errors 44

1 Introduction 44
2 Significant Figures 48
3 Types of Error 56
3.1 Gross Errors 57
3.2 Systematic Errors 61
3.3 Random Errors 64
4 Statistical Analysis 65
4.1 Arithmetic Mean 65
4.2 Deviation from the Mean 66
4.3 Average Deviation 68
4.4 Standard Deviation 69
5.1 Normal Distribution of Errors 70
5.2 Probable Error 73
6 Limiting Errors 77
7 Summary 80

Chapter (3) Bridge Measurements 80


3.1 Introduction 80

4
3.1.1 Wheatstone Bridge for Measuring Resistance 80
3.2 Sources and Detectors in AC Bridges 88
3.3 General Balance Equation for Four-Arm Bridge 89
3.4 Measurement of Self-Inductance 93
3.4.1 Maxwell’s Inductance Bridge 93
3.4.2 Maxwell’s Inductance–Capacitance Bridge 95
3.4.3 Hay’s Bridge 100
3.4.4 Owen’s Bridge 103
3.5 Measurement of Capacitance 107
3.5.1 De Sauty’s Bridge 110
3.5.2 Schering Bridge 114
Measurement Laboratory Experiments 116

References 153

Chapter 1

5
Introduction to Instrument types and Its
performance

1 Introduction
Instruments can be subdivided into separate
classes according to several criteria. These sub
classifications are useful in broadly establishing
several attributes of particular instruments such as
accuracy, cost, and general applicability to different
applications.

1.1 Active and passive instruments


Instruments are divided into active or passive ones
according to whether the instrument output is entirely
produced by the quantity being measured or whether
the quantity being measured simply modulates the
magnitude of some external power source. This is
illustrated by examples.
An example of a passive instrument is the
pressure-measuring device shown in Fig.1. The
pressure of the fluid is translated into a movement of
a pointer against a scale. The energy expended in
moving the pointer is derived entirely from the change
in pressure measured: there are no other energy
inputs to the system.
6
An example of an active instrument is a float-type
petrol tank level indicator as sketched in Fig. 2. Here,
the change in petrol level moves a potentiometer arm,
and the output signal consists of a proportion of the
external voltage source applied across the two ends
of the potentiometer. The energy in the output signal
comes from the external power source: the primary
transducer float system is merely modulating the
value of the voltage from this external power source.
In active instruments, the external power source
is usually in electrical form, but in some cases, it can
be other forms of energy such as a pneumatic or
hydraulic one.
One very important difference between active and
passive instruments is the level of measurement
resolution that can be obtained. With the simple
pressure gauge shown, the amount of movement
made by the pointer for a particular pressure change
is closely defined by the nature of the instrument.
Whilst it is possible to increase measurement
resolution by making the pointer longer, such that the
pointer tip moves through a longer arc, the scope for
such improvement is clearly restricted by the practical
limit of how long the pointer can conveniently be. In
an active instrument, however, adjustment of the

7
magnitude of the external energy input allows much

Fig. 1 Passive pressure gauge.

Fig. 2 Petrol-tank level indicator.

greater control over measurement resolution. Whilst


the scope for improving measurement resolution is
much greater incidentally, it is not infinite because of
limitations placed on the magnitude of the external

8
energy input, in consideration of heating effects and
for safety reasons.
In terms of cost, passive instruments are
normally of a more simple construction than active
ones and are therefore cheaper to manufacture.
Therefore, choice between active and passive
instruments for a particular application involves
carefully balancing the measurement resolution
requirements against cost.

1.2 Null-type and deflection-type instruments


The pressure gauge just mentioned is a good
example of a deflection type of instrument, where the
value of the quantity being measured is displayed in

Fig. 3 Deadweight pressure gauge.

terms of the amount of movement of a pointer. An


alternative type of pressure gauge is the deadweight
gauge shown in Fig. 3, which is a null-type instrument.
Here, weights are put on top of the piston until the

9
downward force balances the fluid pressure. Weights
are added until the piston reaches a datum level,
known as the null point. Pressure measurement is
made in terms of the value of the weights needed to
reach this null position.
The accuracy of these two instruments depends
on different things. For the first one it depends on the
linearity and calibration of the spring, whilst for the
second it relies on the calibration of the weights. As
calibration of weights is much easier than careful
choice and calibration of a linear-characteristic
spring, this means that the second type of instrument
will normally be the more accurate. This is in
accordance with the general rule that null-type
instruments are more accurate than deflection types.
In terms of usage, the deflection type instrument
is clearly more convenient. It is far simpler to read the
position of a pointer against a scale than to add and
subtract weights until a null point is reached. A
deflection-type instrument is therefore the one that
would normally be used in the workplace. However,
for calibration duties, the null-type instrument is
preferable because of its superior accuracy. The extra
effort required to use such an instrument is perfectly

10
acceptable in this case because of the infrequent
nature of calibration operations.

1.3 Analogue and digital instruments


An analogue instrument gives an output that
varies continuously as the quantity being measured
changes. The output can have an infinite number of
values within the range that the instrument is
designed to measure. The deflection-type of pressure
gauge described earlier in this chapter (Fig. 1) is a
good example of an analogue instrument. As the input
value changes, the pointer moves with a smooth
continuous motion. Whilst the pointer can therefore
be in an infinite number of positions within its range
of movement, the number of different positions that
the eye can discriminate between is strictly limited,
this discrimination being dependent upon how large
the scale is and how finely it is divided.
A digital instrument has an output that varies in
discrete steps and so can only have a finite number of
values. The rev counter sketched in Fig. 4 is an
example of a digital instrument. A cam is attached to
example of a digital instrument.

11
Fig. 4 Rev counter.

A cam is attached to the revolving body whose motion


is being measured, and on each revolution the cam
opens and closes a switch. The switching operations
are counted by an electronic counter. This system can
only count whole revolutions and cannot discriminate
any motion that is less than a full revolution.
The distinction between analogue and digital
instruments has become particularly important with
the rapid growth in the application of microcomputers
to automatic control systems. Any digital computer
system, of which the microcomputer is but one
example, performs its computations in digital form. An
instrument whose output is in digital form is therefore
particularly advantageous in such applications, as it
can be interfaced directly to the control computer.
Analogue instruments must be interfaced to the
microcomputer by an analogue-to-digital (A/D)
converter, which converts the analogue output signal

12
from the instrument into an equivalent digital quantity
that can be read into the computer. This conversion
has several disadvantages. Firstly, the A/D converter
adds a significant cost to the system. Secondly, a
finite time is involved in the process of converting an
analogue signal to a digital quantity, and this time can
be critical in the control of fast processes where the
accuracy of control depends on the speed of the
controlling computer. Degrading the speed of
operation of the control computer by imposing a
requirement for A/D conversion thus impairs the
accuracy by which the process is controlled.

1.4 Indicating instruments and instruments with a


signal output
The final way in which instruments can be
divided is between those that merely give an audio or
visual indication of the magnitude of the physical
quantity measured and those that give an output in
the form of a measurement signal whose magnitude is
proportional to the measured quantity.
The class of indicating instruments normally
includes all null-type instruments and most passive
ones. Indicators can also be further divided into those
that have an analogue output and those that have a

13
digital display. A common analogue indicator is the
liquid-in-glass thermometer. Another common
indicating device, which exists in both analogue and
digital forms, is the bathroom scale. The older
mechanical form of this is an analogue type of
instrument that gives an output consisting of a
rotating pointer moving against a scale (or sometimes
a rotating scale moving against a pointer).
More recent electronic forms of bathroom scale
have a digital output consisting of numbers presented
on an electronic display. One major drawback with
indicating devices is that human intervention is
required to read and record a measurement. This
process is particularly prone to error in the case of
analogue output displays, although digital displays
are not very prone to error unless the human reader is
careless.
Instruments that have a signal-type output are
commonly used as part of automatic control systems.
In other circumstances, they can also be found in
measurement systems where the output
measurement signal is recorded in some way for later
use. This subject is covered in later chapters. Usually,
the measurement signal involved is an electrical
voltage, but it can take other forms in some systems

14
such as an electrical current, an optical signal or a
pneumatic signal.

1.5 Smart and non-smart instruments


The advent of the microprocessor has created a
new division in instruments between those that do
incorporate a microprocessor (smart) and those that
don’t.

2 Static characteristics of instruments


If we have a thermometer in a room and its
reading shows a temperature of 20°C, then it does not
really matter whether the true temperature of the room
is 19.5°C or 20.5°C. Such small variations around 20°C
are too small to affect whether we feel warm enough
or not. Our bodies cannot discriminate between such
close levels of temperature and therefore a
thermometer with an inaccuracy of ± 0.5°C is perfectly
adequate. If we had to measure the temperature of
certain chemical processes, however, a variation of
0.5°C might have a significant effect on the rate of
reaction or even the products of a process.
A measurement inaccuracy much less than ±
0.5°C is therefore clearly required.
Accuracy of measurement is thus one
consideration in the choice of instrument for a
15
particular application. Other parameters such as
sensitivity, linearity and the reaction to ambient
temperature changes are further considerations.
These attributes are collectively known as the static
characteristics of instruments, and are given in the
data sheet for a particular instrument. It is important
to note that the values quoted for instrument
characteristics in such a data sheet only apply when
the instrument is used under specified standard
calibration conditions. Due allowance must be made
for variations in the characteristics when the
instrument is used in other conditions.
The various static characteristics are defined in
the following paragraphs.

2.1 Accuracy and inaccuracy (measurement


uncertainty)
The accuracy of an instrument is a measure of
how close the output reading of the instrument is to
the correct value. In practice, it is more usual to quote
the inaccuracy figure rather than the accuracy figure
for an instrument. Inaccuracy is the extent to which a
reading might be wrong, and is often quoted as a
percentage of the full-scale (f.s.) reading of an
instrument. If, for example, a pressure gauge of range

16
0–10 bar has a quoted inaccuracy of ± 1.0% f.s. (±1%
of full-scale reading), then the maximum error to be
expected in any reading is 0.1 bar. This means that
when the instrument is reading 1.0 bar, the possible
error is 10% of this value. For this reason, it is an
important system design rule that instruments are
chosen such that their range is appropriate to the
spread of values being measured, in order that the
best possible accuracy is maintained in instrument
readings. Thus, if we were measuring pressures with
expected values between 0 and 1 bar, we would not
use an instrument with a range of 0–10 bar. The term
measurement uncertainty is frequently used in place
of inaccuracy.

2.2 Precision/repeatability/reproducibility
Precision is a term that describes an instrument’s
degree of freedom from random errors. If a large
number of readings are taken of the same quantity by
a high precision instrument, then the spread of
readings will be very small. Precision is often, though
incorrectly, confused with accuracy. High precision
does not imply anything about measurement
accuracy. A high precision instrument may have a low
accuracy. Low accuracy measurements from a high

17
precision instrument are normally caused by a bias in
the measurements, which is removable by
recalibration.
The terms repeatability and reproducibility mean
approximately the same but are applied in different
contexts as given below. Repeatability describes the
closeness of output readings when the same input is
applied repetitively over a short period of time, with
the same measurement conditions, same instrument
and observer, same location and same conditions of
use maintained throughout. Reproducibility describes
the closeness of output readings for the same input
when there are changes in the method of
measurement, observer, measuring instrument,
location, conditions of use and time of measurement.
Both terms thus describe the spread of output
readings for the same input. This spread is referred to
as repeatability if the measurement conditions are
constant and as reproducibility if the measurement
conditions vary.
The degree of repeatability or reproducibility in
measurements from an instrument is an alternative
way of expressing its precision. Figure 2.5 illustrates
this more clearly.

18
The figure shows the results of tests on three
industrial robots that were programmed to place
components at a particular point on a table. The target
point was at the centre of the concentric circles
shown, and the black dots represent the points where
each robot actually deposited components at each
attempt. Both the accuracy and precision of Robot 1
are shown to be low in this trial. Robot 2 consistently
puts the component down at approximately the same
place but this is the wrong point. Therefore, it has
high precision but low accuracy. Finally, Robot 3 has
both high precision and high accuracy, because it
consistently places the component at the correct
target position.

19
Fig. 2.5 Comparison of accuracy and precision.

2.3 Tolerance
Tolerance is a term that is closely related to
accuracy and defines the maximum characteristic of
measuring instruments, it is mentioned here because
the accuracy of some instruments is sometimes
quoted as a tolerance figure. When used correctly,
tolerance describes the maximum deviation of a
manufactured component from some specified value.
For instance, crankshafts are machined with a
diameter tolerance quoted as so many microns (10-6

20
m), and electric circuit components such as resistors
have tolerances of perhaps 5%. One resistor chosen
at random from a batch having a nominal value 1000W
and tolerance 5% might have an actual value
anywhere between 950W and 1050 W.

2.4 Range or span


The range or span of an instrument defines the
minimum and maximum values of a quantity that the
instrument is designed to measure.

2.5 Linearity
It is normally desirable that the output reading of
an instrument is linearly proportional to the quantity
being measured. The Xs marked on Figure 2.6 show a
plot of the typical output readings of an instrument
when a sequence of input quantities are applied to it.
Normal procedure is to draw a good fit straight
line through the Xs, as shown in Figure 2.6. (Whilst
this can often be done with reasonable accuracy by
eye, it is always preferable to apply a mathematical
least-squares line-fitting technique. The non-linearity
is then defined as the maximum deviation of any of
the output readings marked X from this straight line.

21
Non-linearity is usually expressed as a percentage of
full-scale reading.
2.6 Sensitivity of measurement
The sensitivity of measurement is a measure of
the change in instrument output that occurs when the
quantity being measured changes by a given amount.
Thus, sensitivity is the ratio of

The sensitivity of measurement is therefore the slope


of the straight line drawn on Fig. 6. If, for example, a
pressure of 2 bar produces a deflection of 10 degrees
in a pressure transducer, the sensitivity of the
instrument is 5 degrees/bar (assuming that the
deflection is zero with zero pressure applied).

Fig. 6 Instrument output characteristic.

22
Example 1
The following resistance values of a platinum
resistance thermometer were measured at a range of
temperatures. Determine the measurement sensitivity
of the instrument in ohms/°C.

Solution
If these values are plotted on a graph, the
straight-line relationship between resistance change
and temperature change is obvious. For a change in
temperature of 30°C, the change in resistance is 7 Ω.
Hence the measurement sensitivity =7/30= 0.233 Ω /°C.

2.7 Threshold
If the input to an instrument is gradually
increased from zero, the input will have to reach a
certain minimum level before the change in the
instrument output reading is of a large enough
magnitude to be detectable. This minimum level of

23
input is known as the threshold of the instrument.
Manufacturers vary in the way that they specify
threshold for instruments. Some quote absolute
values, whereas others quote threshold as a
percentage of full-scale readings. As an illustration, a
car speedometer typically has a threshold of about 15
km/h. This means that, if the vehicle starts from rest
and accelerates, no output reading is observed on the
speedometer until the speed reaches 15 km/h.

2.8 Resolution
When an instrument is showing a particular
output reading, there is a lower limit on the magnitude
of the change in the input measured quantity that
produces an observable change in the instrument
output. Like threshold, resolution is sometimes
specified as an absolute value and sometimes as a
percentage of f.s. deflection. One of the major factors
influencing the resolution of an instrument is how
finely its output scale is divided into subdivisions.
Using a car speedometer as an example again, this
has subdivisions of typically 20 km/h. This means that
when the needle is between the scale markings, we
cannot estimate speed more accurately than to the

24
nearest 5 km/h. This figure of 5 km/h thus represents
the resolution of the instrument.

2.9 Sensitivity to disturbance


All calibrations and specifications of an
instrument are only valid under controlled conditions
of temperature, pressure etc. These standard ambient
conditions are usually defined in the instrument
specification. As variations occur in the ambient
temperature etc., certain static instrument
characteristics change, and the sensitivity to
disturbance is a measure of the magnitude of this
change. Such environmental changes affect
instruments in two main ways, known as zero drift and
sensitivity drift.
Zero drift is sometimes known by the alternative
term, bias.
Zero drift or bias describes the effect where the
zero reading of an instrument is modified by a change
in ambient conditions. This causes a constant error
that exists over the full range of measurement of the
instrument. The mechanical form of bathroom scale is
a common example of an instrument that is prone to
bias. It is quite usual to find that there is a reading of
perhaps 1 kg with no one stood on the scale. If

25
someone of known weight 70 kg were to get on the
scale, the reading would be 71 kg, and if someone of
known weight 100 kg were to get on the scale, the
reading would be 101 kg. Zero drift is normally
removable by calibration. In the case of the bathroom
scale just described, a thumbwheel is usually
provided that can be turned until the reading is zero
with the scales unloaded, thus removing the bias.
Zero drift is also commonly found in instruments like
voltmeters that are affected by ambient temperature
changes. Typical units by which such zero drift is
measured are volts/°C. This is often called the zero
drift coefficient related to temperature changes.
If the characteristic of an instrument is sensitive to
several environmental parameters, then it will have
several zero drift coefficients, one for each
environmental parameter.

A typical change in the output characteristic of a


pressure gauge subject to zero drift is shown in
Figure 7(a).

Sensitivity drift (also known as scale factor drift)


defines the amount by which an instrument’s
sensitivity of measurement varies as ambient
conditions change. It is quantified by sensitivity drift
26
coefficients that define how much drift there is for a
unit change in each environmental parameter that the
instrument characteristics are sensitive to. Many
components within an instrument are affected by
environmental fluctuations, such as temperature
changes: for instance, the modulus of elasticity of a
spring is temperature dependent. Figure 7(b) shows
what effect sensitivity drift can have on the output
characteristic of an instrument. Sensitivity drift is
measured in units of the form (angular degree/bar)/°C.
If an instrument suffers both zero drift and sensitivity
drift at the same time, then the typical modification of
the output characteristic is shown in Fig. 7(c).

Example 2
A spring balance is calibrated in an environment
at a temperature of 20°C and has the following
deflection/load characteristic.

It is then used in an environment at a temperature


of 30°C and the following deflection/ load
characteristic is measured.

27
Determine the zero drift and sensitivity drift per
°C change in ambient temperature.

Fig. 7 Effects of disturbance: (a) zero drift;


(b) sensitivity drift; (c) zero drift plus sensitivity drift.
Solution
At 20°C, deflection/load characteristic is a
straight line. Sensitivity = 20 mm/kg.

28
At 30°C, deflection/load characteristic is still a straight
line. Sensitivity = 22 mm/kg.
Bias (zero drift) =5mm (the no-load deflection)
Sensitivity drift = 2 mm/kg
Zero drift/°C = 5/10 = 0.5 mm/°C
Sensitivity drift/°C = 2/10= 0.2 (mm per kg)/°C

2.10 Hysteresis effects


Fig. 8 illustrates the output characteristic of an
instrument that exhibits hysteresis.
If the input measured quantity to the instrument is
steadily increased from a negative value, the output
reading varies in the manner shown in curve (a). If the
input variable is then steadily decreased, the output
varies in the manner shown in curve (b). The non-
coincidence between these loading and unloading
curves is known as hysteresis.
Two quantities are defined, maximum input
hysteresis and maximum output hysteresis, as shown
in Figure 8. These are normally expressed as a
percentage of the full-scale input or output reading
respectively.

29
Fig. 8 Instrument characteristic with hysteresis.

Hysteresis is most commonly found in


instruments that contain springs, such as the passive
pressure gauge (Fig. 1) and the Prony brake (used for
measuring torque).
It is also evident when friction forces in a system
have different magnitudes depending on the direction
of movement, such as in the pendulum-scale mass-
measuring device.
Devices like the mechanical fly ball (a device for
measuring rotational velocity) suffer hysteresis from
both of the above sources because they have friction
in moving parts and also contain a spring. Hysteresis
can also occur in instruments that contain electrical
30
windings formed round an iron core, due to magnetic
hysteresis in the iron. This occurs in devices like the
variable inductance displacement transducer, the
LVDT and the rotary differential transformer.

2.11 Dead space


Dead space is defined as the range of different
input values over which there is no change in output
value. Any instrument that exhibits hysteresis also
displays dead space, as marked on Fig. 8. Some
instruments that do not suffer from any significant
hysteresis can still exhibit a dead space in their
output characteristics, however, Backlash in gears is
a typical cause of dead space, and results in the sort
of instrument output characteristic shown in Fig. 9.
Backlash is commonly experienced in gear sets used
to convert between translational and rotational motion
(which is a common technique used to measure
translational velocity).

3 Dynamic characteristics of instruments


The static characteristics of measuring
instruments are concerned only with the steady state
reading that the instrument settles down to, such as
the accuracy of the reading etc.

31
Fig. 9 Instrument characteristic with dead space.

The dynamic characteristics of a measuring


instrument describe its behavior between the time a
measured quantity changes value and the time when
the instrument output attains a steady value in
response. As with static characteristics, any values
for dynamic characteristics quoted in instrument data
sheets only apply when the instrument is used under
specified environmental conditions. Outside these
calibration conditions, some variation in the dynamic
parameters can be expected.
In any linear, time-invariant measuring system,
the following general relation can be written between
input and output for time t > 0,

32
(1)
where qi is the measured quantity, q0 is the output
reading and a0 . . . an, b0 . . . bm are constants.

The reader whose mathematical background is


such that the above equation appears daunting
should not worry unduly, as only certain special,
simplified cases of it are applicable in normal
measurement situations. The major point of
importance is to have a practical appreciation of the
manner in which various different types of instrument
respond when the measurand applied to them varies.
If we limit consideration to that of step changes in the
measured quantity only, then equation (1) reduces to:

(2)
Further simplification can be made by taking
certain special cases of equation (2), which
collectively apply to nearly all measurement systems.

33
3.1 Zero order instrument
If all the coefficients a1 . . . an other than a0 in
equation (2) are assumed zero, then:

(3)
where K is a constant known as the instrument
sensitivity as defined earlier.
Any instrument that behaves according to
equation (3) is said to be of zero order type. Following
a step change in the measured quantity at time t, the
instrument output moves immediately to a new value
at the same time instant t, as shown in Figure 10.
A potentiometer, which measures motion, is a
good example of such an instrument, where the
output voltage changes instantaneously as the slider
is displaced along the potentiometer track.
3.2 First order instrument
If all the coefficients a2 . . . an except for a0 and a1 are
assumed zero in equation (2) then:

(4)
Any instrument that behaves according to
equation (4) is known as a first order instrument. If
d/dt is replaced by the D operator in equation (4), we
get:

34
and rearranging this then gives

(5)

Fig. 10 Zero order instrument characteristic.


Defining K = b0/a0 as the static sensitivity and  =
a1/a0 as the time constant of the system, equation (5)
becomes:

(6)
If equation (6) is solved analytically, the output
quantity q0 in response to a step change in qi at time t
varies with time in the manner shown in Fig. 11. The
time constant  of the step response is the time taken
for the output quantity q0 to reach 63% of its final
value.
The liquid-in-glass thermometer is a good
example of a first order instrument. It is well known

35
that, if a thermometer at room temperature is plunged
into boiling water, the output e.m.f. does not rise
instantaneously to a level indicating100°C, but instead
approaches a reading indicating 100°C in a manner
similar to that shown in Fig. 11.
A large number of other instruments also belong
to this first order class. This is of particular
importance in control systems where it is necessary
to take account of the time lag that occurs between a
measured quantity changing in value and the
measuring instrument indicating the change.
Fortunately, the time constant of many first order
instruments is small relative to the dynamics of the
process being measured, and so no serious problems
are created.
Example 2.3
A balloon is equipped with temperature and
altitude measuring instruments and has radio
equipment that can transmit the output readings of
these instruments back to ground. The balloon is
initially anchored to the ground with the instrument
output readings in steady state. The altitude-
measuring instrument is approximately zero order and
the temperature transducer first order with a
time constant of 15

36
Fig. 11 First order instrument characteristic.

seconds. The temperature on the ground, T0, is 10°C


and the temperature Tx at an altitude of x meters is
given by the relation: Tx = T0 - 0.01x.
(a) If the balloon is released at time zero, and
thereafter rises upwards at a velocity of 5
metres/second, draw a table showing the temperature
and altitude measurements reported at intervals of 10
seconds over the first 50 seconds of travel. Show also
in the table the error in each temperature reading.
(b) What temperature does the balloon report at an
altitude of 5000 meters?.
Solution
In order to answer this question, it is assumed
that the solution of a first order differential

37
equation has been presented to the reader in a
mathematics course. If the reader is not so equipped,
the following solution will be difficult to follow.
Let the temperature reported by the balloon at some
general time t be Tr. Then Tx is related to Tr by the
relation:

It is given that x = 5t, thus:

The transient or complementary function part of the


solution (Tx = 0) is given by:

The particular integral part of the solution is given by:

Thus, the whole solution is given by:

Applying initial conditions:

38
Thus C =- 0.75 and therefore:

Using the above expression to calculate Tr for


various values of t, the following table can be
constructed:

(b) At 5000 m, t = 1000 seconds. Calculating Tr from


the above expression:

The exponential term approximates to zero and


so Tr can be written as:

This result might have been inferred from the table


above where it can be seen that the error is
converging towards a value of 0.75. For large values
of t, the transducer reading lags the true temperature
value by a period of time equal to the time constant of

39
15 seconds. In this time, the balloon travels a distance
of 75 meters and the temperature falls by 0.75°. Thus
for large values of t, the output reading is always 0.75°
less than it should be.

3.3 Second order instrument


If all coefficients a3 . . . an other than a0, a1 and a2 in
equation (2) are assumed zero, then we get:

(7)
Applying the D operator again:

and rearranging:

(8)

It is convenient to re-express the variables a0, a1, a2


and b0 in equation (8) in terms of three parameters K
(static sensitivity), ω (undamped natural frequency)
and ξ (damping ratio), where:

(9)
This is the standard equation for a second order
system and any instrument whose response can be
described by it is known as a second order

40
instrument. If equation (9) is solved analytically, the
shape of the step response obtained depends on the
value of the damping ratio parameter ξ. The output
responses of a second order instrument for various
values of ξ following a step change in the value of the
measured quantity at time t are shown in Fig. 12. For
case (A) where ξ = 0, there is no damping and the
instrument output exhibits constant amplitude
oscillations when disturbed by any change in the
physical quantity measured. For light damping of ξ =
0.2, represented by case (B), the response to a step
change in input is still oscillatory but the oscillations
gradually die down. Further increase in the value of ξ
reduces oscillations and overshoot still more, as
shown by curves (C) and (D), and finally the response
becomes very over damped as shown by curve (E)
where the output reading creeps up slowly towards
the correct reading. Clearly, the extreme response
curves (A) and (E) are grossly unsuitable for any
measuring instrument. If an instrument were to be
only ever subjected to step inputs, then the design
strategy would be to aim towards a damping ratio of
0.707, which gives the critically damped response (C).
Unfortunately, most of the physical quantities that
instruments are required to measure do not change in

41
the mathematically convenient form of steps, but
rather in the form of ramps of varying slopes. As the
form of the input variable changes, so the best value
for ξ varies, and choice of ξ becomes one of
compromise between those values that are best for
each type of input variable behaviour anticipated.
Commercial second order instruments, of which the
accelerometer is a common example, are generally
designed to have a damping ratio (ξ) somewhere in
the range of 0.6 – 0.8.

Fig. 12 Response characteristics of second order instruments.

4 Summary

This chapter discussed the static and dynamic


characteristics of measuring instruments in some
detail. However, an important qualification that has
been omitted from this discussion is that an
42
instrument only conforms to stated static and
dynamic patterns of behaviour after it has been
calibrated. It can normally be assumed that a new
instrument will have been calibrated when it is
obtained from an instrument manufacturer, and will
therefore initially behave according to the
characteristics stated in the specifications. During
use, however, its behaviour will gradually diverge
from the stated specification for a variety of reasons.
Such reasons include mechanical wear, and the
effects of dirt, dust, fumes and chemicals in the
operating environment. The rate of divergence from
standard specifications varies according to the type of
instrument, the frequency of usage and the severity of
the operating conditions. However, there will come a
time, determined by practical knowledge, when the
characteristics of the instrument will have drifted from
the standard specification by an unacceptable
amount. When this situation is reached, it is
necessary to recalibrate the instrument to the
standard specifications. Such recalibration is
performed by adjusting the instrument at each point in
its output range until its output readings are the same
as those of a second standard instrument to which the
same inputs are applied. This second instrument is
one kept solely for calibration purposes whose
specifications are accurately known.

Chapter two

Measurements and Errors

43
1 Introduction
Measurement generally involves using an
instrument as a physical means of determining a
quantity or variable. The instrument serves as an
extension of human faculties and in many cases
enables a person to determine the value of an
unknown quantity which his unaided human faculties
could not measure.
An instrument, then, may be defined as a device
for determining the value or magnitude of a quantity
or variable. The electronic instrument, as its name
implies, is based on electrical and electronic
principles for its measurement function. An electronic
instrument may be a relatively uncomplicated device
of simple construction such as a basic dc current
meter. As technology expands, however, the demand
for more elaborate and more accurate instrument
increases and produces new developments in
instrument design and application. To use these
instruments intelligently, one needs to understand
their operating principles and to appraise their
suitability for the intended application.
Measurement work employs a number of terms
which should be defined here.

44
Instrument a device for determining the value
or magnitude of a quantity or variable.
Accuracy closeness with which an instrument
reading approaches the true value of the variable
being measured.
Precision a measure of the reproducibility of
the measurements; i.e., given a fixed value of a
variable, precision is a measure of the degree to
which successive measurements differ from one
another.
Sensitivity the ratio of output signal or
response of the instrument to a change of input or
measured variable.
Resolution the smallest change in measured
value to which the instrument will respond.
Error deviation from true value of
the measured value.

Several techniques may be used to minimize the


effects of errors. For example, in making precision
measurements, it is advisable to record a series of
observations rather than rely on one observation.
Alternate methods of measurement, as well as the use
of different instruments to perform the same
experiment, provide a good technique for increasing

45
accuracy. Although these techniques tend to increase
the precision of measurement by reducing
environmental or random error, they cannot account
for instrumental error.

These notes provide an introduction to different


types of error in measurement and to the methods
generally used to express errors, in terms of the most
reliable value of the measured variable.

Accuracy refers to the degree of closeness or


conformity to the true value of the quantity under
measurement. Precision refers to the degree of
agreement within a group of measurements or
instruments.

To illustrate the distinction between accuracy


and precision, two voltmeters of the same make and
model may be compared. Both meters have knife-
edged pointers and mirror-backed scales to avoid
parallax, and they have carefully calibrated scales.
They may therefore be read to the same precision. If
the value of the series resistance in one meter
changes considerably, its readings may be in error by
a fairly large amount. Therefore the accuracy of the

46
two meters may be quite different. (To determine
which meter is in error, a comparison measurement
with a standard meter should be made.)

Precision is composed of two characteristics:


conformity and the number of significant figures to
which a measurement can be made. Consider, for
example, that a resistor, whose true resistance is
1,384Ω, 572 Ω, is measured by an ohmmeter which
consistently and repeatedly indicates 1.4 M Ω. But can
the observer “read” the true value from the scale?. His
estimates from the scale reading consistently yield a
value of 1.4 M Ω. This is as close to the true value as
he can read the scale by estimation. Although there
are no deviations from the observed value, the error
created by the limitation of the scale reading is a
precision error. The example illustrates that
conformity is a necessary, but not sufficient condition
for precision because of the lack of significant figures
obtained. Similarly, precision is a necessary, but not
sufficient condition for accuracy.

Too often the beginning student is inclined to


accept instrument readings at face value. He is not
aware that the accuracy of a reading is not necessarily

47
guaranteed by its precision. In fact, good
measurement techniques demands continuous
skepticism as to the accuracy of the results.

In critical work, good practice dictates that the


observer make an independent set of measurements,
using different instruments or different measurement
techniques, not subject to the same systematic errors.
He must also make sure that the instruments function
properly and are calibrated against a known standard,
and that no outside influence affects the accuracy of
his measurements.

2 Significant Figures
An indication of the precision of the
measurement is obtained from the number of
significant figures in which the result is expressed.
Significant figures convey actual information
regarding the magnitude and the measurement
precision of a quantity. The more significant figures,
the greater the precision of
measurement.

For example, if a resistor is specified as having a


resistance of 68Ω , its resistance should be closer to

48
68 Ω than to 67 Ω or 69 Ω . If the value of the resistor
is described as 68.0 Ω , it means that its resistance is
closer to 68.0 Ω than it is to 67.9 Ω or 68.1 Ω . In 68 Ω
there are two significant figures; in 68.0 Ω there are
three. The latter, with more significant figures,
expresses a measurement of greater precision than
the former.
Often, however, the total number of digits may
not represent measurement precision. Frequently,
large numbers with zeros before a decimal point are
used for approximate populations or amounts of
money.
For example, the population of a city is reported
in six figures as 380,000. This may imply that the true
value of the population lies between 379,999 and
380,001, which is six significant figures. What is
meant, however, is that the population is closer to
380,000 than 370,000 or 390,000. Since in this case the
population can be reported only to two significant
figures, how can large numbers be expressed?
A more technically correct notation uses powers
of ten, 38 × 104 or 3.8 × 105. This indicates that the
population figure is only accurate to two significant
figures. Uncertainty caused by zeros to the left of the
decimal point is therefore usually resolved by

49
scientific notation using powers of ten. Reference to
the velocity of light as 186,000 mi/s, for example,
would cause no misunderstanding to anyone with a
technical background. But 1.86 × 105 mi/s leaves no
confusion.
It is customary to record a measurement with all
the digits of which we are sure nearest to the true
value. For example, in reading a voltmeter, the voltage
may be read as 117.1 V. This simply indicates that the
voltage, read by the observer to best estimation, is
closer to 117.1 V than to 117.0 V or 117.2 V.
Another way of expressing this result indicates the
range of possible error. The voltage may be expressed
as 117.1 ± 0.05 V, indicating that the value of the
voltage lies between 117.05 V and 117.15 V.
When a number of independent measurements
are taken in an effort to obtain the best possible
answer (closest to the true value), the result is usually
expressed as the arithmetic mean of all the readings,
with the range of possible error as the largest
deviation from that mean. This is illustrated in

Example 1:
A set of independent voltage measurements
taken by four observers was recorded as 117.02 V,
117.11 V,
50
117.08 V, and 117.03 V. Calculate
1. the average voltage,
2. the range of error.
Solution
1. Average Voltage:

2. Range:

When two or more measurements with different


degrees of accuracy are added, the result is only as
accurate as the least square measurement. Suppose
that two resistances are added in series as in Example
2.

51
Example 2
Two resistors R1 and R2, are connected in series.
Individual resistance measurements, using a
Wheatstone bridge, give R1 = 18.7 Ω and R2 = 3.624 Ω.
Calculate the total resistance to the appropriate
number of significant figures.
Solution
R1 = 18.7 Ω (three significant figures)
R2 = 3.624 Ω (four significant figures)
RT = R1 + R2 = 22.324 Ω (five significant figures)
= 22.3 Ω.
The doubtful figures are written in italics to indicate
that in the addition of R1 and R2 the last three digits
of the sum are doubtful figures. There is no value
whatsoever in retaining the last two digits (the 2 and
the 4 ) because one of the resistances is accurate only
to three significant figures or tenths of an ohm. The
result should therefore also be reduced to three
significant figures or the nearest tenth, i.e., 22.3 Ω .
The number of significant figures in multiplication
may increase rapidly, but again only the appropriate
figures are retained in the answer, as shown in
Example 3.

Example 3
52
In calculating voltage drop, a current of 3.18 A is
recorded in a resistance of 35.68Ω. Calculate the
voltage drop across the resistor to the appropriate
number of significant figures.

Solution

E = IR = 3.18 × 35.68 = 113.4624 = 113V


Since there are three significant figures involved
in the multiplication, the answer can be written only to
a maximum of three significant figures.
In Example 3, the current, I, has three significant
figures and R has four; and the result of the
multiplication has only three significant figures. This
illustrates the answer cannot be known to an accuracy
greater than the least poorly defined of the factors.
Note also that if extra digits accumulate in the answer,
they should be discarded or rounded off. In the usual
practice, if the (least significant) digit in the first place
to be discarded is less than five, it and the following
digits are dropped from the answer.
This was done in Example 3. If the digit in the first
place to be discarded is five or greater, the previous
digit is increased by one. For three-digit precision,
therefore, 113.46 should be rounded off to 113; and
113.74 to 114.

53
Addition of figures with a range of doubt is
illustrated in Example 4.

Example 4
Add 826 ± 5 to 628 ± 3

Solution
N1 = 826 ± 5 (= ±0.605%)
N2 = 628 ± 3 (= ±0.477%)
Sum = 1454 ± 8 (= ±0.55%)
Note in Example 4 that the doubtful parts are
added, since the ± sign means that one number may
be high and the other low. The worst possible
combination of range of doubt should be taken in the
answer.
The percentage doubt in the original figure N1
and N2 does not differ greatly from the percentage
doubt in the final result.
If the same two numbers are subtracted, as in
Example 5, there is an interesting comparison
between addition and subtraction with respect to the
range of doubt.

Example 5
Subtract 628 ± 3 from 826 ± 5 and express the range of
doubt in the answer as a percentage.

54
Solution
N1 = 826 ± 5 (= ±0.605%)
N2 = 628 ± 3 (= ±0.477%)
Difference = 198 ± 8 (= ±4.04%)

Again in Example 5, the doubtful parts are added


for the same reason as in Example 4. Comparing the
results of addition and subtraction of the same
numbers in Example 4 and 5, note that the precision
of the results, when expressed in percentages, differs
greatly. The final result after subtraction shows a
large increase in percentage doubt compared to the
percentage doubt after addition. The percentage
doubt increases even more when the difference
between the numbers is relatively small. Consider the
case illustrated in Example 6.

Example 6
Subtract 437 ± 4 from 462 ± 4 and express the
range of doubt in the answer as a percentage.
Solution
N1 = 462 ± 4 (= ±0.87%)
N2 = 437 ± 4 (= ±0.92%)
Difference = 25 ± 8 (= ±32%)

55
Example 6 illustrates clearly that one should
avoid measurement techniques depending on
subtraction of experimental results because the range
of doubt in the final result may be greatly increased.

3 Types of Error
No measurement can be made with perfect
accuracy, but it is important to find out what the
accuracy actually is and how different errors have
entered into the measurement. A study of errors is a
first step in finding ways to reduce them. Such a study
also allows us to determine the accuracy of the final
test result.
Errors may come from different sources and are
usually classified under three main headings:
Gross Errors: Largely human errors, among them
misreading of instruments, incorrect adjustment and
improper application of instruments, and
computational mistakes.
Systematic Errors: Shortcomings of the instruments,
such as defective or worn parts, and effects of the
environment on the equipment or the user.
Random Errors: Those due to causes that cannot be
directly established because of random variations in
the parameter or the system of measurement.

56
Each of these classes of errors will be discussed
briefly and some methods will be suggested for their
reduction or elimination.

3.1 Gross Errors


This class of errors mainly covers human
mistakes in reading or using instruments and in
recording and calculating measurement results. As
long as human beings are involved, some gross
errors will inevitably be committed. Although
complete elimination of gross errors is probably
impossible, one should try to anticipate and correct
them. Some gross errors are easily detected; others
may be very elusive. One common gross error,
frequently committed by beginners in measurement
work, involves the improper use of an instrument. In
general, indicating instruments change conditions to
some extent when connected into a complete circuit,
so that the measured quantity is altered by the
method employed. For example, a well-calibrated
voltmeter may give a misleading reading when
connected across two points in a high resistance
circuit (Example 7). The same voltmeter, when
connected in a low-resistance circuit, may give a more
dependable reading (Example 8). These examples

57
illustrate that the voltmeter has a “loading effect” on
the circuit, altering the original situation by the
measurement process.

Example 7
A voltmeter, having a sensitivity of 1,000 /V, reads 100
V on its 150-V scale when connected across an
unknown resistor in series with a milliammeter.
When the milliammeter reads 5 mA, Calculate:
1. Apparent resistance of the unknown resistor,
2. Actual resistance of the unknown resistor,
3. Error due to the loading effect of the voltmeter.

Solution
1. The total circuit resistance equals

Neglecting the resistance of the milliammeter, the


value of the unknown resistor is RX = 20 k.
2. The voltmeter resistance equals

Since the voltmeter is in parallel with the


unknown resistance, we can write

58
3. Error

Example 8
Repeat Example 7 if the milliammeter reads 800 mA
and the voltmeter reads 40 V on its 150-V scale.
Solution
1. The total circuit resistance equals

2. The voltmeter resistance equals

Since the voltmeter is in parallel with the


unknown resistance, we can write

2. Error

Errors caused by the loading effect of the voltmeter


can be avoided by using it intelligently. For example, a
59
low-resistance voltmeter should not be used to
measure voltages in a vacuum tube amplifier. In this
particular measurement, a high-input impedance
voltmeter (such as VTVM or TVM) is required.
A large number of gross errors can be attributed to
carelessness or bad habits, such as improper reading
of an instrument, recording the result differently from
the actual reading taken, or adjusting the instrument
incorrectly. Consider the case in which a multi-range
voltmeter uses a single set of scale markings with
different number designations for the various voltage
ranges. It is easy to use a scale which does not
correspond to the setting of the range selector of the
voltmeter. A gross error may also occur when the
instrument is not set to zero before the measurement
is taken; then all the readings are off.
Errors like these cannot be treated mathematically.
They can be avoided only by taking care in reading
and recording the measurement data. Good practice
requires making more than one reading of the same
quantity, preferably by a different observer. Never
place complete dependence on one reading but take
at least three separate readings, preferably under
conditions in which instruments are switched off-on.

60
3.2 Systematic Errors
This type of errors is usually divided into two
different categories:
1. instrumental errors, defined as shortcomings of the
instrument;
2. environmental errors, due to external conditions
affecting the measurement.
Instrumental errors are errors inherent in
measuring instruments because of their mechanical
structure.
For example, in the d’Arsonval movement,
friction in bearings of various moving components
may cause incorrect readings. Irregular spring
tension, stretching of the spring, or reduction in
tension due to improper handling or overloading of
the instrument will result in errors. Other instrumental
errors are calibration errors, causing the instrument to
read high or low along its entire scale. (Failure to set
the instrument to zero before making a measurement
has a similar effect.)
There are many kinds of instrumental errors,
depending on the type of instrument used. The
experimenter should always take precautions to
insure that the instrument he is using is operating
properly and does not contribute excessive errors for

61
the purpose at hand. Faults in instruments may be
detected by checking for erratic behaviour, and
stability and reproducibility of results. A quick and
easy way to check an instrument is to compare it to
another with the same characteristics or to one that is
known to be more accurate.

Instrumental errors may be avoided by:


1. selecting a suitable instrument for a particular
measurement application;
2. applying correction factors after determining the
amount of instrumental error;
3. calibrating the instrument against a standard.

Environmental errors are due to conditions


external to the measuring device, including conditions
in the area surrounding the instrument, such as the
effects of changes in temperature, humidity,
barometric pressure, or of magnetic or electrostatic
fields. Thus a change in ambient temperature at which
the instrument is used causes a change in the elastic
properties of the spring in a moving-coil mechanism
and so affects the reading of the instrument.
Corrective measures to reduce these effects include
air conditioning, hermetically sealing certain

62
components in the instrument, use of magnetic
shields, and the like.

Systematic errors can also be subdivided into


static or dynamic errors. Static errors are caused by
limitations of the measuring device or the physical
laws governing its behaviour.
A static error is introduced in a micrometer when
excessive pressure is applied in torquing the shaft.

Dynamic errors are caused by the instrument’s


not responding fast enough to follow the changes in a
measured variable.

3.3 Random Errors


These errors are due to unknown causes and occur
even when all systematic errors have been accounted
for. In well-designed experiments, few random errors
usually occur, but they become important in high-
accuracy work. Suppose a voltage is being monitored
by a voltmeter which is read at half-hour intervals.
Although the instrument is operated under ideal
environmental conditions and has been accurately
calibrated before the measurement, it will be found
that the readings vary slightly over the period of
observation. This variation cannot be corrected by
any method of calibration or other known method of
63
control and it cannot be explained without minute
investigation. The only way to offset these errors is by
increasing the number of readings and using
statistical means to obtain the best approximation of
the true value of the quantity under measurement.

4 Statistical Analysis
A statistical analysis of measurement data is
common practice because it allows an analytical
determination of the uncertainty of the final test result.
The outcome of a certain measurement method may
be predicted on the basis of sample data without
having detailed information on all the disturbing
factors.
To make statistical methods and interpretations
meaningful, a large number of measurements is
usually required. Also, systematic errors should be
small compared with residual or random errors,
because statistical treatment of data cannot remove a
fixed bias contained in all the measurements.

4.1 Arithmetic Mean


The most probable value of a measured variable
is the arithmetic mean of the number of readings
taken.

64
The best approximation will be made when the
number of readings of the same quantity is very large.
Theoretically, an infinite number of readings would
give the best result, although in practice, only a finite
number of measurements can be made. The arithmetic
mean is given by the following expression:

Example 1 showed how the arithmetic mean is used.

4.2 Deviation from the Mean


Deviation is the departure of a given reading from
the arithmetic mean of the group of readings. If the
deviation of the first reading, x1, is called d1, and that
of the second reading, x2, is called d2, and so on, then
the deviations from the mean can be expressed as

Note that the deviation from the mean may have a


positive or a negative value and that the algebraic
sum of all the deviations must be zero.

65
Example 9 illustrates the computation of deviations.

Example 9
A set of independent current measurements was
taken by six observers and recorded as 12.8 mA, 12.2
mA, 12.5 mA, 13.1 mA, 12.9 mA, and 12.4 mA.
Calculate
1. the arithmetic mean,
2. the deviations from the mean.

Solution
1. Using Eq. (1), we see that the arithmetic
mean equals:

2. Using Eq. (2), we see that the deviations are


d1 = 12.8 − 12.65 = 0.15 mA
d2 = 12.2 − 12.65 = −0.45 mA
d3 = 12.5 − 12.65 = −0.15 mA
d4 = 13.1 − 12.65 = 0.45 mA
d5 = 12.9 − 12.65 = 0.25 mA
d6 = 12.4 − 12.65 = −0.25 mA

66
Note that the algebraic sum of all the
deviations equals zero.

4.3 Average Deviation


The average deviation is an indication of the
precision of the instruments used in making the
measurements. Highly precise instruments will yield a
low average deviation between readings. By
definition, average deviation is the sum of the
absolute values of the deviation divided by the
number of readings.
The absolute value of the deviation is the value
without respect to sign. Average deviation may be
expressed as:

Example 10 shows how average deviation is


calculated.

Example 10
Calculate the average deviation for the data given
in Example 9.
Solution

67
4.4 Standard Deviation
In statistical analysis of random errors, the root-
mean-square deviation or standard deviation is a very
valuable aid. By definition, the standard deviation of
an infinite number of data is the square root of the
sum of all the individual deviations squared, divided
by the number of readings. Expressed
mathematically:

In practice, of course, the possible number of


observations is finite. The standard deviation of a
finite number of data is given by

Equation (5) will be used in Example 11.


Another expression for essentially the same
quantity is the variance or mean square deviation,
which is the same as the standard deviation except
that the square root is not extracted. Therefore:

variance(V ) = mean square deviation =σ2.


The variance is a convenient quantity to use in
many computations because variances are additive.

68
The standard deviation, however, has the advantage
of being of the same units as the variable, making it
easy to compare magnitudes. Most scientific results
are now stated in terms of standard deviation.

5.1 Normal Distribution of Errors


Table 1 shows a tabulation of 50 voltage readings
that were taken at small time intervals and recorded to
the nearest 0.1 V. The nominal value of the measured
voltage was 100.0 V. The result of this series of
measurements can be presented graphically in the
form of a block diagram or histogram in which the
number of observations is plotted against each
observed voltage reading. The histogram of Figure 1
represents the data of Table1.

Figure 1 shows that the largest number of


readings (19) occurs at the central value of 100.0 V,
while the other readings are placed more or less
symmetrically on either side of the central value. If
more readings were taken at smaller increments, say
200 readings at 0.05 V intervals, the distribution of

69
Figure 1: Histogram showing the frequency of occurrence of
the 50 voltage readings of Table 1.

observations would remain approximately


symmetrical about the central value and the shape of
the histogram would be about the same as before.
With more and more data, taken at smaller and smaller

70
increments, the contour of the histogram would finally
become a smooth curve, known as a Gaussian curve.
The sharper and narrower the curve, the more
definitely an observer may state that the most
probable value of the true reading is the central value
or mean reading.
The Gaussian or Normal law of error forms the
basis of the analytical study of random effects.
Although the mathematical treatment of this subject is
beyond the scope of the text, the following qualitative
statements are based on the Normal law:
1. All observations include small disturbing effects,
called random errors.
2. Random errors can be positive or negative.
3. There is an equal probability of positive and
negative random errors.
We can therefore expect that measurement
observations include plus and minus errors in more or
less equal amounts, so that the total error will be
small and the mean value will be the true value of the
measured variable.
The possibilities as to the form of the error
distribution curve can be stated as follows:
1. Small errors are more probable than large errors.
2. Large errors are very improbable.

71
3. There is an equal probability of plus and minus
errors so that the probability of a given error will be
symmetrical about the zero value.
The error distribution curve of Figure 2 is based
on the Normal law and shows a symmetrical
distribution of errors. This normal curve may be
regarded as the limiting form of the histogram of
Figure 1 in which the most probable value of the true
voltage is the mean value of 100.0 V.

5.2 Probable Error


The area under the Gaussian probability curve of
Figure 2, between the limits +∞ and −∞, represents the
entire number of observations. The area under the
curve between the + σ and − σ limits represents the
cases that differ from the mean by no more than the
standard deviation. Integration of the area under the
curve within the ± σ limits gives the total number of
cases within these limits. For normally dispersed data,
following the Gaussian distribution, approximately
68% of all the cases lie between the limits of + σ and
− σ from the mean. Corresponding values of other
deviations, expressed in terms of σ, are given in Table
2.

72
If, for example, a large number of nominally 100Ω
resistors is measured and the mean value is found to
be 100.00Ω , with a standard deviation of 0.20Ω , we
know that on the average 68% (or roughly two-thirds)

of all the resistors have values which lie between


limits of ±0.20Ω of the mean. There is then
approximately a two to one chance that any resistor,
selected from the lot at random, will lie within these
73
limits. If larger odds are required, the deviation may
be extended to a limit of ±2 σ, in this case ±0.40 Ω .
According to Table 2, this now includes 95% of all the
cases, giving ten to one odds that any resistor
selected at random lies within ±0.40 Ω of the mean
value of 100.00 Ω .

Table 2 also shows that half of the cases are


included in the deviation limits of ±0.6745 σ. The
quantity r is called the probable error and is defined
as:
probable error r = ±0.6745 σ (6)

The value is probable in the sense that there is an


even chance that any one observation will have a
random error no greater than ±r. Probable error has
been used in experimental work to some extent in the
past, but standard deviation is more convenient in
statistical work and is given preference.

Example 11
Ten measurements of the resistance of a resistor
gave 101.2 Ω , 101.7 Ω , 101.3 Ω , 101.0 Ω , 101.5 Ω ,
101.3 Ω , 101.2 Ω , 101.4 Ω, 101.3 Ω , and 101.1 Ω .
Assume that only random errors are present.

74
Calculate:
1. the arithmetic mean,
2. the standard deviation of the readings,
3. the probable error.

Solution

With a large number of readings a simple


tabulation of data is very convenient and avoids
confusion and mistakes.

75
6 Limiting Errors
In most indicating instruments, the accuracy is
guaranteed to a certain percentage of full-scale
reading. Circuit components (such as capacitors,
resistors, etc.) are guaranteed within a certain
percentage of their rated value. The limits of these
deviations from the specified values are known as
limiting errors or guarantee errors. For example, if the
resistance of a resistor is given as 500 Ω ± 10 per
cent, the manufacturer guarantees that the resistance
falls between the limits 450 Ω and 550 Ω . The maker
is not specifying a standard deviation or a probable
error, but promises that the error is no greater than
the limits set.

Example 12
A 0-150V voltmeter has a guaranteed accuracy of
1 per cent full-scale reading. The voltage measured by
this instrument is 83 V. Calculate the limiting error in
per cent.

Solution
The magnitude of the limiting error is

76
It is important to note in Example 12 that a meter
is guaranteed to have an accuracy of better than 1 per
cent of the full-scale reading, but when the meter
reads 83 V, the limiting error increases to 1.81 per
cent. Correspondingly, when a smaller voltage is
measured, the limiting error will increase further. If
the meter reads 60 V, the per cent limiting error is
(1.5/60) ×100= 2.5 per cent; if the meter reads 30 V, the
limiting error is (1.5/30)×100 = 5 per cent. The increase
in per cent limiting error, as smaller voltages are
measured, occurs because the magnitude of the
limiting error is a fixed quantity based on the full-scale
reading of the meter. Example 12 shows the
importance of taking measurements as close to full
scale as possible.
Measurements or computations, combining
guarantee errors, are often made. Example 13
illustrates such a computation.

Example 13

77
Three decade boxes, each guaranteed to ±0.1 per
cent, are used in a Wheatstone bridge to measure the
resistance of an unknown resistor Rx. Calculate the
limits on Rx imposed by the decade boxes.
Solution
The equation for bridge balance shows that Rx
can be determined in terms of the resistance of the
three decade boxes and Rx = R1*R2/R3, where R1, R2,
and R3 are the resistances of the decade boxes,
guaranteed to ±0.1 per cent. One must recognize that
the two terms in the numerator may both be positive
to the full limit of 0.1 per cent and the denominator
may be negative to the full 0.1 per cent, giving a
resultant error of 0.3 per cent. The guarantee error is
thus obtained by taking the direct sum of all the
possible errors, adopting the algebraic signs which
give the worst possible combination.
As a further example, using the relationship P =
I2*R, as shown in Example 14, consider computing the
power dissipation in a resistor.
Example 14
The current passing through a resistor of 100 ±
0.2 Ω is 2.00 ± 0.01 A. Using the relationship P = I2*R,
calculate the limiting error in the computed value of
power dissipation.

78
Solution
Expressing the guaranteed limits of both current and
resistance in percentages instead of units, we obtain
I = 2.00 ± 0.01 A = 2.00 A ± 0.5%
R = 100 ± 0.2 Ω = 100 Ω ± 0.2%
If the worst possible combination of errors is used,
the limiting error in the power dissipation is (P = I2*R):
(2 × 0.5%) + 0.2% = 1.2%
Power dissipation should then be written as follows:
P = I2*R = (2.00)2 × 100 = 400 W± 1.2% = 400 ± 4.8 W.

7 Summary
This chapter discussed the different types of errors
existing in measuring instrument. Also, different
methods which are used in reducing and overcoming
these errors are given.
Also, normal distribution curves are used for
representing the error of the of the measuring
instrument

79
Chapter (3) Bridge Measurements

Introduction
Alternating current bridges are most popular,
convenient and accurate instruments for
measurement of unknown inductance, capacitance
and some other related quantities. In its simplest
form, ac bridges can be thought of to be derived from
the conventional dc Wheatstone bridge. An ac bridge,
in its basic form, consists of four arms, an alternating
power supply, and a balance detector.
Wheatstone Bridge for Measuring Resistance
The Wheatstone bridge is the most commonly used
circuit for measurement of medium- range
resistances. The Wheatstone bridge consists of four
resistance arms, together with a battery (voltage
source) and a galvanometer (null detector). The circuit
is shown in Figure 3.1.

80
Figure 0.1 Wheatstone bridge for measurement of
esistance
In the bridge circuit, R3 and R4 are two fixed known
resistances, R2 is a known variable resistance and RX
is the unknown resistance to be measured. Under
operating conditions, current ID through the
galvanometer will depend on the difference in
potential between nodes B and C. A bridge balance
condition is achieved by varying the resistance R2
and checking whether the galvanometer pointer is
resting at its zero position. At balance, no current
flows through the galvanometer. This means that at
balance, potentials at nodes B and C are equal. In
other words, at balance the following conditions are
satisfied:
1) The detector current is zero, i.e., 1D = 0 and thus
I1 = I3 and I2 = I4
2) Potentials at node B and C are same, i.e., VB =
VC, or in other words, voltage drop in the arm AB
equals the voltage drop across the arm AC, i.e.,
VAB = VAC and voltage drop in the arm BD
equals the voltage drop across the arm CD, i.e.,
VBD = VCD
From The Relation:
VAB = VAC we have I1 × Rx = I2 × R2
(0.1)
At balanced ‘null’ position, since the galvanometer
carries no current, it as if acts as if open circuited,
thus

Thus, from Eq. (2.1), we have

81
(0.2)
Thus, measurement of the unknown resistance is
made in terms of three known resistances. The arms
BD and CD containing the fixed resistances R3 and
R4 are called the ratio arms. The arm AC containing
the known variable resistance R2 is called the
standard arm. The range of the resistance value that
can be measured by the bridge can be increased
simply by increasing the ratio R3/R4.
Errors in a Wheatstone Bridge
A Wheatstone bridge is a fairly convenient and
accurate method for measuring resistance. However,
it is not free from errors as listed below:

82
1) Discrepancies between the true and marked
values of resistances of the three known arms
can introduce errors in measurement.
2) Inaccuracy of the balance point due to
insufficient sensitivity of the galvanometer may
result in false null points.
3) Bridge resistances may change due to self-
heating (I2R) resulting in error in measurement
calculations.
4) Thermal emfs generated in the bridge circuit
or in the galvanometer in the connection
points may lead to error in measurement.
5) Errors may creep into measurement due to
resistances of leads and contacts. This effect is
however, negligible unless the unknown
resistance is of very low value.
6) There may also be personal errors in finding the
proper null point, taking readings, or during
calculations.

Errors due to inaccuracies in values of standard


resistors and insufficient sensitivity of galvanometer
can be eliminated by using good quality resistors and
galvanometer. Temperature dependent change of
resistance due to self-heating can be minimised by
performing the measurement within as short time as
possible.
Thermal emfs in the bridge arms may cause serious
trouble, particularly while measuring low resistances.
Thermal emf in galvanometer circuit may be serious in
some cases, so care must be taken to minimise those
effects for precision measurements. Some sensitive
83
galvanometers employ all-copper systems (i.e.,
copper coils as well as copper suspensions), so that
there is no junction of dissimilar metals to produce
thermal emf. The effect of thermal emf can be
balanced out in practice by adding a reversing switch
in the circuit between the battery and the bridge, then
making the bridge balance for each polarity and
averaging the two results.
Example 3-1
Four arms of a Wheatstone bridge are as follows: AB
= 100 Ω, BC = 10 Ω, CD = 4 Ω, DA = 50 Ω. A
galvanometer with internal resistance of 20 Ω is
connected between BD, while a battery of 10-V dc is
connected between AC. Find the current through the
galvanometer. Find the value of the resistance to be
put on the arm DA so that the bridge is balanced.
Solution
Configuration of the bridge with the values given in
the example is as shown below: To find out current
through the galvanometer, it is required to find out
Thevenin equivalent voltage across nodes BD and
also the Thevenin equivalent resistance between
terminals BD. To find out Thevenin’s equivalent
voltage across BD, the galvanometer is open
circuited, and the circuit then looks like the figure
given below.
At this condition, voltage drop across the arm BC is
given by

84
Hence, voltage difference between the nodes B and D,
or the Thevenin equivalent voltage between nodes B
and D is
VTH = VBD = VB – VD = VBC – VDC = 0.91 - 0.74 =
0.17 V
To obtain the Thevenin equivalent resistance between
nodes B and D, the 10 V source need to be shorted,
and the circuit looks like the figure given below.

85
The Thevenin equivalent resistance between the
nodes B and D is thus

Hence, current through galvanometer is

In order to balance the bridge, there should be no


current through the galvanometer, or in other words,
nodes B and D must be at the same potential.
Balance equation is thus

Example 3-2
The four arms of a Wheatstone bridge are as follows:
AB = 100 Ω, BC = 1000 Ω, CD = 4000 Ω, DA = 400 Ω. A
galvanometer with internal resistance of 100 Ω and
sensitivity of 10 mm/μA is connected between AC,
while a battery of 4 V dc is connected between BD.
Calculate the current through the galvanometer and
its deflection if the resistance of arm DA is changed
from 400 Ω to 401 Ω.
Solution
86
Configuration of the bridge with the values given in
the example is as shown below:
To find out current through the galvanometer, it is
required to find out the Thevenin equivalent voltage
across nodes AC and also the Thevenin equivalent
resistance between terminals AC.
To find out Thevenin equivalent voltage across AC,
the galvanometer is open circuited. At this condition,
voltage drop across the arm AB is given by

Voltage drop across the arm CB is given by

Hence, voltage difference between the nodes A and C,


or the Thevenin equivalent voltage between nodes A
and C is
VTH = VAC = VA – VC = VAB – VCB = 0.798 – 0.8 =
–0.002 V
To obtain the Thevenin equivalent resistance between
nodes A and C, the 10-V source need to be shorted.

87
Under this condition, the Thevenin equivalent
resistance between the nodes A and C is thus.

Hence, current through the galvanometer is

= Sensitivity × Current = 10 mm/µA = 2.04 µA = 20.4


mm

Sources and Detectors in AC Bridges


For measurements at low frequencies, bridge power
supply can be obtained from the power line itself.
Higher frequency requirements for power supplies are
normally met by electronic oscillators. Electronic
oscillators have highly stable, accurate yet adjustable
frequencies. Their output waveforms are very close to
sinusoidal and output power level sufficient for most
bridge measurements.
When working at a single frequency, a tuned detector
is preferred, since it gives maximum sensitivity at the
selected frequency and discrimination against
harmonic frequencies. Vibration galvanometers are
most commonly used as tuned detectors in the power
frequency and low audio-frequency ranges. Though
vibration galvanometers can be designed to work as
detectors over the frequency range of 5 Hz to 1000 Hz,
they have highest sensitivity when operated for
frequencies below 200 Hz. Head phones or audio
amplifiers are popularly used as balance detectors in
ac bridges at frequencies of 250 Hz and above, up to 3
to 4 kHz. Transistor amplifier with frequency tuning
facilities can be very effectively used as balance
detectors with ac bridges. With proper tuning, these
can be used to operate at a selective band of

88
frequencies with high sensitivity. Such detectors can
be designed to operate over a frequency range of 10
Hz to 100 kHz.
General Balance Equation for Four-Arm
Bridge
An ac bridge in its general form is shown in Figure
3.1, with the four arms being represented by four
unspecified impedances Z1, Z2, Z3 and Z4

Figure 0.2 General 4-arm bridge configuration


Balance in the bridge is secured by adjusting one or
more of the bridge arms. Balance is indicated by zero
response of the detector. At balance, no current flows
through the detector, i.e., there is no potential
difference across the detector, or in other words, the
potentials at points B and C are the same. This will be
achieved if the voltage drop from A to B equals the
voltage drop from A to C, both in magnitude and
phase.
Thus, we can write in terms of complex quantities:

or

89
(0.3)
Also at balance, since no current flows through the
detector,

(0.4)
and

(0.5)
Combining Eqs (3.4) and (3.5) into Eq. (3.3), we have

(0.6)
or

(0.7)
When using admittances in place of impedances, Eq.
(3.6) can be re-oriented as

(0.8)
Equations (3.6) and (3.8) represent the basic balance
equations of an ac bridge. Whereas (3.6) is convenient
for use in bridge configurations having series
elements, (3.8) is more useful when bridge
configurations have parallel elements.
Equation (3.6) indicates that under balanced
condition, the product of impedances of one pair of
opposite arms must be equal to the product of
impedances of the other pair of opposite arms, with
the impedances expressed as complex numbers. This
90
will mean, both magnitude and phase angles of the
complex numbers must be taken into account.
Re-writing the expressions in polar form, impedances
can be expressed as where Z represents the
magnitude and θ represents the phase angle of the
complex impedance.
If similar forms are written for all impedances and
substituted in (3.6), we obtain: Thus, for balance we
have,
Equation (3.9) shows that two requirements must be
met for satisfying balance condition in a bridge.

(0.9)

(0.10)
The first condition is that the magnitude of the
impedances must meet the relationship;

(0.11)
The second condition is that the phase angles of the
impedances must meet the relationship;
Example 3-3
In the AC bridge circuit shown in Figure 3.2, the
supply voltage is 20 V at 500 Hz. Arm AB is 0.25 mμ
pure capacitance; arm BD is 400 Ω pure resistance
and arm AC has a 120 Ω resistance in parallel with a
0.15 mμ capacitor. Find resistance and inductance or
capacitance of the arm CD considering it as a series
circuit.
Solution
Impedance of the arm AB is

91
Since it is purely capacitive, in complex notation,

Impedance of arm BD is Z3 = 400 Ω


Since it is purely resistive, in complex notation,

Impedance of arm AC containing 120 Ω resistance in


parallel with a 0.15 µF capacitor is

For balance,

Then impedance of arm CD required for balance is

or,

The positive angle of impedance indicates that the


branch consists of series combination of resistance
and inductance.
Resistance of the unknown branch

Inductance reactance of the unknown branch

Inductance of the unknown branch

92
Measurement of Self-Inductance
Maxwell’s Inductance Bridge
This bridge is used to measure the value of an
unknown inductance by comparing it with a variable
standard self-inductance. The bridge configuration
and phasor diagram under balanced condition are
shown in figure 3.3.
The unknown inductor L1 of resistance R1 in the
branch AB is compared with the standard known
inductor L2 of resistance R2 on arm AC. The inductor
L2 is of the same order as the unknown inductor L1.
The resistances R1, R2, etc., include, of course the
resistances of contacts and leads in various arms.
Branch BD and CD contain known no- inductive
resistors R3 and R4 respectively.
The bridge is balanced by varying L2 and one of the
resistors R3 or R4. Alternatively, R3 and R4 can be
kept constant, and the resistance of one of the other
two arms can be varied by connecting an additional
resistor.
Under balanced condition, no current flows through
the detector. Under such condition, currents in the
arms AB and BD are equal (I1). Similarly, currents in
the arms AC and CD are equal (I2). Under balanced
condition, since nodes B and D are at the same
potential, voltage drops across arm BD and CD are
equal (V3 = V4); similarly, voltage drop across arms
AB and AC are equal (V1 = V2).

93
Figure 0.3 Maxwell’s inductance bridge under
balanced condition: (a) Configuration (b) Phasor
diagram
As shown in the phasor diagram of Figure 3.3 (b), V3
and V4 being equal, they are overlapping. Arms BD
and CD being purely resistive, currents through these
arms will be in the same phase with the voltage drops
across these two respective branches. Thus, currents
I1 and I2 will be collinear with the phasors V3 and V4.
The same current I1 flows through branch AB as well,
thus the voltage drop I1R1 remains in the same phase
as I1. Voltage drop wI1L1 in the inductor L1 will be
90° out of phase with I1R1 as shown in Figure 3.3 (b).
Phasor summation of these two voltage drops I1R1
and wI1L1 will give the voltage drop V1 across the
arm AB. At balance condition, since voltage across
the two branches AB and AC are equal, thus the two
voltage drops V1 and V2 are equal and are in the same
phase. Finally, phasor summation of V1 and V3 (or
V2 and V4) results in the supply voltage V.
At balance,

or,

94
Equating real and imaginary parts, we have

or,

And also,

or,

Thus,
Unknown quantities can hence be calculated as

(0.12)
Care must be taken that the inductors L1 and L2 must
be placed at a distance from each other to avoid
effects of mutual inductance.
The final expression (3.12) shows that values of L1
and R1 do not depend on the supply frequency. Thus,
this bridge configuration is immune to frequency
variations and even harmonic distortions in the power
supply.
Maxwell’s Inductance–Capacitance Bridge
In this bridge, the unknown inductance is measured
by comparison with a standard variable capacitance. It
is much easier to obtain standard values of variable
capacitors with acceptable degree of accuracy. This is
however, not the case with finding accurate and stable
standard value variable inductor as is required in the
basic Maxwell’s bridge described in Section 3.4.1.

95
Configuration of a Maxwell’s inductance–capacitance
bridge and the associated phasor diagram at balanced
state are shown in Figure 3.4.

Figure 0.4 Maxwell’s inductance–capacitance bridge


under balanced condition: (a) Confi guration (b)
Phasor diagram
The unknown inductor L1 of effective resistance R1 in
the branch AB is compared with the standard known
variable capacitor C4 on arm CD. The other
resistances R2, R3, and R4 are known as non–
inductive resistors. The bridge is preferably balanced
by varying C4 and R4, giving independent adjustment
settings.
Under balanced condition, no current flows through
the detector. Under such condition, currents in the
arms AB and BD are equal (I1). Similarly, currents in
the arms AC and CD are equal (I2). Under balanced
condition, since nodes B and D are at the same
potential, voltage drops across arm BD and CD are
equal (V3 = V4); similarly, voltage drops across arms
AB and AC are equal (V1 = V2).

96
As shown in the phasor diagram of Figure 3.4 (b), V3
and V4 being equal, they are overlapping both in
magnitude and phase. The arm BD being purely
resistive, current I1 through this arm will be in the
same phase with the voltage drop V3 across it.
Similarly, the voltage drop V4 across the arm CD,
current IR through the resistance R4 in the same
branch, and the resulting resistive voltage drop IRR4
are all in the same phase [horizontal line in Figure
3.4(b)]. The resistive current IR when added with the
quadrature capacitive current IC, results in the main
current I2 flowing in the arm CD. This current I2 while
flowing through the resistance R2 in the arm AC,
produces a voltage drop V2 = I2R2, that is in same
phase as I2. Under balanced condition, voltage drops
across arms AB and AC are equal, i.e., V1 = V2. This
voltage drop across the arm AB is actually the phasor
summation of voltage drop I1R1 across the resistance
R1 and the quadrature voltage drop wI1L1 across the
unknown inductor L1. Finally, phasor summation of
V1 and V3 (or V2 and V4) results in the supply voltage
V.
At balance,

or,

Equating real and imaginary parts, we have

or,

And also,

or,
97
Thus, unknown quantities can hence be calculated as

(0.13)
Once again, the final expression (3.13) shows that
values of L1 and R1 do not depend on the supply
frequency. Thus, this bridge confi guration is immune
to frequency variations and even harmonic distortions
in the power supply.
It is interesting to note that both in the Maxwell’s
Inductance Bridge and Inductance- Capacitance
Bridge, the unknown Inductor L1 was always
associated with a resistance R1. This series
resistance has been included to represent losses that
take place in an inductor coil. An ideal inductor will be
lossless irrespective of the amount of current flowing
through it. However, any real inductor will have some
non-zero resistance associated with it due to
resistance of the metal wire used to form the inductor
winding. This series resistance causes heat
generation due to power loss. In such cases, the
Quality Factor or the Q-Factor of such a lossy
inductor is used to indicate how closely the real
inductor comes to behave as an ideal inductor. The Q-
factor of an inductor is defined as the ratio of its
inductive reactance to its resistance at a given
frequency. Q-factor is a measure of the efficiency of
the inductor. The higher the value of Q-factor, the
closer it approaches the behavior of an ideal, loss less
inductor. An ideal inductor would have an infinite Q at
all frequencies.
The Q-factor of an inductor is given by the formula
, where R is its internal resistance R (series
resistance) and ɷL is its inductive reactance at the
frequency ω.

98
Q-factor of an inductor can be increased by either
increasing its inductance value (by using a good
ferromagnetic core) or by reducing its winding
resistance (by using good quality conductor material,
in special cases may be super conductors as well).
In the Maxwell’s Inductance-Capacitance Bridge, Q-
factor of the inductor under measurement can be
found at balance condition to be , or,

(0.14)
The above relation (3.14) for the inductor Q factor
indicate that this bridge is not suitable for
measurement of inductor values with high Q factors,
since in that case, the required value of R4 for
achieving balance becomes impracticably high.
Advantages of Maxwell’s Bridge
1) The balance equations (3.13) are independent of
each other, thus the two variables C4 and R4 can
be varied independently.
2) Final balance equations are independent of
frequency.
3) The unknown quantities can be denoted by
simple expressions involving known quantities.
4) Balance equation is independent of losses
associated with the inductor.
5) A wide range of inductance at power and audio
frequencies can be measured.
Disadvantages of Maxwell’s Bridge

99
1) The bridge, for its operation, requires a standard
variable capacitor, which can be very expensive if
high accuracies are asked for. In such a case,
fixed value capacitors are used and balance is
achieved by varying R4 and R2.
2) This bridge is limited to measurement of low Q
inductors (1< Q < 10).
3) Maxwell’s bridge is also unsuited for coils with
very low value of Q (e.g., Q < 1).
Such low Q inductors can be found in inductive
resistors and RF coils. Maxwell’s bridge finds difficult
and laborious to obtain balance while measuring such
low Q inductors.
Hay’s Bridge
Hay’s bridge is a modification of Maxwell’s bridge.
This method of measurement is particularly suited for
high Q inductors.
Configuration of Hay’s bridge and the associated
phasor diagram under balanced state are shown in
Figure 3.5.
The unknown inductor L1 of effective resistance R1 in
the branch AB is compared with the standard known
variable capacitor C4 on arm CD. This bridge uses a
resistance R4 in series with the standard capacitor C4
(unlike in Maxwell’s bride where R4 was in parallel
with C4). The other resistances R2 and R3 are known
no-inductive resistors.
The bridge is balanced by varying C4 and R4.
Under balanced condition, since no current flows
through the detector, nodes B and D are at the same
potential, voltage drops across arm BD and CD are
equal (V3 = V4); similarly, voltage drops across arms
AB and AC are equal (V1 = V2).

100
Figure 0.5 Hay’s bridge under balanced condition: (a)
Confi guration, (b) Phasor diagram
As shown in the phasor diagram of Figure 3.5 (b), V3
and V4 being equal, they are overlapping both in
magnitude and phase and are draw on along the
horizontal axis. The arm BD being purely resistive,
current I1 through this arm will be in the same phase
with the voltage drop V3 = I1R3 across it. The same
current I1, while passing through the resistance R1 in
the arm AB, produces a voltage drop I1R1 that is once
again, in the same phase as I1. Total voltage drop V1
across the arm AB is obtained by adding the two
quadrature phasors I1R1 and wI1L1 representing
resistive and inductive voltage drops in the same
branch AB. Since under balance condition, voltage
drops across arms AB and AC are equal, i.e., (V1 =
V2), the two voltages V1 and V2 are overlapping both
in magnitude and phase. The branch AC being purely
resistive, the branch current I2 and branch voltage V2
will be in the same phase as shown in the phasor
diagram of Figure 3.5 (b). The same current I2 flows
through the arm CD and produces a voltage drop I2R4
across the resistance R4. This resistive voltage drop
I2R4, obviously is in the same phase as I2. The
capacitive voltage drop I2/wC4 in the capacitance C4

101
present in the same arm AC will however, lag the
current I2 by 90°. Phasor summation of these two
series voltage drops across R4 and C4 will give the
total voltage drop V4 across the arm CD. Finally,
phasor summation of V1 and V3 (or V2 and V4) results
in the supply voltage V.
At balance,

or,

Equating real and imaginary parts, we have

(0.15)
and,

(0.16)
Solving Eqs (3.15) and (3.16) we have the unknown
quantities as

(0.17)
and

(0.18)
Q factor of the inductor in this case can be calculated
at balance condition as

(0.19)
Hay’s bridge is more suitable for measurement of
unknown inductors having Q factor more than 10. In

102
those cases, bridge balance can be attained by
varying R2 only, without losing much accuracy.

(0.20)
From (3.17) and (3.19), the unknown inductance value
can be written as For inductors with Q > 10, the
quantity (1/Q)2 will be less than 1/100, and thus can be
neglected from the denominator of (3.20). In such a
case, the inductor value can be simplifi ed to L1 =
R2R3C4, which essentially is the same as obtained in
Maxwell’s bridge.
Owen’s Bridge
This bridge is used for measurement of unknown
inductance in terms of known value capacitance.
Figure 3.6 shows the Owen’s bridge configuration and
corresponding phasor diagram under balanced
condition. The unknown inductor L1 of effective
resistance R1 in the branch AB is compared with the
standard known capacitor C2 on arm AC. The bridge
is balanced by varying R2 and C2 independently.
Under balanced condition, since no current flows
through the detector, nodes B and C are at the same
potential, i.e., V1 = V2 and V3 = V4.

103
Figure 0.6 Owen’s bridge under balanced condition:
(a) Confi guration (b) Phasor diagram
with I1. The inductive voltage drop ɷI1L1 when added
in quadrature with the resistive voltage drop I1R1
gives the total voltage drop V1 across the arm AB.
Under balance condition, voltage drops across arms
AB and AC being equal, the voltages V1 and V2
coincide with each other as shown in the phasor
diagram of Figure 3.6 (b). The voltage V2 is once again
summation of two mutually quadrature voltage drops
I2R2 (resistive) and I2/ɷC2 (capacitive) in the arm
AC. It is to be noted here that the current I2 leads the
voltage V4 by 90° due to presence of the capacitor C4.
This makes I2 and hence I2R2 to be vertical, as shown
in the phasor diagram. Finally, phasor summation of
V1 and V3 (or V2 and V4) results in the supply voltage
V.
At balance,

Simplifying and separating real and imaginary parts,


the unknown quantities can be found out as

104
(0.21)
and

(0.22)
It is thus possible to have two independent variables
C2 and R2 for obtaining balance in Owen’s bridge. The
balance equations are also quite simple. This
however, does come with additional cost for the
variable capacitor.
Example 3-4
A Maxwell’s inductance–capacitance bridge is used to
measure a unknown inductive impedance. The bridge
constants at bridge balance are: Pure resistance arms
= 2.5 kΩ and 50 kΩ. In between these two resistors,
the third arm has a capacitor of value 0.012 μF in
series with a resistor of value 235 kΩ. Find the series
equivalent of the unknown impedance.
Solution
Referring to the diagram of a Maxwell’s inductance–
capacitance bridge: Using the balance equation,

and,
105
Example 3-5
The four arms of a bridge are connected as follows:
 Arm AB: A choke coil L1 with an equivalent series
resistance r1
 Arm BC: A noninductive resistance R3
 Arm CD: A mica capacitor C4 in series a
noninductive resistance R4
 Arm DA: A noninductive resistance R2
When the bridge is supplied from a source of 450 Hz
is given between terminals A and C and the detector is
connected between nodes B and D, balance is
obtained the following conditions: R2 = 2400 Ω, R3 =
600 Ω, C4 = 0.3 μF and R4 = 55.4 Ω. Series resistance
of the capacitor is 0.5 Ω. Calculate the resistance and
inductance of the choke coil.
Solution
The bridge configuration is shown below: Given that
at balance,

At balanced,
or,

106
Equating real and imaginary parts, we have
(i)
(ii)
Solving (i) and (ii), we have the unknown quantities as

and

Measurement of Capacitance
Bridges are used to make precise measurements of
unknown capacitances and associated losses in
terms of some known external capacitances and
resistances. An ideal capacitor is formed by placing a
piece of dielectric material between two conducting
plates or electrodes. In practical cases, this dielectric
material will have some power losses in it due to
dielectric’s conduction electrons and also due to
dipole relaxation phenomena. Thus, whereas an ideal
capacitor will not have any losses, a real capacitor will
have some losses associated with its operation. The
potential energy across a capacitor is thus dissipated
in all real capacitors as heat loss inside its dielectric
material. This loss is equivalently represented by a
series resistance, called the equivalent series
resistance (ESR). In a good capacitor, the ESR is very
small, whereas in a poor capacitor the ESR is large.
Creal Cideal ESR
A real, lossy capacitor can thus be equivalently
represented by an ideal loss less capacitor in series

107
CR with its equivalent series resistance (ESR)
shown Figure 3.7
Equivalent series resistance (ESR) in Figure 2.7.

Figure 0.7 Equivalent series resistance (ESR)

The quantifying parameters often used to describe


performance of a capacitor are ESR, its dissipation
factor (DF), Quality Factor (Q-factor) and Loss
Tangent (tan d). Not only that these parameters
describe operation of the capacitor in radio frequency
(RF) applications, but ESR and DF are also particularly
important for capacitors operating in power supplies
where a large dissipation factor will result in large
amount of power being wasted in the capacitor.
Capacitors with high values of ESR will need to
dissipate large amount of heat. Proper circuit design
needs to be practiced so as to take care of such
possibilities of heat generation.
Dissipation factor due to the non-ideal capacitor is
defined as the ratio of the resistive power loss in the
ESR to the reactive power oscillating in the capacitor,
or

When representing the electrical circuit parameters as


phasors, a capacitor’s dissipation factor is equal to
the tangent of the angle between the capacitor’s
impedance phasor and the negative reactive axis, as
shown in the impedance triangle diagram of Figure 3.8
This gives rise to the parameter known as the loss
tangent d where Figure. 3.8 Impedance

108
Figure 0.8 Impedance triangle diagram
Loss tangent of a real capacitor can also be defined in
the voltage triangle diagram of Figure 3.9 as the ratio
of voltage drop across the ESR to the voltage drop
across the capacitor only, i.e. tangent of the angle
between the capacitor voltage only and the total
voltage drop across the combination of capacitor and
ESR.

Figure 0.9 Voltage triangle diagram


Though the expressions for dissipation factor (DF)
and loss tangent (tan δ) are the same, normally the
dissipation factor is used at lower frequencies,
whereas the loss tangent is more applicable for high
frequency applications. A good capacitor will normally
have low values of dissipation factor (DF) and loss
tangent (tan δ).
In addition to ESR, DF and loss tangent, the other
parameter used to quantify performance of a real

109
capacitor is its Quality Factor or Q-Factor.
Essentially for a capacitor it is the ratio of the energy
stored to that dissipated per cycle.

It can thus be deduced that the Q can be expressed as


the ratio of the capacitive reactance to the ESR at the
frequency of interest.

A high quality capacitor (high Q-factor) will thus have


low values of dissipation factor (DF) and loss tangent
(tan δ), i.e. less losses.
The most commonly used bridges for capacitance
measurement are De Sauty’s bridge and Schering
Bridge.
De Sauty’s Bridge
This is the simplest method of finding out the value of
a unknown capacitor in terms of a known standard
capacitor. Configuration and phasor diagram of a De
Sauty’s bridge under balanced condition is shown in
Figure 3.10.

110
Figure 0.10 De Sauty’s bridge under balanced
condition: (a) Confi guration (b) Phasor diagram
The unknown capacitor C1 in the branch AB is
compared with the standard known standard
capacitor C2 on arm AC. The bridge can be balanced
by varying either of the non-inductive resistors R3 or
R4. Under balanced condition, since no current flows
through the detector, nodes B and C are at the same
potential, i.e., V1 = V2 and V3 = V4.
As shown in the phasor diagram of Figure 3.7 (b), V3 =
I1R3 and V4 = I2R4 being equal both in magnitude and
phase, they overlap. Current I1 in the arm BD and I2 in
the arm CD are also in the same phase with I1R3 and
I2R4 along the horizontal line. Capacitive voltage drop
V1 = I1/wC1 in the arm AB lags behind I1 by 90°.
Similarly, the other capacitive voltage drop V2 =
I2/wC2 in the arm AC lags behind I2 by 90°. Under
balanced condition, these two voltage drops V1 and
V2 being equal in magnitude and phase, they overlap
each other along the vertical axis as shown in Figure
3.7 (b). Finally, phasor summation of V1 and V3 (or V2
and V4) results in the supply voltage V.
At balance,

or,

(0.23)

The advantage of De Sauty’s bridge is its simplicity.


However, this advantage may be nullifi ed by
impurities creeping in the measurement if the
capacitors are not free from dielectric losses. This
method is thus best suited for loss-less air capacitors.

111
In order to make measurement in capacitors having
inherent dielectric losses, the modified De Sauty’s
bridge as suggested by Grover, can be used. This
bridge is also called the series resistance-capacitance
bridge. Configuration of such a bridge and its
corresponding phasor diagram under balanced
condition is shown in Figure 3.11.

Figure 0.11 Modified De Sauty’s bridge under


balanced condition: (a) Configuration, and (b) Phasor
diagram

The unknown capacitor C1 with internal resistance r1


representing losses in the branch AB is compared
with the standard known standard capacitor C2 along
with its internal resistance r2 on arm AC. Resistors R1
and R2 are connected externally in series with C1 and
C2 respectively. The bridge can be balanced by
varying either of the non-inductive resistors R3 or R4.
Under balanced condition, since no current flows
through the detector, nodes B and C are at the same
potential, i.e., V1 = V2 and V3 = V4.
As shown in the phasor diagram of Figure 3.10 (b), V3
= I1R3 and V4 = I2R4 being equal both in magnitude
and phase, they overlap. Current I1 in the arm BD and
I2 in the arm CD are also in the same phase with I1R3
112
and I2R4 along the horizontal line. The other resistive
drops, namely, I1R1 in the arm AB and I2R2 in the arm
AC are also along the same horizontal line. Finally,
resistive drops inside the capacitors, namely, I1r1 and
I2r2 are once again, in the same phase, along the
horizontal line. Capacitive voltage drops I1/wC1 lags
behind I1r1 by 90°. Similarly, the other capacitive
voltage drop I2/wC2 lags behind I2r2 by
90°. Phasor summation of the resistive drop I1r1 and
the quadrature capacitive drop I1/ wC1 produces the
total voltage drop VC1 across the series combination
of capacitor C1 and its internal resistance r1.
Similarly, phasor summation of the resistive drop I2r2
and the quadrature capacitive drop I2/wC2 produces
the total voltage drop VC2 across the series
combination of capacitor C2 and its internal
resistance r2. d1 and d2 represent loss angles for
capacitors C1 and C2 respectively. Phasor summation
of I1R1 and VC1 gives the total voltage drop V1 across
the branch AB. Similarly, phasor summation of I2R2
and VC2 gives the total voltage drop V2 across the
branch AC. Finally, phasor summation of V1 and V3
(or V2 and V4) results in the supply voltage V.
At balance,

or,

Equating real and imaginary parts, we have

(0.24)

(0.25)

113
The modified De Sauty’s bridge can also be used to
estimate dissipation factor for the unknown capacitor
as described below:
Dissipation factor for the capacitors are defined as

From Eq. (3.22), we have

or,

or,

Using Eq. (6.31), we getor, D -D =w (CR -CR )

Substituting the value of C1 from Eq. (3.25), we have

(0.26)
Thus, dissipation factor for one capacitor can be
estimated if dissipation factor of the other capacitor is
known.
Schering Bridge
Schering bridges are most popularly used these days
in industries for measurement of capacitance,
dissipation factor, and loss angles. Figure 3.12
illustrates the confi guration of a Schering bridge and
corresponding phasor diagram under balanced
condition.

114
Figure 0.12 Schering bridge under balanced condition:
(a) Configuration (b) Phasor diagram

The unknown capacitor C1 along with its internal


resistance r1 (representing loss) placed on the arm
AB is compared with the standard loss-less capacitor
C2 placed on the arm AC. This capacitor C2 is either
an air or a gas capacitor to make it loss free. R3 is a
non-inductive resistance placed on arm BD. The
bridge is balanced by varying the capacitor C4 and the
non-inductive resistor R4 parallel with C4, placed on
arm CD.
Under balanced condition, since no current flows
through the detector, nodes B and C are at the same
potential, i.e., V1 = V2 and V3 = V4. As shown in the
phasor diagram of Figure 3.12 (b), V3 = I1R3 and V4 =
IRR4 being equal both in magnitude and phase, they
overlap. Current I1 in the arm BD and IR flowing
through R4 are also in the same phase with I1R3 and
IRR4 along the horizontal line. The other resistive
drop namely, I1R1 in the arm AB is also along the
same horizontal line. The resistive current IR through
R4 and the quadrature capacitive current IC through
C4 will add up to the total current I2 in the branch CD
(and also in A C under balanced condition). Across

115
the arm AB, the resistive drop I1r1 and the quadrature
capacitive drop I1/wC1 will add up to the total voltage
drop V1 across the arm. At balance, voltage drop V1
across arm AB will be same as the voltage drop V2 =
I2/wC2 across the arm AC. It can be confirmed from
the phasor diagram in Figure 3.12(b) that the current
I2 has quadrature phase relationship with the
capacitive voltage drop I2/wC2 in the arm AC. Finally,
phasor summation of V1 and V3 (or V2 and V4) results
in the supply voltage V.
At balance,

or,

Equating real and imaginary parts, we have the


unknown quantities:

(0.27)
and,

(0.28)
Dissipation Factor

(0.29)
Thus, using Schering bridge, dissipation factor can be
obtained in terms of the bridge parameters at balance
condition.

116
Measurement
Laboratory
Experiments

First Year
117
2024

EXPERIMENT 1
ERRORS IN MEASUREMENTS AND BASIC
STATISTICAL SAMPLING

OBJECTIVES
To investigate sources of error in measurements,
To observe the value of statistical analysis.

EQUIPMENT & COMPONENTS


1. 30 composition resistors of the same color-coded value.
2. 1 digital multimeter.

BACKGROUND
Error is defined as the deviation of a reading (or set of
readings) from the expected value of the measure variable.
When we make measurements, some error is inevitable
because no measurement can yield the exact value of any

118
quantity. There are several sources of error in any
experimental data. The primary concerns about analyzing
experimental data are the sources of error and the extent to
which the error has affected the validity of the data.
The error of measurement consists of three major
components:
1. Gross error,
2. Systematic error,
3. Random error.

In order to eliminate gross error, good measurement


practice must be applied. Always obey the rules of the
measurement, do not relay on only one measurement, do
many measurements and judge the results.
Systematic errors are resulted from the measuring
instruments (Instrumental error) and external conditions
(Environmental error).
A statistical analysis is performed on samples of very
large quantities of the measurements to determine the
probable variation in values of the entire measurement, which
is resulted in Random errors. The percentage of the entire
measurement, which will fall within a specific range of values,
can be predicted quite accurately from the statistical analysis
of the sample.
Under ideal conditions, a very large number of
measurements will provide a distribution of readings, with the
greatest number of readings approximately equal to the actual
value. On either side of the actual value, the frequency of
readings will decrease, producing an approximately normal
distribution as shown in Figure 1.1. Arithmetic mean of the n
measurement is the best estimate of the true value. It can be
formulated as:

Where n is the number of measurement, xi is the reading of


the i’th measurement. Standard Deviation or root-mean-
square deviation of the n measurement is the best estimate of
the precision. It can be formulated as:

119
where n is the number of measurement, di is the deviation
from the mean

PROCEDURE
□ 1. Measure the value of each resistors by digital multimeter.
Record the values in the Table 1.1 (Resistance values).
□ 2. Select 8 resistors at random from the total lot and
measure and record their values in Table 1.1 (Sample 1).
□ 3. Mix all the resistors together and select, at random, any
12 resistors. Your selection may or may not include
resistors from the previous sample. Measure and record
the values of the 12 resistors in Table 1.1 (Sample 2).
□ 4. Mix all the resistors together and select, at random, a
sample of 16 resistors and measure an record their vales
in Table 1.1 (Sample 3).
□ 5. On one sheet of graph paper, plot the value of each
resistor in the total lot and make a bar graph, or
histogram.
□ 6. Divide a second sheet of graph paper three ways
vertically. Plot a histogram (resistance values versus
frequency) for each of the three samples of the resistors.
□ 7. Connect the maximum points of each histogram by a
smooth curve. If the numbers of resistors in the samples
120
were much larger, this would give an approximate normal
distribution curve such as the one shown in Figure 1.1.
However, your curves may be skewed because samples
are small.
□ 8. Compute and record in Table 1.1 the standard deviation σ
for the entire lot and for each of the three samples.
□ 9. Record in Table 1.2 which sample (1, 2, or 3) most neatly
describes the total lot with regard to standard deviation.

121
122
EXPERIMENT 2

DC CURRENT AND VOLTAGE


MEASUREMENT

OBJECTIVES
To learn how to measure DC voltages and current through the
circuit,
To learn how to use multimeter properly.

EQUIPMENT & COMPONENTS


1. Digital Multimeter.
2. Analog Multimeter.
3. DC Power Supply (12 V).
4. Resistors (100Ω, 1kΩ, 10kΩ.)

BACKGROUND
The two most important commonly used quantities are
the current and the voltage. The current is a serial quantity
and measured by using ampermeter. The voltage is defined
between two nodes and measured by connecting a voltmeter
across those two nodes.

Figure 2.1 (a) The electronic circuit, (b) measuring current and
(c) measuring voltage.

123
Current is measured by connecting the ammeter in
series to the resistor; the voltage is measured by connecting
the voltmeter in parallel to it.
In order to eliminate the loading effect, the internal
resistance of the ammeter should be very small and the
resistance of the voltmeter should be very high compared to
the circuit resistance. Otherwise the ammeter or the voltmeter
changes the circuit operation conditions and an error is
introduced.
The error of measurement consists of three major
components:
1. Gross error,
2. Systematic error,
3. Random error.
In order to eliminate gross error, good measurement
practice must be applied. Always obey the rules of the
measurement, do not relay on only one measurement, do
many measurements and judge the results.
Systematic errors are resulted from the measuring
instruments (Instrumental error) and external conditions
(Environmental error). Instrumental error is given in the user’s
manual of the instrument for maximum reading:

where ε0 is the error or accuracy of the instrument. Error of


the analogue instrument is usually expressed as the class of
the instrument. The class of the instrument shows the relative
error for full-scale deflection (maximum reading).
Lading error is the other type of instrumental error.
When measuring system is connected to the system to be
measured, some loading effects happen due to the power
sharing of the two systems. The loading error for an ammeter
and voltmeter can be calculated as follows, respectively.

124
where r is the internal resistance of the meter and Rn is the
Norton and Rth Thevenin equivalent resistance of the circuit.
Since the internal resistance of analog voltmeter change with
the selected range, instead of the internal resistance usually
the sensitivity of the instrument, S (input resistance-per-volt),
is given such as “20kΩ/V”. This value is indicated on the
panel of the instrument. If the range of the voltmeter is set to
VR, then the internal resistance of analog voltmeter is:

The internal resistance of a digital voltmeter is usually


constant and greater than 1MΩ. The internal resistance of an
ammeter changes with the range of the instrument and should
be obtained from the users manual of the instrument.

The relative worst case or limiting error of a


measurement is the sum of the loading error and instrument
(accuracy) error assuming other errors are negligible:

125
PROCEDURE
□ 1. Calculate the current and the voltages across the
resistors R1 and R2 of Figure 2.2 for (a) R2=10kΩ and (b)
R2=100Ω.

Figure 2.2

□ 2. Connect the circuit given in Figure 2.2. Set R2=10kΩ.


Measure the exact value of the voltage source, using the
digital voltmeter.
□3. Open one end of R1. Connect the analog ammeter in
series and measure the current I1.
□4. Open one end of R2. Connect the analog ammeter in
series and measure the current I2.
□5. Connect the analog voltmeter across the resistor R1 and
measure V1.
□6. Connect the analog voltmeter across the resistor R2 and
measure V2.
□7. Repeat steps 3 to 6 using digital multimeter.
□8. Set R2= 100Ω. Repeat steps 3 to 7.
□9. Calculate the errors of measurements and fill in the Table
2-1, 2-2.

126
127
EXPERIMENT 3

RESISTOR CHARACTERISTICS
AND OHM’S LAW

OBJECTIVES
To learn how to use ohmmeter properly,
To study characteristics of resistance,
To learn how to use Ohm’s Law in circuit analysis.

EQUIPMENT & COMPONENTS


1. Digital Multimeter
2. Adjustable DC Power Supply
3. Potentiometer
4. Resistors

BACKGROUND
All materials possess electrical resistance which is the
opposition to the flow of electrical current in a circuit. The unit
of measure for electrical resistance is the ohms (Ω). One ohm
may be defined as the electrical resistance of a copper wire
which is 300 m long and 2.5 mm in diameter. The instrument
used to measure electrical resistance is called an ohmmeter.
The ohmmeter must be connected to any circuit element
under no power conditions. The resistors can be simply
divided into two types: fixed resistors and variable resistors.
The fixed resistor has two terminals and its resistance is
constant. A variable resistor (VR) or potentiometer has three
terminals and its resistance is variable.
The circuit symbol of a variable resistor is shown in
Figure 3.1. The three terminals are the end terminals A,C and
a wiper terminal B. The resistance between the end terminals
RAC is fixed and is always equal to its nominal value. The
wiper resistances between the wiper terminal and the end
terminals, RAB and RBC, are variable. The wiper resistances
correspond to a given position of the potentiometer shaft.

128
Figure 3.1 Symbol of a variable resistor

Ohm’s law, discovered by a German physicist Simon


Ohm (1787 – 1854), is an important law that describes the
relationship of voltage E to current I and resistance R. It is
often referred to as the foundation of circuit analysis and can
be expressed by three different ways:

where E is the potential difference from one end of a


resistance element to the other (volt), I is the current through
the same resistance element (amperes), R is the resistance of
the same element (ohm).
Remember that lowering the resistance raises the
current, and raising the voltage also raises the current.

PROCEDURE
□ 1. Using the ohmmeter measure the resistance of resistors
given to you and record the results in the column of
measured value in Table 3.1.
□ 2. Read the resistor values by using color codes. Compare
the measured values with the reading values and
129
tolerances for determining whether each measured value
is within the tolerance or not. Complete Table 3.1.

□ 3. Designate three terminals of VR1 on the KL-21001 as A


(the right), B (the middle) and C (the left).
□ 4. Using the ohmmeter, measure the resistance between
terminals A and C and record the value on Table 3.2.
□ 5. Turn the VR1 control knob completely to the left. Measure
and record the resistance between terminals A and B, and
A and C on Table 3.2.
□ 6. Turn the VR1 control knob completely to the right.
Measure and record the resistance between terminals A
and B, and A and C on Table 3.2.
□ 7. Turn the VR1 control knob to middle position. Measure
and record the resistance between terminals A and B, and
A and C on Table 3.2.
□ 8. Set the circuit shown in Figure 3.2 by using R1 = 100Ω
and Vs = 12 V. Using Ohm’s

Figure 3.2.

□ 9. Connect the ammeter to the circuit shown in Figure 3.2.


Measure the current through resistor R1. and record the
result on the Table 3.3 Is there good agreement
between your measured and calculated current values?
□ 10.Change Vs as you read 150 mA on the ammeter. Measure
the voltage across the resistor R1 by using voltmeter.
Calculate and record the voltage value (V) on the Table
3.3. Is there good agreement between your measured
and calculated voltage values?
□ 11. It is simple to build an equivalent ammeter by
connecting a known resistor in parallel with a
130
voltmeter, see Figure 3.3. According to this build your
own ammeter with RA = 4.7 Ω. To measure the current in
this ammeter you measure the voltage (V) and by using
the Ohm’s Law you calculate the current through RA (I

=V/RA ). You assume that the internal resistance of

voltmeter is very high and loading effect is negligible.

Figure 3.3

□ 11. Connect your ammeter to the circuit shown in Figure 3.2


V (R1 = 100Ω and Vs = 12 V.). Calculate the current
through the resistor R1 by considering your ammeter
connected to the circuit. Record the result on the Table
3.4.
□ 12.Measure the current through resistor R1 and record the
result on the Table 3.4. Is there good agreement between
your measured and calculated current values?

131
132
EXPERIMENT 4

OSCILLOSCOPE

OBJECTIVES

To learn how to use oscilloscope properly,


To measure amplitude, frequency and phase angle by using
oscilloscope.

EQUIPMENT & COMPONENTS


1. Dual Trace Oscilloscope
2. Digital Multimeter
3. Signal Generator
4. Resistors
5. Capacitor

BACKGROUND
The cathode ray oscilloscope is the most versatile instrument
to measure electrical quantities. It is possible to use an
oscilloscope to measure the following quantities of a voltage;

Figure 4.1. Sinusoidal wave.


1. Instantaneous value, v(t)

133
2. The positive and negative peak value, Vp+, Vp-
3. The peak-to-peak value, Vpp
4. The average value, VA
5. The period, T
6. The phase difference between two voltages, φ.

Probe:
The input impedance of an oscilloscope is very high.
1MΩ parallel 20pF input impedance is an industry standard
for oscilloscopes. To compensate the large input capacitance
and to increase the input resistance of an oscilloscope it is
necessary to use a serial cable and circuit that is called
“PROBE”. The probe is made of a co-axial shielded cable and
compensated attenuator as shown in Figure 4.2.

Figure 4.2. Oscilloscope probe.


The large input time constant R2C2 creates a low pass
filter (with single pole) and attenuates all high frequency (HF)

134
signals. To compensate this “pole” it is necessary to create a
zero at the same frequency by R1C1, satisfying:

To obtain x10 attenuation R1 = 9MΩ is selected. It is


necessary to adjust either C1, to obtain the equality, since C2
is not exactly known. This adjustment must be done once, for
every oscilloscope and every probe connected to the specific
input of this oscilloscope.
Note: Compensation works only on x10 position.
Therefore HF signals can only be used by using a
compensated x10 probe.

Measurement using an Oscilloscope


Since it has a very large input resistance, an
oscilloscope can only be used to measure voltages and time.
The vertical axes of an oscilloscope are calibrated to indicate
volts/cm and the horizontal axes are calibrated as s/cm.

Instantaneous and peak value:


Measuring the first three quantities, Instantaneous, Peak
and Peak-to-Peak Voltage value is straightforward. Be sure
that zero level adjustment is done properly and DC or AC
input selection is correctly set.

Average value:
To measure the average value first the input connection
is made AC coupled and one extreme of the waveform is set
to the reference line. Then the input is set to Dc coupled mode
and the voltage shift is measured. This shift is equal to the
average value of the voltage.

Period:
The period can also be directly measured by measuring
the time difference between two zero crossing points, and
frequency can be easily calculated by taking the reciprocal of
the period.

Phase:
The phase measurement can be performed in two different
ways.

135
1. If the oscilloscope has dual trace facility two voltages are
displayed simultaneously and the time difference, tD, is
measured between two identical points of the waveforms.
The phase difference is then:

where T is the period of the signal.


2. If one of the signals applied to the horizontal input while the
other is connected to vertical input an elliptical trace is
obtained on the screen as shown in Figure 4.3. The phase
difference can be calculated using this trace as follows:

Figure 4.3. The Lissajou curve.

PROCEDURE
□ 1. Read the short instruction for the oscilloscope
operations.
□ 2. Set all push buttons to OUT position. Then turn on the
oscilloscope according to its procedures given in data
sheet.
□ 3. Connect the x10 probe to the CAL output of the
oscilloscope. Adjust the TIME/DIV. and VOLT/DIV. knobs
to obtain a suitable square wave on the screen. Then
136
adjust the compensation capacitor by using a suitable
screwdriver. Try to obtain an ideal square wave.

□ 4. Apply 1 kHz sine wave to the INP.-I input of the


oscilloscope. Adjust the TIME/DIV. to obtain
approximately one full period, and set VOLTS/DIV.-I to
0.2V/cm and adjust the output of the signal generator to
obtain 1.2V peak-to-peak amplitude. Measure the period
of the waveform and calculate the frequency.
□ 5. Measure RMS output voltage of the signal generator
using the digital voltmeter. Calculate peak-to-peak
amplitude. Compare the result with the oscilloscope.
□ 6. Connect the circuit in Figure 4.4. using R = 1kΩ and C =
100nF. Connect node A to vertical input CH.I and node B
to vertical input CH.II. Adjust VOLTS/DIV.-II to obtain
reasonable amplitude. Measure the time difference
between the zero crossing points of the waveforms and
calculate the phase difference for (i) f=100Hz, (ii) f=1kHz,
(iii) f=10kHz.

Figure 4.4.

□ 7. Push “Hor/MENU” button and obtain the Lissajous


pattern. Measure the phase using Equation 4.3 for (i)
f=100Hz, (ii) f=1kHz, (iii) f=10kHz.

137
138
EXPERIMENT 5

AC VOLTAGE MEASUREMENT

OBJECTIVES
To learn how to measure AC voltages,
To become familiar with the use of AC voltmeters.
EQUIPMENT & COMPONENTS
1. Dual Trace Oscilloscope
2. Digital Multimeter
3. Analog Multimeter
4. Signal Generator.
BACKGROUND
The important parameters of a time varying AC voltage,
v(t), to be measured are the
following;
1. Instantaneous value, v(t)
2. The positive and negative peak value, Vp+, Vp-
3. The peak-to-peak value, Vpp
4. The average value, VA
5. The RMS value, VRMS
5. The period, T
6. The frequency, f.

Figure 5.1. Sinusoidal wave.


The oscilloscope is the most versatile instrument to
measure most of these electrical quantities. An oscilloscope
139
can measure all of the above mentioned properties except the
RMS value and the frequency. The frequency may be easily
calculated after measuring the period since f = 1/T .
The phase difference between two voltages may be also
measured by an oscilloscope.

Measurement using a Voltmeter:


To measure the average value, VA, of an AC signal a DC
voltmeter may be used directly. An AC voltmeter must be
used for measuring the RMS value. However, an AC voltmeter
can work properly only in a limited frequency range for which
it is designed. Outside of this range, the measurement results
will be wrong. Therefore, for AC measurements, it is important
to know the frequency of the signal to be measured and the
frequency characteristics of the voltmeter to be used.
The waveform also effects the measurement’s result.
Ordinary AC voltmeters are designed and calibrated to
measure the RMS values of sinusoidal signals only. For other
waveforms the result is not usually correct.
The RMS value of a waveform is defined as follows:

It is fairly difficult to perform the squaring operation to


obtain the true RMS value. Only expensive voltmeters
measure the true RMS value. The ordinary voltmeters
measure the absolute average of a voltage, which is defined:

as the average of the absolute value. The absolute average


value can be easily be obtained by a full wave rectifying the
input voltage. The conversion from the absolute average to
the RMS value can be done by scaling the readout (or the
input voltage) by a factor called the shape factor. But a Shape
Factor is only valid for a certain waveform. For a sinusoidal
waveform this factor is 1.11.

140
After measuring the Absolute Average voltage, the RMS
value can be easily determined if the waveform is known. For
complex waveforms, such as speech and noise, it is not
possible to determine the shape factor and it is necessary to
use a True RMS voltmeter.
AC Voltmeter must be connected in parallel with the
terminals of the circuit elements whose ac voltages will be
measured. Besides the polarity, AC voltmeters use the same
rules as DC voltmeters do. Since AC voltage reverses its
polarity periodically, AC voltmeters are therefore designed
without limit in polarity.
Measurement using an Oscilloscope:
Since it has a very large input resistance, an
oscilloscope can only be used to measure voltages and time.
The vertical axes of an oscilloscope are calibrated to indicate
volts/cm and the horizontal axes are calibrated as s/cm.
Instantaneous and peak values:
Measuring the first three quantities, Instantaneous, Peak
and Peak-to-Peak Voltage value is straightforward. Be sure
that zero level adjustment is done properly and DC or AC
input selection is correctly set.
Average value:
To measure the average value first the input connection
is made AC coupled and one extreme of the waveform is set
to the reference line. Then the input is set to Dc coupled mode
and the voltage shift is measured. This shift is equal to the
average value of the voltage.
Period:
The period can also be directly measured by measuring
the time difference between two zero crossing points, and
frequency can be easily calculated by taking the reciprocal of
the period.
Phase:
The phase measurement can be performed in two
different ways.
1. If the oscilloscope has dual trace facility two voltages are
displayed simultaneously and the time difference, tD, is
measured between two identical points of the waveforms.
The phase difference is then:

where T is the period of the signal.


141
2. If one of the signals applied to the horizontal input while the
other is connected to vertical input an elliptical trace is
obtained on the screen as shown in Figure 5.3. The phase
difference can be calculated using this trace as follows:

Figure 5.2. The Lissajous curve.


CALCULATIONS
Calculate the shape factor of a triangular and square
wave. The standard AC voltmeters are calibrated to indicate
the RMS value of a sine wave. Using the calculated shape
factors calculate the voltage to be displayed by an absolute

142
average reading meter when the input is 10 V peak (i) sine, (ii)
square and (iii) triangular waveform.
Calculate the theoretical RMS values for the procedure
steps 4, 5 and fill in the blanks in table.
PROCEDURE
□ 1. Read the short instruction for the oscilloscope given in
appendix 1. Set all push buttons to OUT position.
□ 2. Apply 1 kHz sine wave to the INP.-I input of the
oscilloscope. Adjust the TIME/DIV. to obtain
approximately one full period, and set VOLTS/DIV.-I to
2V/cm and adjust the output of the signal generator to
obtain 12V peak-to-peak amplitude. Measure the period
of the waveform and calculate the frequency.
□ 3. Connect the analog and digital voltmeters and the
oscilloscope to the signal generator as shown in Figure
5.3. Set the waveform to Sine wave, output voltage to
12V peak-to-peak on the oscilloscope. Change the
frequency and note the voltmeter readings in Table 5.1
(Check the voltage on the oscilloscope to remain
constant, 12V, peak-to-peak for every frequency).

Figure 5.3
□ 4. Set the frequency to 50 Hz. Adjust the output voltage to
10V peak-to-peak on the oscilloscope. Then read the
voltage with digital and analog voltmeter. Compare the
result with the theoretical values. Complete Table 5.2.
□ 5. Change the signal to square and triangular waveforms
then repeat step-4. Complete Table 5.3.
□ 6. Draw the frequency response of the voltmeters, using the
measured values (Table 5.1), for the sinusoidal input, on
a lin-log graph-paper.

143
144
EXPERIMENT 6

MEASUREMENT USING DC BRIDGES

OBJECTIVES
To study resistance measurement techniques using DC
Bridge circuits

EQUIPMENT & COMPONENTS


1. Digital Multimeter
2. 12V DC power supply
3. Potentiometer
4. Resistors.

BACKGROUND
Accurate measurements of resistances may be
performed by using impedance-measuring Bridges. There are
a number of bridges, which are called usually by their
inventor’s name, to measure different type of resistances and
impedances.

Typical bridge circuit is given in Figure 6.1.

Figure 6.1. The basic impedance bridge.


When the equation

145
is satisfied, the voltages of nodes A and B are equal and the
current of the detector, is zero.
The unknown impedance, Z4 is:

For measuring real impedances, i.e. resistors, all


impedances are resistors and the bridge is called Wheastone
Bridge. For measuring capacitance, inductance or complex
impedances at least one of the Z1, Z2, Z3 must also be complex
in order to satisfy the balance equation.

Resistance Measurement using the Wheastone Bridge


The basic bridge circuit called Wheastone Bridge is
suitable to measure medium range resistance. A DC voltage
source and DC meter may be used in this case, since the
impedances to be used are real.

Figure 6.2 Wheastone Bridge.

For the resistive case the Equation 6.2 becomes:

146
For the maximum sensitivity of the bridge all resistance
values should be in the same range (as close to each other as
possible) as a result of maximum power transfer theorem.

Resistance Measurement using the Kelvin Double Bridge

Figure 6.3. The Kelvin Double Bridge.

The Kelvin Double Bridge is a modification of the


Wheatstone Bridge and provides greatly increased accuracy
in the measurement of low-value resistances, generally below
1Ω. The term double bridge is used because the circuit
contains a second set of ratio arms.

The unknown resistance Rx can be calculated as follows:

147
where Ry is the yoke resistance measured between the node
connecting R3 and b and the node connecting Rx and a.
PROCEDURE
□ 1. Connect the Wheatstone bridge circuit shown in Figure
6.2. Set R1=R2= 1kΩ. Use an unknown resistor with a
relatively low resistance value, a potentiometer for R3 and
12V DC power supply for the voltage source. Connect a
DC voltmeter between nodes A-B as the detector.
□ 2. By adjusting Rv obtain the balance of the bridge.
□ 3. Calculate the value of unknown resistance Rx
□ 4. Replace R1=R2= 100Ω and repeat steps 2, 3.
□ 5. Connect the Kelvin double bridge circuit shown in Figure
6.3. Set R1=R2=R3=a=b= 1kΩ and an unknown resistor
having a relatively low resistance value. Connect a 12V
DC power supply for the voltage source.
□ 6. Calculate the value of unknown resistance Rx.

EXPERIMENT 7

MEASUREMENT USING AC BRIDGES

148
OBJECTIVES
To study impedance measurement techniques using AC
Bridge circuits

EQUIPMENT & COMPONENTS


1. Digital Multimeter.
2. Signal Generator.
3. Potentiometers.
4. Resistors, Capacitors and Inductors.

BACKGROUND
Accurate measurements of complex impedances and
frequencies may be performed by using
impedance-measuring AC Bridges. There are a number of
bridges, which are called usually by
their inventor’s name, to measure different types of
impedances and frequencies.

Typical bridge circuit is given in Figure 7.1.

Figure 7.1. The basic impedance bridge.


When the equation

is satisfied, the voltages of nodes A and B are equal and the


current of the detector, is zero. The unknown impedance, Z4
is:

149
For measuring capacitance, inductance or complex
impedances at least one of the Z1, Z2, Z3 must also be
complex in order to satisfy the balance equation.

Capacitance Measurement using the Schering Bridge

Figure 7.2. The Schering Bridge.

The Schering Bridge is a complex impedance bridge to


measure capacitance. The complex balance equation yields
two real equation. Using these two equations the unknown
capacitance
and its equivalent series resistor (which corresponds to the
losses of the capacitor) can be calculated as follows:

150
The error of measurement can be calculated as:

The accuracy of the detector does not effect the error of


measurement. But its sensitivity determines the minimum
detectable difference between the voltages of the nodes A and
B, hence, minimum detectable variation of adjustable
components, ΔCxdet. The best sensitivity is obtained if the
values of the impedances of the arms of the bridge on the
operating frequency are close to each other.

The losses of a capacitor can either be represented by a


shunt or series equivalent resistors.

The Schering Bridge measures the series equivalent


capacitor and resistor. To obtain the parallel equivalent
capacitor and leakage resistor (representing the losses):

151
The result of measurement should be given as:

Inductance Measurement using the Maxwell Bridge


The Maxwell Bridge is a complex impedance bridge to
measure an unknown inductance in terms of a known
capacitance. The complex balance equation yields two real
equation. Using these two equations the unknown inductance
and the resistor in series with it can be calculated as follows:

Figure 7.3 Maxwell Bridge.

152
respectively.

PROCEDURE

□ 1. Connect the Schering bridge circuit shown in Figure 7.2.


Set C3 to 47nF, R1 to 1kΩ. and oscillator frequency to
1kHz. Connect an AC voltmeter between nodes A-B as the
detector.
□ 2. Set C1 as 22nF and by adjusting R2 obtain the balance of
the bridge. Can you balance the bridge? Then change C1
to 47nF and adjust R2 to obtain balance of the bridge.
Can you balance the bridge?
□ 3. Calculate the value of unknown capacitor Cx and
equivalent resistance Rx.
□ 4. Calculate the value of unknown capacitor Cp and
equivalent resistance Rp.
□ 5. Set the frequency of the generator to 50 Hz and repeat
step 3 and 4.
□ 6. Connect the Maxwell bridge circuit shown in Figure 7.2.
Set C1 to 47nF, R2 to 1kΩ. and oscillator frequency to
1kHz. Connect an AC voltmeter between nodes A-B as
the detector.
□ 7. By adjusting R1 and R3 obtain the balance of the bridge.
Can you balance the bridge?
□ 8. Calculate the value of unknown capacitor Lx and
resistance Rx.
□ 9. Calculate the value of unknown capacitor Lp and
resistance Rp.
□ 10. Set the frequency of the generator to 50 Hz and repeat
step 8 and 9.

153
154
References

1- Electronic Measurements and Measurements, 2nd


edition, Larry D. Jones and A. Foster Chin.
2- Introduction to Instrumentation and
Measurements, 2nd edition, Robert B. Northrop,
2005.
3- Measurement & Error, S. Z. Sayed Hassen, Dept.
of Electrical & Electronic Engineering, University
of Mauritius, January 30, 2006.
4- Measurement and Control Basics, 3rd Edition,
Thomas A. Hughes, 2002.
5- Electrical measurements, T. Norby, Department
of Chemistry, University of Oslo.
6- https://en.wikipedia.org/wiki/Attenuator_(elec
tronics)
7- https://www.elprocus.com/what-is-an-
attenuator-design-types-and-applications/
8- https://www.connectortips.com/faq-what-are-
attenuators/
9- https://www.watelectronics.com/what-is-an-
attenuator-design-types-its-applications/

155

You might also like