You are on page 1of 246

Pressure Measurement

Randall Anderson
Noel Bignell (Ed.)
Kitty MK Fen
Walter J Giardini
John KL Man
Edwin C Morris
David B Prowse
Dennis Tyrrell
Brad R Ward

Monograph 7: NML Technology Transfer Series


Second Edition

National Measurement Laboratory


D-06843-P
.g7

Pressure Measurement

Randall Anderson
Noel Bignell (Ed.)
Kitty MK Fen
Walter J Giardini
John KL Man
Edwin C Morris
David B Prowse
Dennis Tyrrell
Brad R Ward

Monograph 7: NML Technology Transfer Series


Second Edition

National Measurement Laboratory CSIRO

BMLSTORGEN
AP 336 i
The monograph is one of a series that is being developed to give practical advice on the
measurement of physical quantities. Current monographs are:
Monograph 1: “Uncertainty in Measurement: the ISO Guide”
Monograph 2: “Statistical Background to the ISO Guide to the Expression of
Uncertainty in Measurement”
Monograph 3: “Traceable Measurements”
Monograph 4: “The Calibration of Weights and Balances”
Monograph 5: “Thermocouples in Temperature Measurement”
Monograph 6: “The Measurement of Electrical Quantities”
Monograph 7: “Pressure Measurement”
Monograph 8: “Humidity and its Measurement”
Monograph 9: “Liquid-in-Glass Thermometry”

with details given on NML’s website: http://www.nml.csiro.au .

National Measurement Laboratory CSIRO, September 2003

ISBN 0-9750744-6-6

Copies of the monographs are available from


National Measurement Laboratory
Bradfield Road (PO Box 218)
Lindfield. NSW, Australia
(phone; -f6 1-2-9413 7459)

© CSIRO Australia, 2003


Printed in Sydney

11
iii
iv
Contents
1. An overview of pressure measurement 1
1.1. Introduction 1
1.2. Definition of pressure 3
1.3. Units of pressure 3
1.4. Realisation of the unit - primary standards 4
1.5. Dissemination of the units of pressure 7
1.6. Types of pressure-measuring instruments 8
1.6.1 Vacuum instruments 8
1.6.2 Manometric instruments 10
1.6.3 Industrial pressure measurement 12
1.6.4 High-pressure measurement 12
1.7. Conclusion 12
1.8. Appendix 13
1.8.1 The international system of units - SI 13
1.8.2 Units of pressure outside the SI system 14
1.9. References 15

2. Vacuum pressure measurement 17


2.1. Introduction 17
2.2. Vacuum ranges and characteristics 18
2.2.1 Viscous flow 18
2.2.2 Molecular flow 20
2.2.3 Thermal transpiration 21
2.2.4 Gas laws 22
2.3. Vacuum pumps 23
2.3.1 Rotary pump 23
2.3.2 Roots pump 23
2.3.3 Sorption pump andvapour trap 24
2.3.4 Diffusion pump 25
2.3.5 Ion pump 26
2.3.6 Turbo-molecular pump 26
2.3.7 Cryo-pump 26
2.4. Pressure measurement - rough and medium vacuum 26
2.4.1 Mechanical gauges 26
2.4.2 Thermal conductivity gauges 31
2.5. Pressure measurement - high vacuum region 33
2.5.1 Gauge types 33
2.5.2 Partial pressure measurement 36
2.5.3 Primary calibration techniques 36
2.5.4 Comparison calibration techniques 37
2.6. Pressure measurement in ultra-highvacuum 38
2.7. References 39

3. Barometry, manometry and liquid columns 41


3.1. Introduction 41
3.2. Mercury barometers 42
3.2.1 History 42
3.2.2 Standard conditions 42
3.2.3 Pressure units 43
3.3. Types of barometers 43
3.3.1 Fortin barometer 45
3.3.2 U-tube barometers and monometers 45
3.3.3 Micrometer barometer 46
3.3.4 Siphon barometer 46
3.3.5 Kew or ‘fixed-cistern’ barometer 47
3.4. Reading Fortin and Kew barometers 48
3.5. Pressure measurement using Fortin andKew barometers 49
3.5.1 Temperature correctionand measurement 49
3.5.2 Correction for gravity 51

VI
3.5.3 Capillary errors 51
3.5.4 Elevation correction 52
3.5.5 Estimation of air in the vacuum spaces 53
3.6. Calibration of mercury barometers 55
3.6.1 Frequency of calibration 55
3.6.2 Comparison with a remote barometer 56
3.6.3 Checking mercury barometers in-situ 58
3.7. Uncertainty in calibrating a Fortin barometer in-situ 58
3.7.1 The model 59
3.7.2 The components of uncertainty 60
3.8. Care and handling Fortinand Kew barometers 62
3.8.1 Preparation for transport 62
3.8.2 Preparation for measurement 63
3.9. Aneroid barometers 63
3.10 Manometers 64
3.10.1 Micromanometers 64
3.10.2 Inclined-tube manometers 68
3.10.3 Gas manometer 69
3.11 Properties of mercury 69
3.12 Properties of water 70
3.13 Other manometric fluids 71
3.14 References 72

4. The gas-operated piston gauge and its uses 73


4.1. Introduction 73
4.2. Gas-operated piston gauges 74
4.2.1 Variant designs 76
4.3. Evaluation of pressure 80
4.3.1 Force 80
4.3.2 Effective area 86
4.3.3 Fluid head corrections 88
4.3.4 Final expression forpressure 89

vii
4.4. Calibration of gas-operated piston gauges 91
4.4.1 Rate of fall method 91
4.4.2 Differential-pressure method 92
4.4.3 Data evaluation 93
4.4.4 Uncertainty evaluation 93
4.4.5 NML Reports 94
4.5. Principal uses of piston gauges 94
4.5.1 Calibration of pressure gauges 95
4.5.2 Calibration of pressure transducers 95
4.5.3 Calibration of ball and nozzle gauges 96
4.6. Practical notes on the uses of piston gauges 96
4.7. Calculation of uncertainty 98
4.7.1 Uncertainty in pressure for a gas-operated piston gauge 98
4.7.2 Uncertainty in calibration for a Bourdon pressure gauge 102
4.7.3 Uncertainty in calibration for a pressure transducer 106
4.8. Appendix. Sample measurement reports 109
4.8.1 report of calibration for a piston cylinder and weight set 109
4.8.2 Report of calibration for a bal and nozzle piston gauge 114
4.9. References 118

5. The hydraulic piston gauge and its use 119


5.1. Introduction 119
5.2. Basic theory 122
5.2.1 Models of piston gauge operation 122
5.3. Use of the pressure balance 126
5.4. Loads 127
5.5. Operational considerations 135
5.6. Calibration of pressure-measurement devices 137
5.7. Cross-float comparison calibration of piston gauges 138
5.8. Uncertainty of generated pressure and device calibration 149
5.9. Appendix. Buoyancy correction for immersed parts 151
5.10. References 154

viii
6. Pressure transducers 155
6.1. Introduction 155
6.2. The transducer as a black box 155
6.2.1 Ideal relationship between measurand and output 156
6.2.2 Dynamic response and frequency response 156
6.3. Real transducers 159
6.3.1 Zero error 159
6.3.2 Linearity 159
6.3.3 Hysteresis 168
6.4. Change of characteristics - instability 162
6.4.1 Aging 163
6.4.2 Temperature stability 163
6.4.3 Excitation 163
6.4.4 Other factors affecting stability 164
6.4.5 Repeatability or instability 164
6.5. Calibration and pressure standards 164
6.5.1 Pressure transducers 164
6.5.2 Ancillary equipment 165
6.5.3 Static pressure calibration 166
6.5.4 Dynamic pressure calibration 168
6.6. Pressure transducers using elastically-deformable elements 170
6.6.1 Diaphragms 170
6.7. Transducers not using elastic deformation 178
6.7.1 Force feedback or force balances 178
6.7.2 Pressure-sensitive resistance transducers 179
6.7.3 Semiconductors 180
6.7.4 Piezo-electric transducers 180
6.7.5 Pressure-sensitive capacitive transducers 180
6.8. Example specification sheet 180
6.8.1 ‘NML’ industrial pressure transmitters 180
6.8.2 Other items covered in specifications 184
6.9. Pressure transducers in pressure measurement 184

IX
6.10. Uncertainty of pressures measured with a pressure transducer 185
6.11. Example report on a pressure transducer 188
6.12. References 191

7. Differential pressure measurement 193


7.1. Introduction 193
7.2. Differential pressure measurement using liquid columns 193
7.2.1 Inclined tube 193
7.2.2 Pressure correction 194
7.3. Differential pressure transducers 194
7.4. Standards of differential pressure at high line pressures 195
7.4.1 Two separate pressurestandards 195
7.4.2 Two balanced piston gauges 196
7.4.3 NPL design 197
7.4.4 Mercury columns 199
7.5. Calibration of differential pressuretransducers 199
7.6. References 200

8. Pressure calibrators 201


8.1. Introduction 201
8.2. Functional design of pressure calibrators 201
8.3. Desired outcomes of a calibration 202
8.4. Current standard test procedures 202
8.5. Statement of compliance 202
8.6. A guide for testing a digital pressure calibrator 203
8.6.1 test design 203
8.6.2 Test procedure 203

9. Measurement and usesof high pressure 211


9.1. Introduction 211
9.2. High-pressure standards 211
9.3. Fixed points 214
9.4. Transducers 215

X
9.5. High-pressure generation and equipment 217
9.5.1 Pumps 217
9.5.2 Tubing 219
9.5.3 Pressure vessels 219
9.5.4 Seals 219
9.5.5 Valves 220
9.5.6 Pressure fluids 221
9.6. Pressures greater than 2 GPa 221
9.6.1 Standards 221
9.6.2 Fixed points 222
9.6.3 Pressure generation 222
9.6.4 Pressure measurement 222
9.6.5 Dynamic pressures 223
9.7. Safety 223
9.8. Industrial applications of high pressure 224
9.8.1 Chemical synthesis 225
9.8.2 Crystal growth 225
9.8.3 Production of superhard materials 225
9.8.4 Hydrostatic extrusion 225
9.8.5 Differential hydrostatic extrusions 226
9.8.6 Isostatic pressing 226
9.8.7 Cutting materials with high-pressure jets 226
9.8.8 Explosive forming 226
9.8.9 Explosive welding 226
9.8.10 Autofrettage 227
9.9. References 228

XI
xii
Chapter 1

An overview of pressure measurement


Noel Bignell

1.1 Introduction
The word “pressure” predates the technical concept of pressure as it will be used in this course.
When the Oxford Dictionary quotes an example of the use of the word in 1602, “The soft
pressure of a melting kisse”, it is not of the Bourdon tube or barometer that we must think, but
rather, the wine press, used to force the juice from grapes by pressing them.

Even today there is confusion in the public mind between the concepts of pressure and force,
as, indeed, between force and weight. Until the middle of the seventeenth century this
confusion was quite general and great men, of considerable insight in many matters, such as
Galileo, were quite wrong in their ideas concerning the pressure of air. It is curious today to
realise that Galileo, though he believed in the existence of a vacuum, did not believe in the
pressure of air or water.

The vacuum was thought to be logically impossible by Aristotle, and his teachings came to
have the status of dogma. It was perhaps as much to throw off the teachings of Aristotle as for
any other reason that Galileo embraced the idea of a vacuum. He completely failed to give the
correct explanation to the observation of the failure of water siphons to work above a certain
height. In about 1640 Berti performed some important experiments with long sealed tubes of
water and found that the water column could not be made to exceed a height of about 10 m. It
then began to be accepted that air exerted a pressure that supported such columns of liquid, but
only to a certain height. Torricelli thought to use glass tubes filled with mercury, and found
that these could measure the pressure of the air. Thus, Torrecelli invented the barometer in
1644.

In France, Pascal (born June 19, 1623) heard of Torricelli’s experiments from Petit, who had
repeated them. Pascal was pleased because he, like Galileo, had believed in the vacuum for a
long time, on largely philosophical grounds. However, even then Pascal was not clearly aware
of the role that atmospheric pressure played in these experiments. In 1647 he was still more
concerned with the vacuum, which he thought was perfect, than with atmospheric pressure,
and there is some doubt as to whether he or Descartes suggested the famous experiment
performed by Perier on 19 September 1648, in which a Torricellian tube (a barometer in other
CHAPTER L OVERVIEW OF PRESSURE MEASUREMENT

words) was taken up the Puy-de-Dome, a mountain in the Auvergnes. The result showed quite
clearly and. apparently, surprisingly to the experimenter, that the air pressure decreased as the
mountain was climbed.

That such great thinkers as Pascal and Galileo did not sooner proclaim the concept of pressure
shows that there is some difficulty in the idea. Part of the difficulty is that atmospheric
pressure surrounds us, but a confusion that commonly occurs comes from the failure to
distinguish pressure from force. An essential feature of pressure, realised by Pascal shortly
after the barometer was taken up the Puy-de-Dome, and which has come to be called Pascal’s
Principle, states that pressure, unlike force, is not directed and manifests itself in all directions.
Thus, every square millimetre of surface in contact with a fluid under pressure is pushed by it.
If the surface has a large area then there is a large push or force and this is the principle of the
hydraulic press, shown in Figure 1.1.

The concept of pressure once conceived was soon part of science and in 1662 Boyle stated that
for a given sample of gas the pressure multiplied by the volume was a constant. The constant
depended on the quantity of gas present in the sample. In other words pressure, P, is
proportional to 1/V, where V is the volume of the gas. We had to wait for the development of
the concept of temperature for further progress in gas laws. Curiously, though Galileo had not
a good idea of the concept of pressure he is credited with the invention of the thermometer.
However, a temperature scale is more than a thermometer and it was not until the development
of a scale that Charles was able to develop his Law. This he stated in 1787. In modern terms it
is that the volume of a fixed sample of gas held at constant pressure is proportional to its
absolute temperature. These two laws combined give the Ideal Gas Law which states that
PV/T is a constant, where T is the absolute temperature or the Celsius temperature plus 273.15
and its unit is the kelvin. Room temperature is thus about 300 K.

Figure 1.1: Two piston gauges of different effective areas can be connected together as
shown and put in a balanced condition if the load on each piston is in
proportion to its area. Used in reverse, this arrangement becomes an
hydraulic press.

2
NBIGNELL

1.2 Definition of pressure


The modern definition of the magnitude of a pressure is the force per unit area that it exerts on
any surface. This definition can be seen to work most clearly in the piston gauge (or dead­
weight tester), which is a well-fitting piston in a well-made cylinder. Referring to Figure 1.1,
the downward force on the small piston is that due to gravity acting on the weights, which is
Mg, where M is the mass and g is the acceleration due to gravity. The pressure, P, comes
directly from the definition as

P = force/area = Mg!A
where A is the area of the cross-section of the piston. If we compare a piston in a cylinder with
a column of mercury we can see that we might regard the column as a frictionless piston in a
cylinder, in which case its mass would be pAh, where again A is the area of cross-section of
the column, p is the density and h is the height. Substituting this mass into the formula for the
piston gauge we have

P = pAhglA = pgh.
That the area of the column does not change the pressure can be demonstrated by a U-tube
where a thin arm will balance a thick one of equal height.

Part of the difficulty experienced by Galileo and others in accepting the existence of
atmospheric pressure was probably due to the apparent lack of pressure exerted by it. It is only
by reference to a vacuum that a true measure of the pressure of the atmosphere can be made.
When pressure is measured by reference to a vacuum it is called absolute pressure, because the
vacuum provides an absolute reference. Pressures measured with reference to the atmosphere,
which varies from time to time and from place to place, are called gauge pressures, the
pressure measured by an ordinary gauge. The ‘gauge pressure’ of the atmosphere is, indeed,
zero, and in this respect Galileo was right, the atmosphere has no pressure.

All pressure measurements are, in fact, measurements of differences of pressure or differential


pressure, and absolute and gauge pressure are just special cases of differential pressure. The
latter term is, however, often used to mean small differences in pressure between two large
pressures. The accurate measurement of these small pressures is one of the most difficult tasks
in pressure measurement.

1.3 Units of pressure


From the definition of pressure we can immediately see that the unit of pressure is the unit of
force divided by the unit of area, which for the SI system is the newton divided by the metre
squared, (the Appendix contains a short account of the SI system of units). This is given the
special name pascal in honour of the great French philosopher already mentioned. Note that it
is not spelt with a capital letter when written in full but, when abbreviated to Pa, the capital is
used. This is a general rule for SI units that take their names from great scientists of the past.

The unit of pressure is a derived unit. That means it is derived from other, already defined,
quantities, and thus its value is fixed by their values. It must have the magnitude of the unit of

3
CHAPTER 1. OVERVIEW OF PRESSURE MEASUREMENT

force divided by the unit of area. The unit of area, of course, must have the magnitude of the
unit of length squared and the unit of force must be the unit of mass times the unit of
acceleration, from Newton’s law. The unit of acceleration must have the magnitude of the unit
of velocity divided by the unit of time and the unit of velocity is itself the unit of length
divided by the unit of time. The quantities mass, (A/), length, (L) and time (7) are basic and
their units are called base units. Pressure in terms of (L), (A/) and (7) is seen to have the
dimensions of mass per length per time squared {ML'^T^}. This is true for all pressure units
even for the millimetre of mercury (mmHg) to be discussed shortly.

The pascal may be too large or too small for any particular purpose and multiples and sub­
multiples are used, as with other units, as follows:

Nanopascal, nPa, 10*’ Pa


Micropascal, pPa, 10’^ Pa
Millipascal, mPa, 10'3 Pa
Hectopascal hPa, 10-Pa
Kilopascal, kPa 10’ Pa
Megapascal, Mpa 10^ Pa
Giga pascal. Gpa 10’ Pa

Many other units have been, and are, used for pressure measurement, usually for some
historical reason. A list of some of these, and conversion factors is given in the Appendix.

1.4 Realisation of the unit - primary pressure standards


To make a pressure measurement we need equipment that must be checked against pressure
standards that must themselves be checked against better standards. It is clear we must stop
this process somewhere at an instrument that does not need to be checked, but which can
measure pressure without calibration by another pressure instrument. Such a device is called a
primary pressure standard. The unit of pressure must be realised, i.e., made real, by such an
apparatus that accurately relates it to the base units. This apparatus has its characteristics
measured only in terms of length, mass and time, that is, in terms of the base units. There are
two main ways of doing this, one goes back to Torricelli and his column of mercury and the
other goes back even further to the notion of pressing, the trick being to press without wasting
any force in friction. A piston gauge may be made virtually friction less by rotation of the
piston in its cylinder so that a film of fluid is maintained between the two. The right- or left­
hand side of Figure 1.1 could be such a gauge and the pressure, F, produced by them will be
equal if

P = Mg!A = mgla.

So we can compare piston gauges this way without ever needing to know P (or g, since this
cancels). The area A (or a} may be measured by measuring the diameter of the piston, or
should we measure the internal diameter of the cylinder. They are not the same because,
though the piston is close fitting, there must be a film of fluid between the two. A simple
answer is to average the two results, but more sophisticated and more accurate treatments are

4
NBIGNELL

available. Another complication arises at higher pressures, which can cause the cylinder to
stretch a bit and so increase the cross-sectional area. There is a solution to this problem,
several in fact, and it is left to articles suggested in the References to explain them.

As previously noted, a column of liquid exerts a pressure pgh and a unit of pressure, often used
in vacuum work, is the torr, in honour of Torricelli, which is the pressure of a column of
mercury at 0°C that is one millimetre high and at a gravity value of 9.80665 ms'". The
temperature is specified because the density of mercury changes with temperature. This is
often called simply the millimetre of mercury, abbreviated as “mmHg”. It is not an SI unit; the
SI value in pascals is obtained by working out the value of pgh with p = 13,595.1 kgm '^ and g
= 9.80665 ms'" and so its value is 133.322 Pa.

Thus, a column of mercury, as shown in Figure 1.2, with a means to measure the height of the
column and the temperature of the mercury, is a primary pressure standard - it does not need
to be calibrated against another pressure standard. Of course the length measuring device
needs to be calibrated and the value of g needs to be measured. Someone also needs to have
measured the density of mercury as a function of the temperature and we must use a calibrated
thermometer to measure the temperature of the mercury column so that we can calculate the
density. The scale may also change with temperature, so its temperature needs to be measured
too, so that a correction may be made. Thus, an apparently simple device becomes quite
complex in practice, particularly for accurate measurements.

Figure 1.2: A primary standard mercury column showing a means to measure the
heights of the two mercury surfaces, the mercury temperature and the scale
temperature.

5
CHAPTER 1. OVERVIEW OF PRESSURE MEASUREMENT

Many mercury columns are used as pressure standards. To measure higher pressures we
naturally require longer tubes and the highest ever constructed appears to be that which was
installed in 1891 up the western pillar of the Eiffel tower to the summit, a height of 300 m (see
Figure 1.3). This was used by Louis Cailletet (1832-1913) for the calibration of Bourdon
Tubes, the ordinary pressure gauge still used today, in which he found errors of up to 12%.
The maximum pressure attainable by this instrument was about 400 times atmospheric
pressure or about 40 MPa. Longer columns were constructed by Amagat by going down
mines and not up towers. He reached pressures of 300 MPa.

Figure 1.3: The path of the mercury column constructed in the Eiffel tower and used by
Cailletet to obtain, for the first time, accurate pressures up to 40 MPa.

6
NBIGNELL

It is possible to reach higher pressures in a smaller tower by using a property of the piston
gauge. The piston gauge can reproduce and maintain a pressure even though it may not be
able to measure it. By putting a piston gauge at the bottom of a mercury column, as shown in
Figure 1.4a, its pressure can be made equal to that of the column by adjustment of the weights
on it. The column pressure is known from its height, as discussed above. Then, the same
piston gauge is connected to the top of the column and made to reproduce the previous
pressure while a second piston gauge is placed at the bottom, as shown in Figure 1.4b. This
second gauge now has a pressure applied equal to twice the column pressure and it may,
subsequently, be placed at the top and the first one brought to the bottom. By continuing this
process very high pressures may be reached without the use of a very long column. There are
many technical difficulties, the mercury temperature and its compressibility must be known, to
obtain the density, and the measurement of column length is not easy at high pressures, where
glass tubes cannot be used.

(a) (b)

Figure 1.4: A piston gauge may be balanced against a mercury column (a), and then
used to apply the same pressure to the top of the column whilst a second
piston gauge balances the combined pressure of column and first piston
gauge, (b). The process may be repeated many times to simulate the pressure
of a very tall mercury column.

1.5 Dissemination of the unit of pressure


Establishing a pressure unit is for the sole purpose of measuring pressures in the real world,
and pressure is one of the most widely measured quantities. Its range of magnitudes is
enormous; from the pressure in the picture tubes of television sets to the pressure used in the

7
CHAPTER 1. OVERVIEW OF PRESSURE MEASUREMENT

manufacture of industrial diamonds represents about 16 orders of magnitude. Such a wide


range demands to be divided up and it is convenient to speak of at least four regions. These
may be extended at either end by adding prefixes “high” and “ultra” at the lower end and
“very” and “super” at the high end. We shall confine ourselves to:

(a) vacuum, up to 100 Pa


(b) manometric, from 100-10^ Pa
(c) industrial, from 10'*’-10^ Pa
(d) high pressure. from 10^ Pa upwards.

To cover such a wide range of pressures, there are many different types of instruments in use,
but these are, in the main, secondary or tertiary instruments. Secondary standards are
calibrated directly from primary standards and are used to calibrate tertiary instruments. The
terms are not quite as rigid as that because obviously a primary standard could be used to
calibrate a tertiary standard directly and a secondary standard can calibrate another secondary
standard, but the basic idea is that there is a hierarchy of standards. The term “traceability” [1]
is often used to indicate a properly documented chain of instruments, back to a primary
instrument, which is, usually, in a National Measurement Institute. An instrument, other than a
primary standard, cannot properly be said to be calibrated unless it has traceability. For
example, a Bourdon gauge used to measure steam pressure in a boiler would have been
calibrated using a test gauge, a superior sort of Bourdon gauge, which would have been
calibrated in the factory laboratory against a dead-weight tester (a piston gauge for industrial
use), which would have been calibrated in an external laboratory against a piston gauge acting
as a secondary standard that was in its turn calibrated against a primary standard in a national
laboratory. Sometimes there is an additional calibration step between the primary standard and
the secondary standard involving another secondary standard. All this is shown in Figure 1.5,
which shows some other steps in a national pressure measurement system.

Obviously, every step in such a system means an increase in uncertainty, just how much
depends on the care taken and the instruments involved. This uncertainty increase down the
chain of pressure instruments from the primary standard sometimes means that it is hard to
achieve sufficiently low uncertainty. The lowest uncertainty is achieved for pressures in the
barometric region and rises towards both the vacuum and high pressure regions. An idea of
the accuracy attainable in the different pressure regions is given in Figure 1.6.

1.6 Types of pressure-measuring instruments


Each of the regions into which the domain of pressure measurement has been divided has its
own instrumentation, though there is some overlap to adjacent regions.

1.6.1 Vacuum instruments

Vacuum pressures are always in a gas, and the properties of gases can be used to advantage in
the measurement of the pressure. For example, gases are compressible, and, when compressed,
their pressure increases so that if we compress a gas to an accurate one-tenth its volume then
its pressure will rise ten times and may then be high enough to measure with a liquid column.

8
NBIGNELL

Figure 1.5: A hierarchy of pressure standards and some of their inter-relationships.

This is the principle of the McLeod gauge, a primary standard, which uses a rising mercury
column to compress the gas. Some standards generate pressures in this range by a reverse
process. An easily measured pressure in a small volume is allowed to expand into a larger
volume. If the ratio of the volumes is known then the final pressure can be calculated from

9
CHAPTER L OVERVIEW OF PRESSURE MEASUREMENT

Boyles Law or the Ideal Gas Law if the temperature has changed. The expansion process can
be repeated with the newer lower pressure to get even lower ones.

The spinning rotor gauge, a secondary standard, uses the rate of slowing down of a rapidly
spinning object, held suspended by a magnetic field, to measure pressure. The Pirani gauge
uses the rate of loss of heat of a hot object as a measure of pressure and there are several
gauges that use the ionisation properties of gases.

1.6.2 Manometric instruments


In this region pressures support convenient heights of liquid columns so instruments using
them are quite common. They are stable and reliable, because of their simplicity, and may be
used as primary standards so that they may never need calibration. Notice that a primary
standard need not be very accurate. A crude mercury column attached to a wooden scale, such
as used for blood pressure measurement, is a primary standard, not a very accurate one, but
one sufficient for the purpose.

The small forces these pressures exert are sufficient to usefully operate diaphragms, often
corrugated for extra sensitivity, whose movement can be magnified mechanically or measured
electrically in a number of ways. If the diaphragm encloses a space this can be evacuated and
an absolute instrument produced for air pressure measurement. The aneroid barometer and the
altimeter used in aircraft are instruments of this type.

In the normal mode of operation a piston gauge cannot work at, or below, atmospheric
pressure because there would be no pressure to support the piston. One solution to this is to
enclose the gauge in an evacuated chamber, so that the effective pressure supporting the piston
has increased by the pressure of the atmosphere. This is called the absolute mode because the
piston gauge now measures absolute pressures. An alternative solution, when measuring
below atmospheric pressure, is to invert the piston/cylinder assembly so that the atmospheric
pressure acts upwards to hold the piston from falling out. In this case, the more weights on the
piston the smaller the pressure, which in this case is measured as a negative gauge pressure.

There are various ways that mercury columns can be used to measure barometric (i.e.,
atmospheric) pressure. The most straightforward is shown in Figure 1.2, but has a
disadvantage that both surfaces of the mercury must be measured to find the distance between
them. The Fortin barometer overcomes this by squeezing more mercury into or out of the
system to bring the bottom surface up to a pointer so that it is always at the same level. Then,
only the top surface need be measured. The Kew pattern barometer uses a contracted scale so
that, even though the bottom surface drops as the column of mercury rises, the scale is rigged
so that it still reads correctly.

Barometers of course read absolute pressure, but mercury columns can read absolute or gauge,
depending on whether they are closed or open at the top. Absolute pressures from the vacuum
range to atmospheric pressure can be read as gauge pressures and in this region they are
negative, indicating that the pressure is less than atmospheric. Many gauges give the length of
mercury column supported, or they state, for example, “60 kPa vacuum’’ instead of “-60 kPa”.

10
NBIGNELL
Figure 1.6: The accuracy of various pressure standards throughout the pressure range. BUDENBERG DUAL RANGE
slope and instruments of constant relative accuracy by vertical lines. Many
instruments function under a combination of these conditions and appear on

Instruments of constant accuracy are represented by straight lines of small


the diagram as curves.

Hl0,000 MPa MANGANIN RESISTANCE GAUGE


BOURDON TUBE GAUGES
-1000 MPa
PRIMARY
-100 i%Pa

-10 MPa Ruska


-1 MPa

-100 KPa INDUSTRIAL DEADWEIGHT GAUGE

-10 KPa FORTIN BAROMETER


' HASS BAROMETER
INTERFEROMETRIC MANOMETER

-100 Pa POINT-CONTACT MICROMETER MANOMETER

-10 Pa

-1 Pa /

-100 mPa SPINNING ROTOR GAUGE

-10 mPa

1 mPa
\
-.1 mPa

-100 mPa IONISATION GAUGE

•■*0 "’P® ACCURACY

1 rnPa 0.0001% 0.0003% 0.001% 0.003% 0.01% 0.03% 0.1% 0.3% 1%


__________ I___________ I __________ I_____________I--------------- 1------------ 1-------------- 1------------- 1-------------- '----------
CHAPTER 1. OVERVIEW OF PRESSURE MEASUREMENT

1.6.3 Industrial pressure measurement

Pressures in the industrial range are capable of exerting considerable forces and are used in
hydraulic machinery for this reason. They are easily measured by the elastic distortions they
produce in devices such as the Bourdon tube, which is a metal tube with an elliptical cross­
section. When bent into a “C” shape it tends to straighten under pressure. This movement can
be magnified by a mechanism and made to work a pointer. They are the common pressure
gauges of the world with literally many millions in use. Diaphragms of a heavier sort than in
1.6.2 can also be used, usually in pressure transducers, where the movement is detected
electrically.

Both of the above types have to be calibrated, and here the dead-weight tester is almost
universally used. It only measures, or produces, gauge pressures, but in this region there is
little use for absolute pressures.

1.6.4 High pressure measurement

This is a region dominated by techniques. The technique of pressure generation may use very
large hydraulic presses or surprisingly small, elegant devices that push diamonds together.
There are special techniques for containing the pressure, for making electrical contacts, for
making optical experiments using windows and for many other purposes.

For measurement in the low-pressure end of this region very sturdy Bourdon gauges can be
used and other devices based on the elastic deformation of materials. All pressures can cause
objects to shrink a little, but here the effect becomes so large it may be used as a measurement
of the pressure. Other more subtle properties than size change too, and may be used for
measurement, for example, the electrical resistance of manganin.

Piston gauges have been used up to about 30,000 times atmospheric pressure or 3 GPa, as
primary standards. At this pressure, gases have densities greater than gold. These piston
gauges are very inconvenient to use and they have been used to measure “fixed points”. Some
materials undergo phase transitions, which are distinct, easily detected changes at well-defined
pressures. Those that have been accurately measured are called fixed points. They are then
used as reference points in pressure calibration, much as the melting of ice is used in
temperature calibration, and the high pressure piston gauge does not have to be used again
when that pressure is required to be known. Up to 8 GPa, some more fixed points have been
measured using packed pistons in cylinders in which there is considerable error due to friction,
though an average of the pressure increasing and the pressure decreasing values can be taken
in an attempt to allow for this. Beyond 8 GPa, extrapolation of properties and theoretical
predictions of behaviour are the only guides to the value of the pressure.

1.7 Conclusion
Pressure is a quantity that can be measured reasonably well over about 16 orders of magnitude,
that is, from 10'^ Pa to 10^ Pa. Pressure outside this region can be generated, but not measured
very well, for several more orders of magnitude. This monograph will confine itself to the

12
NBIGNELL

more useful ranges, covering about 12 orders of magnitude. We have seen that even over this
smaller range many different types of equipment and techniques are needed to make
measurements of pressure with useful uncertainties. These techniques will be explored further
in the following chapters.

1.8 Appendix
1.8.1 The international system of units - SI

The international system of units (le Systeme International d’unites), or SI, as it is known, had
its origins in 1789 in the French revolution. One of the many reforms undertaken then was
that of the weights and measures of France that were in a state of disorder with different units
in use in different cities. It was decided to have a system of units based on powers of ten for
their multiples and submultiples. After only ten years the first standard metre and standard
kilogram were produced. The system prospered, after an initial reluctance by the French
population to accept it, and by 1851, at the Great Exhibition in London in particular, it was
becoming clear that an international system was needed. In 1875, at Paris, the world’s oldest
operating international treaty was signed by 28 states. They agreed to set up a laboratory in
Paris (the BIPM) as a base for international standards research and as a repository for the
standards themselves.

Since then the metric system has continued to be developed, in the laboratory in Paris as well
as in many other standards laboratories around the world. A glimpse at the progress made can
be gauged from considering how the length standard has changed. The metre came into
existence as one ten millionth part of the arc of the meridian from the North Pole to the
equator, passing through Paris, though only part of the arc, from Barcelona to Dunkerque, was
surveyed. This measurement was transferred to a platinum bar and, eventually, to an improved
form of platinum-iridium alloy length standard of X-shaped cross-section. This object then
became the standard in 1889, since it was realised that each time the earth was surveyed to
better accuracy the length standard would change under the previous definition. The next
change in definition occurred in 1960 when an object-based standard was abolished by
defining the metre as a length equal to 1650 763.73 wavelengths, in vacuum, of a certain
spectral line of krypton. In 1983 the metre was re-defined as the length of the path travelled by
light, in vacuum, in 1/299 792 458 of a second.

One of the original ideals motivating the French revolutionaries and other intellectuals of the
time, such as Thomas Jefferson, was of a system of units able to be realised anywhere by
anyone and not tied to objects or governments. This has been partly achieved. Only the
standard of mass remains object-based, but an international effort is underway to replace even
this with a definition. This is possible today, but not with a sufficiently low uncertainty.
Perhaps in the next fifteen years the uncertainty will be reduced to a level where there can be
international acceptance of a new definition of the kilogram.

The international body, set up under the Metric Treaty of 1875, which decrees these new
definitions, is called the General Conference of Weights and Measures or CGPM (from the
French wording. Conference General des Poids et Mesures). In 1960, after over ten years

13
CHAPTER 1. OVERVIEW OF PRESSURE MEASUREMENT

development the CGPM brought into being the International System of Units, SI, by accepting
the base units already in use and laying down rules for prefixes, derived and supplementary
units.

The SI is a self-consistent system of units, that is, if, in some physical equation such as

P = Mg/A -I- pgh.

we use SI units on the right-hand side then we will obtain an answer in the SI unit for the
quantity on the left-hand side. This is a valuable property that does not apply if we work in
pounds, inches, pounds force per square inch, etc.

1.8.2 Units of pressure outside the SI system

Below is a list of pressure units sometimes encountered, with their values in pascals.

Name of Unit Symbol Value in Pa


Standard atmosphere atm 10L325 kPa
Bar Bar orb 100 kPa
Dynes per square centimetre or barye Dyn/cm^ 1(K) mPa
Megabarye 10^ dvn/cmT 100 kPa
Eopt of water flHiO 2.98907 kPa
Foot of water at 4 °C ftH2O (4 °C) 2.9890 kPa
Foot of water at 20 °C , / . .' flH:O pox:) or 2.98371 kPa
ftHiOat20,X
Guericke (=J cmTfO) Ger 98.0665 Pa
Gaede , S'*!. ..
1 liPa
Inch of water (std) InH2O 249.089 Pa
or inch of water gauge or in Wg
Inch of water (4 °C) inH2O(4X) 249.082 Pa
Inch of water (20 °C) inH20 (20 °C) 248.642 Pa
Kilogram force per square centimetre kgf/cm^ 98.0665 kPa
or technical atmosphere
Standard barometric inch of mercury inHg Std 3.38639 Pa
Inch of mercury (at 20 T?) inHg at 20 XI 3.37536 Pa
Kilopond force per square inch Kip/in~ or ksi 6.89476 MPa
Kilopond per square centimetre Kp/cm^ 98.0665 kPa
Pound-force per square foot Ibf/ft^ or psf 47.8803 Pa
Pound-force per square inch Ibf/in^ or psi 6.89476 kPa
Micrometre of mercury (std) umHg 133.322 mPa
Micron of mercury (std) 133.322 mPa
Millimetre of water (std) mmH2O 9.80665 Pa
Millimetre of water at 4 °C mmH2O(4°C) 9.8074 Pa

14
NBIGNELL

Millimetre of water at 20 mmH20 (20 X) 9.7890 Pa


Standard barometric millimetre of mmHg standard 133.322 Pa
mercury or mmHg (OX)
or millimetre of mercury at 0X1 or mmHg
Millimetre of mercury’ at 20 X mmHg at 20 X 132.889 Pa
Millibar mbar or mb 100 Pa
Newton per square metre N/tn 1 Pa
Poundal per square foot 1.48S16Pa
Poundal per square inch pdl/in^ 214.2957 Pa
Pound per square inch psi 6.89476 kPa
Pound per square inch gauge psig 6.89476 kPa above
atmospheric pressure
Pieze PZ 1 kPa
Ton-force per square foot tonf/ft^ 107.252 kPa (in UK)
95.7605 kPa (in USA)
Ton-force per square inch tonf/in^ 15.4443 MPa (in UK)
13.7895MPa (in VSA)
Torr torr or Torr 133.322 Pa

1.9 References
[1] Sandars, G. (2002). Traceable Measurements (Monograph 3: NML Technology Transfer
Series, CSIRO)

Further reading
Bridgeman, P.W. (1931). The Physics of High Pressure (G. Bell and Sons, also repub. Dover,
N.Y.).

Dadson, R.S., Lewis, S.L. and Peggs, G.N. (1982). The Pressure Balance Theory and Practice
(H.M. Stationery Office, London).

Eiffel, G. (1900). Travaux scientifiques executes a la Tour de 300 metres, de 1889 a 1900.
(Maretheux, Paris).

Middleton, W.E.K. (1964). The History of the Barometer (John Hopkins Press, Baltimore).

Rocke, F.A. (1984). Handbook of Units and Quantities (AAEC, Sydney).

15
CHAPTER 1. OVERVIEW OF PRESSURE MEASUREMENT

16
Chapter 2

Vacuum pressure measurement


David B Browse (1986), revised by Brad Ward

2.1 Introduction
The pressures measured in vacuum technology cover over 19 orders of magnitude, from 101
kPa to 10 *“* Pa. To measure pressure over this large range it is necessary to use a series of
gauges operating on different physical principles, depending largely on the mean free path
(mfp), A, of the gas. The mfp is defined as the average distance a molecule travels before it
encounters another molecule. The measuring range of each gauge is limited, at both ends, by
the physical phenomenon used to measure the pressure.

If the mfp of the gas is small compared to the dimensions of the vessel then the gas flow is
viscous, or continuous, and heat conduction, etc., is independent of the pressure. When the mfp
is approximately equal to the dimension of the vessel then there is a transition to molecular
flow and heat conduction depends on the pressure. At still lower pressures the heat conduction
becomes proportional to the number of molecules present, which in turn is proportional to the
pressure. At these low pressures, pumping occurs by the molecules straying into the mouth of
the pump and being prevented from re-emerging by various means.

The vacuum pressure range can be divided into the following four regions, the pressures being
approximate:

Rough vacuum (RV) or


101,000- 100 Pa
Barometric pressure
Medium vacuum (MV) 100-0.1 Pa

High vacuum (HV) 0.1 - 10-'’Pa

Ultra-high vacuum (UHV) less than IO '"’ Pa.

The reasons for division into these regions can be seen by reference to Table 2.1.

Most people involved in industry and research never have their vacuum gauges calibrated, but
rely upon the calibration provided by the manufacturer. They consider the system is under

17
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

control if the gauge reads the same today as yesterday. Bourdon tube gauges used for vacuum
measurement are covered by AS 1349 and are the exception. This chapter describes how
vacuum gauges may be calibrated, the accuracy that may be expected, and what can go wrong
with a gauge. Although the emphasis is on the medium and rough vacuum ranges high and
ultra-high vacuum techniques are described briefly.

2.2 Vacuum ranges and characteristics


As the pressure in a system is reduced below atmospheric, the flow eventually changes from
viscous to molecular. This can be characterised roughly by the Knudsen number, which is
defined as the ratio of the diameter of the pipe, D, to the mean free path, i.e., D/2,.

For D/2, >110 we have viscous flow


1 < D/2 <110, intermediate, or Knudsen, flow
D/2 < 1 molecular flow.

Figure 2.1: Relationship of some of the physical properties of vacuum, and vacuum
ranges, as a function of pressure.

2.2.1 Viscous flow

Viscous flow prevails over the rough vacuum region and into the medium vacuum region. It is
the mutual interaction of the molecules with one another that determines the character of the
flow. Although initially the flow in the pipe between a rotary pump and the vacuum system
may be turbulent, the pressure quickly reduces to the point where the Reynolds number falls

18
DB BROWSE, REVISED B WARD

below 2200 and the flow becomes laminar, so turbulent flow need not be considered in
vacuum systems.

Table 2.1 Physical properties of vacuum pressure ranges for air at 20°C.

Rough Medium High Ultra-high


Pressure Range
Torr 760-1 i-io” 10”-10‘^ < 10*2
Pa 10-10- 10--10* 10’-10'5 < 10°

No. of particles 2.5 xlO'^ 3.3x10” 3.3x10'° <3.3x10°


Per mm’ to 3.3x10” to 3.3 xlO'® to 3.3x10°
Impingement
2 lO^’-lO*® 10”-10*'' 10*-10” < 10”
rate per mm .s
continuous transition to molecular practically no
flow molecular flow flow flow; flow due
Type of flow (Knudsen flow) to single
molecules

Transport from dependent on proportional to no transport


phenomena independent pressure; mfp pressure phenomena
(heat conduction, to becoming order of vessel
viscosity) dependent on dimension
pressure
Mean free path
7x10--7x10- 7x10*2-70 70-7x10° >7x10°
2 mm
Factors vessel volume vessel volume surface monolayer time
determining Size only and shape; condition; out (1 s to several
of pump surface condition gassing hours)
becomes
important
Suitable pumps rotary rotary diffusion diffusion
roots roots turbo- getter-ion
sorption sorption molecular turbo-molecular
ejector cryo-pump
Suitable gauges bourdon Pirani Penning, hot Bayard-Alpert
manometer thermocouple cathode mass
capacitance thermistor ionisation spectrometer
Capacitance spinning- rotor
McLeod U-tube gauge
SRG

Laminar flow in a cylindrical tube normally has a parabolic velocity profile, and is called
Poiseuille flow. For Poiseuille flow the rate of flow of a fluid, i.e., the conductance C (L/s),

19
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

through a tube is proportional to the pressure difference causing the flow and to the fourth
power of the diameter of the tube.

C = 3.27x10^ D-'Pav/lri, (2.1)

where D = diameter of the tube or orifice, mm


I is the length of the tube, mm
7/ is the viscosity, Pa.s
Pav is the average pressure in the tube. Pa.

For air at 20°C:


C = 1.34 D*Pa,/l.

For a pipe evacuated with a pump of speed S,, (in L/s), the pressure drop along the pipe is
given by

P,-P2 = SplP2/0.1 yQD'Par {2.2}

For the efficient utilisation of the pump capacity it is usual to adopt the rule that the pressure
drop in the pipe must not be greater than one-fifth the pump inlet pressure.

2.2.2 Molecular flow

Molecular flow is dominant in the high and ultra-high vacuum region. In the medium vacuum
region the flow changes from viscous to molecular. It is this region that is the most important
industrially, for such processes as molecular distillation, sublimation, freeze drying of food and
pharmaceutical products, vacuum casting, etc.

For molecular flow the important parameter is the mean free path A. For air, the mean free
path in mm can be estimated by 7/P, where P is the pressure in Pa. For molecular flow the
conductance of a very long tube for air at 20°C is

C = 0.1212 DVI L/s (2.3)


with D and I in mm.

For a short tube, C = 0.1212 12/(1 -+- 4L)/3). and for circular aperture, C = 0.0016D~ L/s. For
an aperture of any shape of area A mm’, the conductance is C = 0.116A L/s.

For a vacuum system consisting of a number of pipes in parallel, the total conductance is given
by

Ct — CI + C2 + ... C„ (2.4)

And for conductances in series:

1/C, = 1/CJ + I/C2 + ... 1/C„. (2.5)

20
DB BROWSE, REVISED B WARD

This equation shows that as the smallest aperture limits the conductance, there is little use in
having a large diameter tube terminated by a small valve.

2.2.3 Thermal transpiration

Consider two vessels separated by a partition with a hole of area A. If the linear dimensions of
the hole are less than 2, then, with the same gas on each side of the partition:

P,/P2 = (2.6)

where P is the pressure and T is the absolute temperature.

The phenomenon where by a pressure gradient results from a temperature gradient is known as
‘thermal transpiration’. This may be of importance in an unequally heated vacuum system
when the pressure is in the molecular flow region or lower. For example, the part of an
apparatus cooled by liquid nitrogen will be lower in pressure by a factor (77/293)'^', or 0.51,
than those parts of the apparatus at room temperature.

The above equation is only approximate, and many researchers have investigated the effect.
Yasumoto (1980) has provided numerical values for the empirical equation:

(1 - P,/P2)/(l - = l/(AX^ + BX+ + 1) (2.7)

with X - P-JdVr) and T = (Tj + T2)/2, where Pi, P2 and T1, T2 denote pressure and temperature
in the respective parts of a two-part system connected by a narrow tube of diameter d along
which the temperature gradient exists. Selected values of the constants A, B and C are given in
Table 2.2.

Table 2.2 Values of the constants in equation (2.7) for selected gases.
Gas lO^A lO'^B C
(torr mm deg'’)'^ (torr mm deg'’)’’ (torr mm deg’’)
Nitrogen 10 7.6 40
Oxygen 10 7.0 38
Argon 9 7.0 50
Helium 1 4.0 16

If three vessels are connected in series by means of two apertures whose dimensions are again
small compared with the mean free path, then:

Pj/P2 = (T,/T2f^ and P2/P3 = ^2^3)*'^

Hence:

P1/P3 = (Ti/Ts)^^

21
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

And the pressure-temperature relationship between vessels 1 and 3 is not affected by the
presence of vessel 2, Thus, a liquid nitrogen trap between a gauge and a vacuum vessel does
not affect the significance of the gauge readings (for permanent gases). This is not the case for
mercury, the effect of which in a system is described in section 2.4,1. By extending the above
argument it can be shown that the length of tube connecting a gauge to a chamber does not in
theory affect the reading. In practice effects such as out gassing in the tube and pumping by
the gauge must be considered.

2.2.4 Gas laws

For an ideal gas, i.e., any gas at a pressure sufficiently low so that the molecules can be
considered to have zero volume and zero attractive force between them,

P,V,/T, = P2V2/T2. (2.9)

The equation of state is given by

PV = nRT (2.10)

where n is the number of moles of gas and R is the gas constant = 8,314 J/K.mol.

Real gases at higher pressure can be better described by Van der Waals equation:

(P + A/V-)(V-b) = RT. (2.11)

The values of A and b for some common gases are given in Table 2.3.

Table 2.3. Vales of A and b of Van der Waals equation for some common gases.
A b
(cm^/moO^atm cm^/mole

Nitrogen 1.390x10^ 39.13


Oxygen 1.360x10^ 31.83
Air 1.33 X 10^ 36.60
Argon 1.345 x10^ 32.19
Helium 3.412x10^ 23.70
Mercury 8.000x10^ 17.00

The equation of state for a real gas may be expanded in terms of virial coefficients in powers
of density 1/V.

PV = RT{\ +B/V+C/V" + -- - ) (2.12)

22
DB BROWSE, REVISED B WARD

For nitrogen at 20°C, B = -7.4 cm^/mol and C = 1400 cm^/mol. For those situations in vacuum
applications where real gas effects need to be taken into account then either Van der Waals or
the virial equation may be used. Except for very precise calculations most gases may be
considered ideal for pressures up to two or three atmospheres.

2.3 Vacuum pumps


Different types of vacuum pump are used to obtain different pressures in a system. Some
pumps will operate from atmospheric pressure down to their ultimate vacuum, others require
another pump to reduce the pressure to a predetermined level before they will operate. Some
pumps require a second pump to maintain a continuous low pressure at the output, this is
called “backing”. The more commonly used vacuum pumps are described in this section and
their typical pressure ranges are illustrated in Figure 2.2.

Diaphragm
Rotary
Sorption
Roots
Molecular Drag
Diffusion
Turbo Molec.
Ti Sublimation
Ion
Cryogenic

IO-* 103 102 101 1 101 102 IOS 104 105 106 10? 106 10’6 lo-i®

Pressure (Pa)
Figure 2.2: Pressure ranges of a number of vacuum pumps.

2.3.1 Rotary pump

This is the most commonly used pump for evacuating systems down to pressures of between 1
and 10 Pa. The pump, which operates by compressing the gas and expelling it into the
atmosphere, is shown schematically in Figure 2.3. Most modern pumps have a gas-ballast
system to expel condensable vapours. This simply bleeds a small amount of air into the pump
through a one-way valve, the operation of which can be seen in Figure 2.3. Typically, a two-
stage pump with full gas-ballast will attain 1 Pa, but this is reduced to about 0.05 Pa when the
gas-ballast valve is closed. All rotary pumps produce a certain amount of oil back-streaming
(i.e., oil that migrates from the pump back into the system), so for an oil-free environment
some form of oil trap (section 2.3.3) must be used.

2.3.2 Roots pump

A Roots pump is a positive displacement pump with two symmetrical rotors rotating in
opposite directions within the pump housing (Figure 2.4). The rotors are synchronised by a

23
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

gear drive so that they move past one another and with close proximity to the casing walls
without touching. This enables the pump to operate at high rotational speed with little friction.

Figure 2.3: Schematic diagram of a rotary pump with gas ballast.

INLET

CASING
LOBE

DISCHARGE
Figure 2.4: Diagram of a Roots pump.

Such pumps require backing by another pump to below 10 kPa and obtain an ultimate pressure
of about 10'" Pa.

2.3.3 Sorption pump and vapour trap

A sorption pump and vapour trap consists of a metal enclosure containing an activated sorbent
that has a large surface area and a pore size similar to the molecular diameter of the gas being
absorbed. Such properties are exhibited by activated charcoal, activated alumina and synthetic
molecular sieves. Molecular sieves, of zeolites (alkali alumino-silicates), the most commonly
used material, have a structure that is made of very-fine roughly-spherical cavities, connected
by minute channels of about the same diameter as the gas molecules. As a consequence they
have a large surface area of about 10'’ m"/kg. The pore structures would normally be occupied

24
DB BROWSE, REVISED B WARD

by water of crystallisation. Gas molecules are bound to sorbents by physical attraction


(physisorption) or by chemical interaction (chemisorption). Trapping is more easily achieved
when gas molecular energies are low enough to allow a relatively long residence time when
they contact the sorbent surface. Hence, sorbents are usually cooled with liquid nitrogen,
increasing their sorptive capacity by up to 10^ times its value at room temperature.

Sorption pumps operate from atmospheric pressure down to about 0.1 Pa, depending upon the
gas and the type of zeolite. Because of the small pore diameters of types 3A and 4A (0.3-0.4
nm), sorption pups are used for general pumping, while type 13X (pore diameter 1.3 nm) is
used for trapping, rather than pumping, large organic molecules in oil vapour traps. In contrast
to pumps, traps are normally operated at room temperature.

Sorption pumps can only be used until they are saturated, when they must be warmed to room
temperature to reactivate the sorption material. So for continuous pumping of large quantities
of gas they must be used in tandem. They have the advantage of providing clean, vibration-
free pumping with no moving parts and requiring virtually zero maintenance. The main
drawback is that they require a supply of liquid nitrogen.

2.3.4 Diffusion pump

In a diffusion or vapour pump (Figure 2.5) a stream of the pump fluid vapour, produced by a
heater, travels up the central tube and emerges from a nozzle, expands into the pump casing
and is condensed on the cooled walls. The vapour streaming towards the pump discharge
outlet, maintained at a pressure of 50 Pa or less by another pump, usually rotary, imparts
momentum in this direction to the gas molecules, so creating a pumping action between the
intake aperture at high vacuum and the discharge outlet at backing pressure. By suitable
design of the jet, the vapour stream emerging from the nozzle becomes supersonic producing a
shockwave in which the gas is rapidly compressed. The shock wave acts like a dam, providing
the seal required across the pump.

Figure 2.5: Schematic diagram of the operation of a diffusion pump.

25
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

To reach the ultimate pressure attainable with the diffusion pump, and to prevent back
streaming of the pump fluid into the evacuated chamber, it is essential to have a chilled baffle
or liquid nitrogen trap between the pump and the chamber. Pressures of 10'^ Pa can be
achieved with some oil diffusion pumps and 10'^ Pa with mercury diffusion pumps.

2.3.5 Ion pump

Sputter ion pumps have a pumping action based on sorption processes that are initiated by
ionised gas particles (molecules that have lost one or more electrons). The ions, produced by a
high-voltage discharge in a magnetic field, impinge upon the titanium cathode of the electrode
system and sputter the cathode material (i.e., knock out the molecules). The titanium
deposited on the walls and elsewhere acts as a getter film and absorbs gases such as nitrogen,
oxygen and hydrogen. Pumping also occurs by the embedding of gas ions in the cathode.
Such pumps operate from about 1 Pa to 10’’ Pa. As for a Penning gauge (section 2.5.1), the
current in the pump can be used to measure the pressure.

2.3.6 Turbo-molecular pump

The turbo-molecular pump is basically an axial flow compressor (a jet engine!), which consists
of a stator and a rotor that rotates at speeds greater than 30,000 r.p.m. The principle, which has
been well known since 1913, depends on the fact that the gas particles to be pumped receive,
through impact with the rapidly spinning rotor, an impulse in the required flow direction. The
compression ratio is the ratio of the outlet to the inlet pressure, so the lower the pressure
produced by the backing pump the lower the ultimate pressure. The pumping speed is limited
by the compression ratio (around 10^ for nitrogen), which varies with pump design and gas.
Turbo molecular pumps can produce an ultimate pressure of around 10'^ Pa.

2.3.7 Cryo-pump

A cryo-pump consists of a surface that is cooled below a temperature of 77 K so that gases and
vapours condense on the surface. In general, liquid helium is used, with often a built-in
compressor to produce the liquid. Pressures of 10'^ Pa can be reached with such pumps.

2.4 Pressure measurement - rough and medium-vacuum


The rough- and medium-vacuum ranges are considered together because the methods overlap
to a considerable extent. The types of gauge used can be separated into two groups -
mechanical and thermal conductivity. Above 100 Pa the standard instruments are almost
exclusively manometers and piston gauges.

2.4.1 Mechanical gauges

Liquid Manometers

There are many forms of liquid column micromanometers, but whatever the form they all
balance the test pressure with a column of fluid of known density. The variation in form

26
DB BROWSE, REVISED B WARD

occurs with the method of height measurement, selection of manometer fluid, treatment of
capillarity, and provision for vapour pressure effects. The majority of these manometers
operate in both the absolute and gauge mode and are covered in detail in Chapter 3.

McLeod gauge or compression manometer

An alternative measurement approach is the compression manometer developed by McLeod in


1874. With this instrument a gas is compressed to a pressure that can be measured by
differential column heights according to standard manometric practice. The structure of a
McLeod gauge is depicted in Figure 2.6.

With the vacuum system at the pressure, P, to be measured, gas at atmospheric pressure is
admitted to the reservoir and drives the mercury up the central tube. When the mercury reaches
the cut-off level a sample of gas from the vacuum system at pressure P is trapped within
volume V. Continued rise of the mercury compresses the gas into the compression capillary so
that the final pressure is measured by the differential height h between the compression
capillary and the comparison capillary.

To Test Chamber

Comparison capillary
Compression capillary
cross section a

Scale

Compression Volume v

Cut off level

Pressurising gas

Mercury reservoir

Figure 2.6: Schematic diagram of a McLeod Gauge

If a is the area of cross section of the capillary, then assuming equal temperature in both
capillaries the gas law gives:

27
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

PV=(P + h)ah (2.13)

where P is assumed to be in mmHg. If P is in Pa then equation (2.13) is written:

py = (P + pgh)ah.

and from (2.13): P = ah^/(y- ha). (2.14)

A linear relationship between P and h can be obtained if the compression ratio is made the
same for all pressures, by raising the mercury in the compression capillary to the same level.
Then:

py = h^(P + h)

and P = h,>ah/(y- h„a), (2.15)

where ho is the constant length of trapped gas in the compression capillary.

Because in the first technique the compression ratio, and therefore the pressure amplification,
is inversely proportional to the square root of the pressure, this technique has a higher
sensitivity at low pressure and a larger useful working scale.

A more accurate relationship between pressure and volume is given by Van der Waals
equation (2.11). The error, AP in using the gas equation rather than Van der Waals is:

AP/P, = Pf/RT(A/RT -b) (2.16)

where pf is the final pressure in the compression capillary.

The volume V and the area of the capillary a are usually determined prior to the final assembly
of the gauge. For the highest accuracy instruments the distortion due to the sealing of the
capillary must be determined.

The differential height is usually measured by a scale attached to the capillaries. For more
precise work the heights can be measured with a cathetometer or travelling telescope. For best
viewing the mercury surfaces are back lit through a translucent screen.

The main sources of error are capillarity and non-uniform capillaries, particularly near the top
where the compression capillary is sealed. As the effect of capillarity due to surface tension is
very large for small diameter tubes, then it is possible for the variation due to capillarity to be
compounded by the use of a comparison capillary, because both capillaries could have
significant non-cancelling capillarity errors. Use of a larger bore for the reference arm reduces
the error provided the theoretical depression for the two bores is known.

The movement of the mercury column in the capillary tube can cause an electrical charging of
the tube, producing erratic behaviour of the mercury level, particularly on the downward
motion. The final position should be approached smoothly while tapping the capillary tube.

28
DB BROWSE, REVISED B WARD

As for manometers, it is normal to use a cold trap with a McLeod gauge to keep mercury out
of the rest of the system. These instruments are equally affected by the mercury vapour
pumping effect (see below).

Ruthberg [1] estimates that for a compression volume of 2 L the uncertainty at 0.1 Pa is 0.8%,
with a bias of 1 % to be added for the mercury vapour pumping effect.

A McLeod gauge cannot perform a continuous measurement of pressure. Even after the gauge
has been read and the mercury lowered, it is necessary to wait some time for equilibrium to be
established. Also, the gauge does not measure the pressure due to condensable vapours in the
gas sample.

The McLeod gauge is more useful as a calibrating instrument for other gauges than for actually
reading pressure in a vacuum system. Even then, because of the errors discussed above, great
care must be taken if a commercial instrument is used at pressures below about 0.05 Pa.

When using a McLeod gauge with a sealed system, raising the mercury will change the
pressure in the system. This can cause a substantial error but is overcome by proper technique.
The mercury should be raised to just below the cut-off level and the system allowed to
equilibrate. The small additional movement to achieve cut off then causes a negligible change
in volume.

Numerous modifications to the McLeod gauge have been developed, but most of the variations
in common use relate to how the mercury is raised, e.g., a swivel action, while the basic form
of the gauge remains unchanged.

Mercury vapour pressure


Mercury, from instruments such as McLeod gauges and manometers, will contaminate a
system unless steps are taken to prevent it. The mercury vapour can be excluded from the test
chamber either by use of a differential pressure gauge or by use of a cold trap. The first
method requires an auxiliary pressure control on the instrument side for precise tracking of the
test pressure. The use of a cold trap causes a mercury vapour flow to the trap and the gaseous
molecules inside the instrument are dragged away. Therefore, the gas pressure is lower at the
instrument than in the test chamber.

The effect of mercury vapour pumping on manometer readings is still controversial. The
results calculated depend upon the range of pressures and temperatures and upon the geometry
of the system. Berman [2] pointed out that for pressures below 0.13 Pa (0.001 mmHg) all the
mathematical models agree and he obtained good agreement between calculated and
experimental values. His graph plotting the percentage error of the difference between the
pressure at the manometer and the pressure at the cold trap as a function of the manometer
pressure for various temperatures and tube radii, is given in Figure 2.7. For a tube of radius 5
mm and 293 K the error is about 9%. Various methods, such as cooling the mercury and
connecting the cold trap and manometer with a capillary, have been proposed to compensate
for the pressure offset. None of these is very convenient to apply for routine measurement.

29
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

Figure 2.7: Mercury vapour pumping effect of a manometer or Mcleod gauge. The
percentage difference between the pressure of the instrument and that at the
cold trap, Ap, relative to the instrument pressure (P), is plotted as a function of
temperature and the inner radii of connecting tubing (r), after Berman [2].

Capacitance diaphragm gauges

These gauges consist of a thin highly pre-stressed metal diaphragm positioned between fixed
capacitor plates. The diaphragm separates two gas tight enclosures that are connected to the
external pressure ports. A change of pressure within the enclosure produces a force that
deflects the diaphragm, and this varies the relative capacitance of the diaphragm and the fixed
capacitor plates. The variable capacitance forms part of a capacitance bridge, hence, an output
may be measured by rebalancing the bridge, or the unbalance may be used to develop an
electrostatic potential to restore the diaphragm to its null position. The former technique is the
one normally adopted in commercial instruments. A discussion of the operation of these
instruments is given in Chapter 6.

Capacitive diaphragm gauges may be either differential or absolute instruments. In the


absolute case the reference side is sealed at a very low pressure and contains a getter to cope
with any outgassing. Differential instruments may operate in either absolute or gauge mode
and. so, along with manometers, are the only vacuum measuring instruments capable of such
operation. A change in line pressure can change the zero, due to strain-induced changes in the
sensor’s geometry. Some recent models incorporate a secondary “guard volume” that
surrounds the entire side, thus minimising the effect of line pressure changes.

To minimise the effects of temperature change the instruments are usually temperature
controlled at around 45 °C. For accurate measurements, at pressures below 100 Pa,
corrections for thermal transpiration effects may have to be made. See section 2.2.3 and [2].

30
DB BROWSE, REVISED B WARD

Modem instruments can cover 5 decades of pressure and the most sensitive heads have a
resolution of 10’^ Pa corresponding to a diaphragm deflection of around 0.5 nm. Depending
upon the pressure, the claimed accuracies of these gauges range from 0.1% to over 5% of the
reading. Hyland and Tilford [4] measured an average long-term instability of 0.4%, in
agreement with other workers, although, shifts of up to 2% were obtained.

Piston gauges

The theory and operation of piston gauges is described in Chapter 4. While gas operated
gauges are mainly used for gauge pressure measurement, a few instruments are equipped with
a bell-jar that may be placed over the piston and weights, so that measurements may be made
relative to vacuum. The lowest pressure that may be reached by such instruments is limited by
the mass of the weight carrier and the piston area, usually around 1.5 to 2 kPa abs. Due to this
pressure limitation, piston gauges are standards that can be used for measurement only in the
rough vacuum or barometric pressure ranges, and above.

Because of gas leakage past the piston and cylinder a pumping system with a large throughput
is required to reduce the pressure in the bell-jar to 0.1 Pa. Any error in the measurement of
this pressure is then negligible in relation to the pressure generated by the piston gauge.

By connecting two instruments to the opposite sides of a differential pressure gauge (e.g., a
capacitance gauge), it is possible to generate pressure differences down to 0.1 Pa, although,
this is still at a line pressure of 2 kPa or larger.

2.4.2 Thermal conductivity gauges

The heat loss from a resistance element, or filament, due to the conduction of the gas, as a
function of pressure, forms the basis of these instruments.

For all thermal conductivity gauges, the calibration depends on the state of the filament,
oxidation, corrosion and deposited films from oil or grease vapours. The calibration is a
function of the type of gas in the system and is usually given in terms of nitrogen.

Pi rani gauge

In 1906, Pirani used the rate of loss of heat from a thin wire suspended in a vacuum chamber
as a measure of the pressure. The heat transfer due to convection in a rough vacuum is almost
independent of pressure until it drops to 300 or 400 Pa. At this pressure the mean free path is
approximately equal to the filament diameter and the heat transfer becomes dependent upon
pressure. This continues until the mean free path becomes comparable with the inner diameter
of the gauge head. This occurs at about 0.1 Pa when the heat transfer through the gas by
conduction has almost ceased and so again becomes independent of pressure.

The filament forms one arm of a Wheatstone bridge network, which serves both to heat the
wire and to measure its resistance. A schematic diagram of a Pirani gauge is shown in Figure
2.8. To minimise temperature effects, some gauges have four filaments, matched for

31
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

resistance, each mounted in a separate glass envelope and connected one in each arm of the
bridge. To prevent filament sag, Pirani gauges should always be mounted vertically.

Figure 2.8: Diagram of a Pirani gauge and associated circuitry

Pirani gauges can be operated in two modes:


Constant current: With increasing gas pressure, the temperature of the filament decreases
because of the larger heat transfer so that the bridge becomes out of balance. The bridge out-
of-balance signal is a measure of the pressure and may be used over the range of 0.1 Pa to
about 200 Pa.

Constant Temperature: The applied voltage is adjusted so that the filament resistance, and
hence its temperature, remains constant, and so the bridge always remains balanced. The
output is then the current needed to maintain the temperature. The method can be used to
measure pressure over the range 0.1 Pa to 100 kPa. Because of their very short response times,
Pirani gauges are particularly suitable for control and monitoring of pressure and for use as
coarse leak detectors.

Poulter et al [5] obtained, for both types of gauges, a reproducibility of about 6% of reading
under well defined, clean conditions. In particular, the gauge should be calibrated every six
months and the zero and atmospheric pressure points set each day the gauge is used. They
concluded that when these gauges are used under well defined conditions, they may be used as
reference gauges for the routine checking of other gauges of this type used on industrial
vacuum systems.

Thermocouple gauge

The filament of a thermocouple gauge is heated electrically and its temperature is measured by
means of a thermocouple spot welded to it. As the pressure increases, the temperature of the
filament decreases and the reading of the thermocouple is a measure of the pressure.
Thermistor gauge

A thermistor gauge is a Pirani-type gauge that uses a thermistor as the resistance element.
Thermistors have a high negative temperature coefficient of resistance. The curve of the bridge
current needed for a balance as a function of pressure is mainly linear from 0.1 to 100 Pa when
plotted on a log-log scale.

32
DB BROWSE, REVISED B WARD

2.S Pressure measurement - high vacuum region


2.5.1 Gauge types

Cold-cathode ionisation (Penning) gauge

Penning gauges contain two unheated electrodes, a cathode and an anode, between which the
discharge is excited by means of a de voltage of about 2 kV. A magnetic field makes the paths
of the electrons so long that their collision probability with gas molecules and the production
of charge carriers is large enough to maintain the discharge to low pressure. The reading of
these gauges is dependent on the nature of the gas. The gauges also pump gases similarly to a
sputter-ion pump. The upper limit of the measuring range is a few Pascals due to the onset of
the flow discharge. The lower limit is about 10'^ to 10 '^ Pa, due to difficulty in maintaining the
discharge and the presence of another emission current that is indistinguishable from an ion
current.

The main problem with these gauges in use is the gradual formation of a non-conducting film
on the electrodes, which either stops the discharge or causes erratic readings. The electrodes
then have to be cleaned or replaced. The presence of a permanent magnet prevents high-
temperature bake-out and eliminates out gassing.

Hot-cathode ionisation gauges

Hot-cathode ionisation gauges are the main gauges for measuring pressure in the HV and the
UHV regions. Their calibrations change with age, use and type of gas measured. The stability
can range from 2 to 5%. Methods of calibration are described in section 2.5.3.

Figure 2.9: Hot-cathode ionisation gauges (schematic): (a) conventional gauge, (b)
Bayard-Alpert gauge, (c) modulated Bayard-Alpert gauge, (d) diagram of a
typical Bayard-Alpert gauge. A - anode, F - fllament, G - grid, IC - ion
collector, M -modulator electrode.

The gauges come in a number of different forms with differing numbers of electrodes (Figure
2.9). The basic type consists of a hot cathode, an anode and an ion collector. The gauges work

33
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

with low voltages and without an external magnetic field. The electrons from the hot cathode
are accelerated in the electric field and receive sufficient energy to ionise the gas in which the
electrode system is located. The positive gas ions formed are transported to the ion collector,
which is negative with respect to the cathode, and give up their charge. The ion current
thereby produced is a measure of the gas density and so of the gas pressure. Bayard-Alpert
gauges have electrode arrangements to reduce some of the disturbing effects, such as X-rays,
which influence low pressure measurement. The reading of ionisation gauges is a function of
the nature of the gas in the system, which must be analysed for accurate measurements. The
useful pressure range is approx. 10'^ to 10 " Pa.

Spinning-rotor gauge (SRG)

The essential part of the SRG is a steel ball (typically 4.5 mm diameter) that is levitated using
permanent magnets and electromagnets. Inductive sensing and electronic feedback provide
suspension stability along the three orthogonal axes. The basic parts of the SRG head are
illustrated in Figure 2.10. The ball is contained within a non-magnetic tube attached to the
vacuum system by a UHV flange. A high frequency two-phase inductive drive spins the ball
to approximately 400 Hz. The rotation of the ball is sensed by inductive pickup of the rotating
component of the ball’s magnetic moment, as it describes a small cone about the vertical. The
pickup signal is amplified and the rotational period of the ball is timed.

Figure 2.10: Operation of the head of a spinning-rotor viscosity gauge. The head,
without the magnet and coils, may be removed to enable the tube to be
baked at high temperatures.

It has been shown experimentally that the gas molecules leaving a smooth adsorbate-covered
metal surface do not, on the average, exchange tangential momentum with the surface.
Accordingly, in the molecular flow region, the gas friction on a rotating ball may be calculated
from classical mechanics and molecular dynamics by considering the tangential momentum
exerted by the incident molecules. Hence the relative rotor deceleration per unit time is

34
DB BROWSE, REVISED B WARD

- 0)/CU= IOP/ttR pc,

where c = mean molecular velocity,


p = density of the ball,
P = pressure,
R = radius of the ball,
co = angular velocity of the rotating ball,
and C5 its time derivative .

In theory, the SRG is an absolute gauge that does not need calibration, as all the quantities in
equation (2.17) can be obtained by measurement. However, the above considerations are
based on the assumption that the tangential momentum transfer is complete, whereas this is not
always the case, and the accuracy is affected by the following factors [6]:

• coverage of the rotor surface with non-metallic material


(contamination with oil or grease);
• absence of adsorbates; and
• roughening of the rotor surface.

The effect of these conditions is accounted for by a coefficient cr on the right-hand side of
equation (2.17). For a rotating ball under ideal conditions and a perfect sphere, c7= 1.0. For
an isotropically rough ball of maximum surface roughness, cr= 1.27.

There is, however, another effect that must be taken into account in assessing the accuracy of
the SRG and that is a pressure independent residual drag. The main causes of this effect are
eddy currents induced in the spinning ball by asymmetries in the suspension field and eddy
currents induced in surrounding metallic components, particularly the tube, by the rotating
components of the ball’s magnetic moment. The residual drag is determined by applying a
pressure of less than 10’^ Pa to the SRG and observing the indicated pressure. This is then
entered as an offset correction.

The useful lower pressure of the SRG (approximately 10'^ Pa) is set by the zero stability,
thermal disturbances and changes in the residual drag. McCulloh et al [7] list the optimum
conditions for obtaining a pressure stability within ±10'^ Pa.

Adequate time (5 h after a complete spin up from zero) must be allowed for thermal stability
after the inductive drive mechanism is used. The residual drag must be redetermined after
each suspension of the ball and should be redetermined just before a set of low pressure
observations.

Laboratory temperature changes should be minimised, as should vibration levels. The


apparatus should not be touched or disturbed during measurements, mechanical fore pumps
should be isolated, etc. The external magnetic environment should not be changed. And long
sampling or pre-scale periods must be used to minimise the random noise contribution,
particularly at the lowest pressures.

35
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

The main application of the SRG is as a secondary standard with uncertainties of 1%. The
range of recent instruments extends well into the medium vacuum range and so they could be
used to replace the McLeod gauge. However, given the requirements for HV operation
(particularly cleanliness) the instrument is only likely to be used for calibrating other transfer
standards in the medium vacuum range.

2.5.2 Partial pressure measurement

Because most vacuum gauges have differing sensitivities for different gases, to adequately
calibrate such gauges, especially in the HV and UHV regions, it is important to know the
composition and partial pressure of the gas in the chamber.

The partial pressure of a specific kind of gas in a gas mixture within a system is the pressure
that that gas exerts on the container walls. The sum of the partial pressures of all kinds of gas
gives the total pressure. These gases are essentially distinguished from one another by their
molecular masses. Partial pressures are determined qualitatively by measuring the molecular
masses of the gas particles (i.e., what gases are present) and quantitatively by the abundance of
the individual mass numbers.

The best method of measuring partial pressure and composition is to use a mass spectrometer.
This consists of an ion source, an ion separation unit and an ion collector. The ion separation
unit is the heart of the technique and most gauges use a quadrupole system with four electrodes
of hyperbolic cross-section. A constant high frequency and superimposed de voltage are
applied to the electrodes. Only molecules of a particular mass may reach the collector for
particular values of frequency and voltage. By scanning the frequency and voltage a spectrum
of the molecules present may be built up.

2.5.3 Primary calibration techniques

Static expansion system

The basis of the procedure is to fill a small pre-evacuated volume with a known amount of gas
and then to expand it into a larger volume, which has also been evacuated and sealed. From a
knowledge of the volume ratio and the initial pressure, the final pressure can be calculated.
The procedure can be repeated in a multistage process. Errors are produced by out gassing,
gauge pumping, adsorption and temperature effects. Pressures down to 10'^ Pa can be
generated with an uncertainty of 1 or 2%, which drops to less than 0.2% at higher pressure.

Dynamic expansion or orifice flow method

In this method a steady calculable pressure is generated by balancing the rate of flow of gas
into a chamber against the rate of removal of gas. The most popular method uses a thin circular
orifice at the junction of two spheres. The upper sphere is the calibration chamber while the
lower one ensures a proper distribution of molecular velocities on the low-pressure side of the
orifice. To calculate the conductance of the orifice it has to be thin and have a diameter less
than the mean free path. These restrictions have limited the maximum pressure to 50 mPa.

36
DB BROWSE, REVISED B WARD

For all orifice-based systems it is necessary to have some form of flowmeter to accurately
measure the very small throughput.

It is very important in this method to ensure that the velocity of the gas molecules is
randomised and that the incoming molecules cannot stream directly into the orifice or any
gauge. This can be done by introducing the gas into the system via a tube positioned so that all
the incoming molecules must make at least one collision with the system walls before moving
towards a gauge. The principle is most important for all calibration systems of any form.

The best systems can generate pressures with an uncertainty of around ±1%. Of the two
calibration methods described above, the expansion method probably has the greater potential
for improved accuracy. The dynamic method has an advantage of being suitable for any gas
while the series expansion method can only be used with the preferred gases (nitrogen and the
inert gases). The static expansion method has no upper pressure limit and a practical lower
limit of 1 jiPa; the dynamic method has an upper limit of 50 mPa and is usable down to the
lowest obtainable pressures. The dynamic method is also less susceptible to contamination of
the system.

2.5.4 Comparison calibration techniques

The methods described in section 2.5.3 require very expensive and specialised equipment and
are performed throughout the world by only a few standards laboratories. An acceptable
method, providing sufficient accuracy for most requirements, is to use a gauge calibrated on a
primary system as a standard. The basic principles for the calibration of vacuum gauges by
comparison techniques are not limited to any particular pressure range.

• The test chamber should have a volume at least 20 times the total volume of the gauges and
any associated plumbing connected to it.
• Because gauges may show either sorption or desorption phenomena, the tubing connecting
the gauges to the chamber should have a conductance of at least 100 times the volume rate
of the flow of these effects. This usually results in a conductance of at least several
litres/second.
• The base pressure of the system should not exceed 2% of the lower limit of the calibration
range, as indicated on the reference gauge.
• During calibration the valve between the pump and the chamber may be either partially
closed or fully closed. In the former case the gas inlet must be adjusted to give a constant
reading on the reference gauge.
• The geometry should be such that a molecule emanating from the measuring chamber
cannot enter the gauge without having encountered a wall at least once.
• There should be at least three calibration points per decade of the range.
• Pressure and temperature gradients should be minimised so as not to introduce any
appreciable errors.

37
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

2.6 Pressure measurement - ultra-high vacuum


Calibration in the ultra-high vacuum region is very specialised, with the main technique being
the molecular beam method [8]. The arrangement, shown schematically in Figure 2.11,
consists of a cell and a calibration chamber. Circular orifices connect the two chambers via an
extremely low-pressure region. In the reference pressure chamber a constant gas pressure
measured by a calibrated gauge, is maintained. The test gas effusing through the orifice, area
Al, of the reference cell, strikes the orifice of area A2 of the calibration chamber at a distance Z.
The vacuum gauge to be calibrated is mounted in the calibration chamber.

The calibration pressure in a stationary and thermal equilibrium condition is:

P = const. P,A j/i^ (2.18)

where I » (4Ai/7rf\

Using this method the Physikalisch-Technische Bundesanstalt of Germany has been able to
extend the routine calibration of vacuum gauges down to 10'*® Pa.

REFERENCE CHAMBER j- CRYO PANEL CALIBRATION


CALI CHAMBER
GAS INLET4 ______

REFERENCE GAUGE
TEST GAUGE

Figure 2.11: Schematic diagram of the molecular beam method for the calibration
of ultra-high vacuum gauges.

38
DB BROWSE, REVISED B WARD

2.7 References
There are a large number of books on vacuum fundamentals and techniques, all of which
broadly cover much the same material. A number of these books are included for general
reference. Where detailed information is not in the test books references have been given to
some original papers.

[1] Ruthberg, S. (1975). Experimental Thermodynamics II: Experimental Thermodynamics


of Non-reacting Fluids. Ch. 4. Part 6. Edited by B. Le Neindre and B. Vodar.
Butterworths (London).

[2] Berman, A. (1974). Vacuum, 24 (6), 241.

[3] Poulter, K.F., Rodgers, J-J., Nash, P.J. Thompson, T.J. and Perkin, M.P. (1983). Vacuum,
33 (6), 311.

[4] Hyland, R.W. and Tilford, C.R. (1985). J. Vac. Sci. Technol. A3(3), 1731.

[5] Poulter, K.F., Rodgers, M-J. and Ashcroft, K.W. (1980). J. Vac. Sci. Technol., 17(2),
638.

[6] Fremery, J.K. (1985). J. Vac. Sci. Technol. A3(3), 1715.

[7] McCulloh, K.E., Wood, S.D. and Tilford, C.R. (1985). J. Vav. Sci. Technol. A3(3), 1738.

[8] Grosse, Von G. and Messer, G. (1981). Vakuum-Technik. 30, Jahrgang, Heft 8, 226.

Further reading

Chambers, A., Fitch, R.K. and Halliday, B.S. (1998). Basic Vacuum Technology. lOP
Publishing.

Diels, K. and Jaeckel, R. (1966). Leybold Vacuum Handbook. Pergamon Press (Oxford).

Ernsberger, F.M. and Pitman, H.W. (1955). Rev. Sci. Instrum. 26, 584.

Leek, J.H. (1957). Pressure Measurement in Vacuum Systems. Chapman & Hall Ltd.
(London).

Leybold-Heraeus GMBH. Vacuum Technology its Foundations Formulae and Tables.

Poulter, K.F. (1977). J. Phys. E: Scientific Instruments, 10, 112.

Roth, A. (1976). Vacuum Technology. North Holland (Amsterdam).

39
CHAPTER 2. VACUUM PRESSURE MEASUREMENT

Turnbull, A.H., Barton, R.S. and Riviere, J.C. (1962). An Introduction to Vacuum Technique,.
George Newnes Limited (London).

Van Atta, C.M. (1965). Vacuum Science and Engineering. McGraw-Hill (New York).

Weissler, G.L. and Carlson, R.W. (1979). Vacuum Physics and Technology. Vol. 14 of
“Methods of Experimental Physics”. Academic Press (New York).

Yarwood, J. (1967). High Vacuum Technique. Chapman & Hall Ltd (London).

Yasumoto, 1. (1980). J. Phys. Chem., 84, 589.

40
Chapter 3

Barometry, manometry and liquid columns


David B Browse (1986), revised by Kitty Fen

3.1 Introduction
Traditionally, manometers (from mano-, thin, rare + meter, measure) measure differential
pressures or the pressures of gases or vapours. If they are relative to atmosphere, we say that
they measure gauge pressure. Barometers (from baro-, weight + meter, measure) measure the
pressure of the atmosphere and it is relative to zero pressure. Therefore, they measure absolute
pressure. In both types of instrument the pressure is balanced against a column of liquid in a
tube or against the elastic force of some element as in an aneroid barometer or pressure
transducer. The column of liquid can be water, mercury or other liquids of suitable properties.
Figure 3.1 shows the general concept of a liquid-column manometer.

Pi-P2 = pgh
(3.1)

where

p - density of liquid inside the tube


g - acceleration due to gravity
h - vertical height between the two columns

Figure 3.1: A U-tube manometer

If P2 = 0 and Pi is any pressure, the manometer is one that measures absolute pressure.
If P2 = 0 and Pi is open to the atmosphere, it is a barometer. A barometer is a particular case of
a manometer.
If P2 = atmospheric pressure and Pi is any pressure, then the manometer is one that measures
gauge pressure.
If Pi and P2 are connected to any pressure other than zero pressure, it is a manometer that
measures differential pressure.

41
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

Mercury barometers and manometers are now mainly used as primary standards to calibrate
pressure-measuring units such as transducers. Whilst there are a few long range liquid
columns, confined to research laboratories, only columns of length 2 m or less are described
here, and the measurement of pressure with such columns over the pressure range 100 Pa to
about 120 kPa, either gauge or absolute, will be discussed.

A brief description will be given of aneroid barometers and their use, but the large number of
so-called electronic barometers and manometers will not be discussed. These are basically
sensitive, high accuracy pressure transducers, or an aneroid barometer to which an electronic
transducer has been fitted, and as such are described in Chapter 6. Their use in barometric
pressure measurement is identical to aneroid barometers.

Manometers, particularly micromanometers, are described briefly. Whilst their use in industry
is still important for the measurement of airflow, their main use is as calibrating instruments
for the electronic instruments that have displaced them.

3.2 Mercury barometers


3.2.1 History

In 1643, Torricelli demonstrated that the atmosphere could support a column of mercury about
760 mm high. The subsequent development of the mercury barometer in recent times has been
unfortunate in the number of scales that have come into use and in the variety of gravity values
that have been associated with the scales. The use of three standard temperatures (0 °C, 12 °C
and 62 °F) with three variants of standard gravity (980.665 cm/s", 980.62 cm/s“ and gravity at
mean sea level in latitude 45°) combined with the choice of five scales (millibar, millimetre,
inch, altitude in metres and altitude in feet), has provided a fruitful source of misconception
and error. The situation has been further complicated by the differing temperature coefficients
of barometers of the Kew and Fortin type. The correlation and correction of readings on the
various scales of these two types of barometers have been the subject of confusion.

As a consequence of discussions between the National Physical Laboratory (UK) and related
national and international bodies, B.S. 2520 “Barometer conventions and Tables” was
published in 1954 and revised in 1983 [1]. This standard defined pressure units, standard
conditions and the corrections to standard conditions.

3.2.2 Standard conditions

For a given pressure or pressure difference, the height of a mercury column is a function of the
temperature and the value of gravity. Therefore, to enable columns to be compared it is
necessary to determine what the height would be under agreed conditions (BS 2520). These
conditions are a temperature of 0 °C and at a gravity of 9.80665 m/s“. As a direct consequence
of the definition of the pressure units, reduction of a reading to standard conditions is
equivalent to calculating the pressure, in Pa, from the equation:

P = pgh (3.2)

42
r

DB BROWSE, REVISED KEEN

where /? is the height of the mercury column in m,


p is the density of mercury at the temperature of measurement, kg/m^ and
g is the local value of gravity, m/s^.

With equation (3.2) it is possible to convert readings to different scales,

P = Psgshs = p,gLh,

where the subscripts S, L and t refer to standard, local and temperature t °C, respectively.

3.2.3 Pressure units

The units are discussed in Chapter 1, but are mentioned here because barometry and
manometry have given rise to a large number of units. The SI unit of pressure is the pascal (Pa)
and its multiples. They will be used exclusively throughout this chapter.

In barometry, the millimetre of mercury at various temperatures (mmHg), inch of mercury


(inHg) and the millibar (mbar) were commonly used in the past and some of them still exist. In
manometry, besides mmHg and inHg (the mbar is almost never used for manometers,
although, it is used by various manufacturers on some vacuum gauges), other common units
are mmH2O, cmH20 and inH2O.

Some of these units will take a long time to disappear, especially mmHg and mmH20. The
main reason being that they are visual units, i.e., we can visualise a pressure equivalent to the
height of a column of mercury 10 mm high, for example. However, the pascal has the very real
advantage that its introduction has abolished the need to worry about standard conditions (see
section 3.2.2) and greatly simplified the comparison of instruments with different scales and
built-in correction factors.

3.3 Types of barometers


(1) Primary standard instruments

Many mercury barometers are primary standards, but some are more accurate than others. The
latter are usually found in national standards laboratories, most of which have a primary
standard barometer/manometer. All of these are basically a U-tube with various methods of
detecting the mercury surfaces, such as capacitance measurements, white-light interferometry,
laser interferometry, ultrasonic interferometry etc. In general, the accuracy of the instruments
is limited by the uniformity and stability of the temperature and knowledge of the density of
mercury. The aim of this chapter is to discuss pressure measurement with laboratory
barometers/manometers and, so, the primary instrument used in the CSIRO National
Measurement Laboratory will only be described briefly. More information on the primary
standards in the national laboratories can be found in [2].

A diagram of the interferometric manometer of the National measurement Laboratory is shown


in Figure 3.2. It consists of a Michelson interferometer, which uses the two mercury surfaces

43
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

of a U-tube manometer as the mirrors of the interferometer. The pressure is measured by


counting the interference fringes as the surfaces move from zero to the pressure to be
measured. At the wavelength of laser light used, 2.3 X 10^ fringes are counted in going from
vacuum to atmospheric pressure. A more sophisticated laser system is being installed to
improve the fringe counting efficiency.
PHOTOMULTIPLIER

POLARCMO
ANALYSERS
BEAM SPLITTER
PHOTOMULTIPLIER

Figure 3.2: Interferometric manometer at the National Measurement Laboratory

(2) Laboratory mercury barometers

Laboratory mercury barometers fall into two broad categories:

1. Those where both the upper and lower surfaces are read e.g., Fortin, U-tube, micrometer
and siphon barometers.
2. Fixed cistern barometers where only the position of the top of the column is read, e.g.,
Kew pattern barometers, both station and gauge.

44
DB BROWSE, REVISED KEEN

Figure 3.3: Cistern of a Fortin barometer

3.3.1 Fortin barometer

The Fortin barometer is the most commonly used laboratory mercury barometer. The essential
feature of the design is that the level of the lower mercury surface is brought up to the tip of a
fiducial point, which is securely fixed to the roof of the cistern and corresponds with the zero
of the scale. The cistern mercury is contained in a leather bag and the adjustment is achieved
by means of a screw acting on the bag. Because of the limit of the size of the bag, these
barometers are used mainly for measurement of atmospheric pressure over the range 800 to
1100 hPa. The height of the mercury column is obtained by means of a sighting ring, or
cursor, used to set on the mercury meniscus. Pressure equilibrium with ambient air is achieved
either by a vent in the top of the cistern or through a porous plug or washer. Tube diameters
vary from about 6 to 12 mm, or even larger, where high accuracy is desired, as the bigger the
bore the smaller the errors due to variation of the mercury meniscus. Figure 3.3 shows the
cistern of a Fortin barometer.

3.3.2 U-tube barometers and manometers

A few instruments are made commercially, a simple glass U-tube with a scale and evacuated
with a good rotary pump works very well provided it is made of sufficiently large bore (greater
than 6 mm) so that meniscus effects are small.

One manufacturer makes an instrument in which the two arms are connected by a flexible tube
and one arm is raised or lowered to return the mercury level in that arm to a fixed position in

45
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

the tube, corresponding to a null in the other arm. Only one arm of the instrument is read.
While a convenient instrument to use, its accuracy is limited by meniscus effects and change in
volume of the flexible tube.

3.3.3 Micrometer barometer

This instrument uses a micrometer to measure the height of the mercury in the cistern. The
instruments are normally used as gauge barometers for precision calibration work. At equal
pressure in both arms, the zero of the instrument can be noted using the micrometer and a
cathetometer to measure the distance between the height of the meniscus inside the tube and
scale zero point. The capillary depression is eliminated for large diameter tubes (greater than
12 mm) as any variation is negligible (Table 3.1.). This instrument has as the advantage of
being a primary standard in that only the scale and the micrometer require calibration. Figure
3.4 shows a diagram of a micrometer barometer.

Figure 3.4: Micrometer barometer

3.3.4 Siphon barometer

This instrument requires only the scale to be calibrated for it to be a primary standard. It is, of
course, of short range and used mainly in standards laboratories. Figure 3.5 shows a siphon
barometer.

46
DB BROWSE, REVISED KEEN

Figure 3.5: Siphon barometer

3.3.5 Kew type or “fixed cistern” barometer

Figure 3.6 shows a Kew type barometer. Such barometers give direct indication of the
pressure by means of a single setting on the top of the mercury column. The small variations
in the level of the mercury in the cistern, which are proportional to the changes in the height of
the mercury column, are automatically allowed for by means of a linear, but contracted scale.
The amount of the contraction is fixed by the cross-sections of the tube and cistern. In practice
the contraction is usually not less than 5%.

AiR TRAP

VENT HOLES LEATHER WASHER

■ MERCURY

Figure 3.6: Kew station barometer

In contrast to the Fortin barometer, the reading on the Kew barometer depends critically upon
the amount of mercury and may be adjusted by adding or subtracting mercury. It follows that
Kew Barometers respond to the change in volume of mercury with temperature and, therefore.

47
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

have a different temperature coefficient from that of the Fortin type. Also, the reading depends
upon the uniformity of the tube and cistern, the Fortin is independent of any such variations.

A Kew station barometer is similar in appearance and range to a Fortin and is designed to hang
on a wall, usually in a gimbal to keep it vertical. A Kew gauge barometer is usually a long
range instrument designed to stand on a bench, and has a nipple to connect it to another
instrument. One instrument has two cisterns placed symmetrically either side of the tube, and
this enables the scale to extend to zero as well as making the reading less sensitive to tilt of the
barometer.

Some barometer manufacturers provide various correcting mechanisms for temperature and
gravity, e.g., a “slack” scale that can be stretched to compensate for the temperature. Such
devices neither affect the operation of the barometer nor improve their accuracy. They merely
improve the convenience of use, and so will not be discussed.

3.4 Reading Fortin and Kew barometers


Reading a barometer requires practice, as setting too high or too low produces errors. This is
illustrated in Figure 3.7.

To read the barometer, the mercury level is raised until it touches the pointer (Fortin
barometer). The setting is correct when the pointer and its image just touch. The top of the
mercury column is then set by bringing the bottom of the cursor tangential to the meniscus.
This is most conveniently done by adjusting the cursor until the top of the meniscus and the
front and back of the cursor are in line (Figure 3.7). Another method is to bring the cursor
down until it just cuts off a chord about 1 mm in length on the meniscus. These two methods
produce the same result in practice. When the barometer has been adjusted it should be firmly
tapped and the mercury level in the cistern and the pointer and cursor setting readjusted if
necessary. Tapping frees the mercury from sticking to the glass walls and allows the menisci
to be properly formed. The height of the column is then read from the scale beside the tube.

A number of other methods are used for detecting the mercury surface, particularly on
reference instruments in standards laboratories. The photoelectric detector will be mentioned
as it is the only really viable commercial method. A photodetector is used to measure the
amount of light passing between the mercury meniscus and the cursor, and the height of the
cursor is adjusted to return the detector to a fixed output. In general, a resolution of 0.01 mm
is obtainable with such detectors, but they need to be set for the correct pressure, and unless
the barometer can read zero pressure this means adjustment with a reference instrument is
required.

The use of floats in mercury barometers or manometers (with various electrical and optical
devices for sensing the position) is usually not as accurate as a good visual setting. This is due
to variable meniscus effects, but often the convenience of an electrical output is the overriding
consideration.

48
DB BROWSE, REVISED KEEN

Figure 3.7: Parallax errors in using a cursor to determine the top of the mercury
meniscus of a barometer or manometer

3.5 Pressure measurement using Fortin and Kew barometer


To obtain correct pressure values from a barometer, the barometer readings should be
corrected for temperature and gravity, to the value they would read at standard conditions.
Moreover, depending on the construction and location of the barometer, the readings may also
need to be corrected for capillary errors, elevation or even the vacuum above the mercury in
the barometer.

3.5.1 Temperature correction and measurement

For most Fortin and Kew barometers the temperature is usually measured by a mercury-in-
glass thermometer attached to about the mid point of the brass scale, with the bulb embedded
in a metal shield. At 1000 hPa, the reading of a Fortin barometer changes by 0.16 hPa per °C,
so a temperature measurement with an uncertainty of 0.1 °C should be adequate for most
barometers.

It is generally assumed that the scale and mercury column temperatures are identical. This
assumption is probably without significant error provided that the thermometer reading has not
changed appreciably (0.1 to 0.2 °C) in the previous 30 to 60 minutes. However, standing close
to a barometer to read it will significantly influence the temperature. Provided the time is
reduced to a minimum, and the thermometer is read first, the temperature uncertainty should
be reduced to the order of 0.2 °C.

To correct the pressure readings of the barometer to standard temperature (0 °C), the
corrections given below should be made. The symbols used in the equations are:

49
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

a is the internal area of the tube at 0°C,


ht is the length from the zero plane to the lowest point of the graduated section of the
scale,
r is the barometer reading at 0°C
r, is the corresponding barometer reading at t°C,
V is the total volume of mercury in the barometer,
A is the effective area of horizontal cross section of the cistern (the area of the cistern
minus any area occupied by members projecting through the mercury surface),
a is the coefficient of linear expansion of the brass scale, 0.000 018 4/°C,
/3 is the coefficient of dilation of mercury, 0.000 181 8/°C,
Y is the coefficient of linear expansion of the tube,
S is the coefficient of linear expansion of the cistern,
3 rf is the coefficient of volume expansion of the cistern, allowing appropriately for the
expansion of steel and glass, 0.000 030 0/°C and
t is the temperature read by the thermometer attached to the instrument.

(1) Fortin-type barometers

This category of barometer includes U-tube, micrometer and siphon barometers. They have a
temperature coefficient that depends solely on the thermal expansion of the mercury and of the
scale (conventionally brass). The reading at 0 °C is given by:

1 + pt

Most tables, such as those in BS 2520 [1], tabulate r^Ta}i/{ \ -\-[31) and this is to be subtracted
from the barometer reading. At atmospheric pressure and 20 °C this correction is about 3.3
hPa (2.5 mmHg).

(2) Kew-type barometers

The temperature correction for fixed cistern barometers is given by

rj{/3-a}t hj{a-l3}t /I + « f4(y^-2^)r + 2r,6z(J-/)r1


1 + /3t 1 + /3t A [ ^[l + i^/3 + 2^)r] J

The corresponding formula given in BS 2520 is:

0(^ <^)^ - - 3z7)r (3.6)


1 + /3t A

which is an approximation to (3.5). However, at 1070 hPa (800 mmHg) the difference
between equations (3.5) and (3.6) is 0.019 hPa (0.014 mmHg), which is about the resolution of

50
DB BROWSE, REVISED KEEN

a very high grade barometer. Therefore, equation (3.6) could be used for the correction of
Kew type barometers.

3.5.2 Correction for gravity

Since the reading of a mercury barometer depends on gravity, a correction for the departure of
local gravity from standard gravity, gs = 9.80665 m/s\ is necessary. If the value of gravity at
the site of a barometer is not known, then ignoring local anomalies, the gravity at latitude 0
and at height H (m) above mean sea level is given by the following formula, which was
recommended by the International Association of Geodesy:

= g^H = 9.780 318 4(1+0.005 302 4 sin"<A o.OOO 005 9 sin-20) - 3.086x10’^// (3.7)

Alternatively, the local gravity value may be found from the National Geoscience Datasets
maintained by Geoscience Australia (GA). The information can be accessed online through
their website. If the exact location cannot be found, one can find the g value and the latitude of
the closest station maintained under the Australian Fundamental Gravity Network (AFGN),
which can also be accessed through Geoscience Australia.

The pressure at standard gravity is then calculated by

Ps = PLgL/gs- (3.8)

3.5.3 Capillary errors

A capillary rise (or depression) will be present when a curved interface exists in a liquid at
rest. Theoretically, the amount of depression is a function of the bore of the container, the
meniscus height and the value of the surface tension of the mercury. Because barometers are
usually not symmetrical instruments the capillary depression of the mercury in the tube and the
cistern will not be equal and so must be allowed for. The amount of depression tends to vary
with the age of the barometer, with the direction of the change in pressure and with local
differences in the condition of the surface of the container. Except for freshly distilled mercury
the surface tension of mercury is not well known.

Thus, the capillary depression, a function of the surface tension for a given bore of tube and
meniscus height, cannot be accurately determined. Errors in the capillary depression have
been estimated to be about ±20%, which allows an uncertainty that is adequate for normal
barometers and manometers.

For highest accuracy the only safe recourse is to make the bore of tube 40 mm in diameter or
larger, so that the capillary depression is negligible. In precision manometry and barometry it
is desirable to measure the heights of the menisci and to apply the capillary correction. This,
however, is not considered worth doing for tubes of less than 10 mm in diameter. The
uncertainty in the capillary correction is such that, if an uncertainty of 0.13 hPa (0.1 mmHg) is
required, then a tube of not less than 12 to 16 mm in diameter should be used.

51
CHAPTER 3. BAROMETRY, MANOMETRY & COEUMNS

The correction for capillary error is given in Table 3.1. A knowledge of the surface tension is
required, but it has been accepted that the value of 450 mN/m is as applicable as any. For
Fortin and fixed-cistern barometers the capillary depression given in Table 3.1, should be
added to the reading of the barometer.

Table 3.1. Correction in pascals for capillary depression at a surface tension of 450 N/m,
after Gould & Vickers [3].

Bore of tube Meniscus height in mm


(mm) 0.1 0.3 0.5 0.7 0.9 1.1 1.3 1.5 1.7 1.9 2

6 17 51 82 111 137 160 177 191 200 - -


7 12 35 57 78 97 114 128 140 148 155 -
8 8 25 41 57 71 83 94 104 111 117 119
9 6 18 30 42 52 62 71 78 84 89 91
10 5 14 23 31 39 47 53 59 64 68 70

12 3 8 13 18 23 27 31 35 38 40 41
14 2 5 8 11 13 16 19 21 23 24 25
16 1 3 5 6 8 10 11 12 14 15 15
18 1 2 3 4 5 6 7 8 8 9 9
20 0 1 2 2 3 3 4 5 5 5 6
22 0 1 1 1 2 2 2 3 3 3 3

It is common practice with ordinary barometers and manometers to incorporate the capillary
depression into the scale. For U-tube instruments it is commonly assumed that this correction
can be neglected, on the assumption that it is the same for each surface. However, for
precision work it should be noted that one surface is rising and the other falling and this will
cause a different meniscus height.

Capillary depression cannot normally be applied to the cisterns of barometers, as the meniscus
height cannot usually be measured. The cistern should be large enough so that the depression
is negligible. With time and contamination the meniscus in the cistern can flatten, altering the
level, or volume, of the mercury. At constant pressure this volume effect causes the barometer
reading on the fixed scale to fall as the meniscus flattens. This effect is reduced for cisterns of
larger diameter and does not affect Fortin barometers.

3.5.4 Elevation correction

A measurement of either differential or absolute pressure made with any manometer or


barometer, liquid or mechanical, applies only to the elevation at which it is made. If pressures
are desired at any other level above or below, the decrease or increase in absolute pressure
with elevation to the intervening head of gas (or liquid) must be allowed for. The elevation
correction is the correction that needs to be applied to transfer the pressure measured at one
elevation to that at another elevation.

52
DB BROWSE, REVISED KEEN

For most barometric pressure measurements where it is necessary to correct for instruments at
different levels it is sufficient to make a linear allowance as given in Table 3.2.

Table 3.2 The allowance per metre of vertical distance between the cistern of the
barometer and the point at which the pressure is required.
mbar (hPa) mmHg Pa
Barometric pressure 1000 760 100 000

Allowance per metre 0.118 0.089 11.8

In the use of manometers for the calibration of gauge pressure for precision measuring
instruments, it is necessary to allow for the change in pressure on the low pressure side
(usually open to the atmosphere) due to the change in column height, as well as any head on
the pressure side.

For greater elevation changes, the correction also involves computing the change in gas
pressure in going from one level to another. Since a gas compresses under its own weight, its
density at a higher elevation is less than at a lower elevation. The basic relation is:

dP/dh = - pg. (3.9)

Substituting p = PM/RT and integrating, one obtains:


-Mgh

P = P^e (3.10)

Hence P ^P„0 -hMg/RT) (3.11)

where g is the acceleration due to gravity,


h is the vertical height of the gas column,
M is the molar weight of the gas, air = 0.029 kg,nitrogen= 0.028 kg,
P is the absolute pressure at height equal to h,
P(, is the absolute pressure at the reference level of the column, i.e. h = Q
R is the universal gas constant = 8.314 J/K.mol and
T is the absolute temperature of the gas.

Equation (3.11) is an approximation that is valid for h up to 100 m. If the level at P is above
the level at the sign of h is positive, so that the correction reduces the measured pressure; if
below, h is negative.

3.5.5 Estimation of air in the vacuum space of a barometer

The vacuum space above the mercury column in a barometer is never perfect, as mercury
vapour and air are always present. The pressure in the vacuum space must be added to the
measured pressure to obtain the true pressure. In some barometers the vacuum space can be
evacuated with a vacuum pump and the pressure measured with a McLeod gauge (which does

53
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

not measure the pressure of the mercury vapour - Chapter 2). Some barometers have a
mercury-sealed valve to enable the tube to be sealed and to allow re-evacuation. Some
commercial instruments are sold as manometers and are provided with a vacuum line to enable
them to be used as barometers. These lines are often too small in cross-section and too long to
provide a vacuum commensurate with the accuracy of the instrument. Whatever system is
used, care should be taken to ensure that the pressure in the vacuum space is sufficiently low.

For barometers that have a sealed tube, it is necessary to occasionally check for air in the
vacuum space, as, once such a barometer is installed, this is the main source of error. A
qualitative and a quantitative approach for checking are described here.

A qualitative method of estimating the amount of air in a barometer tube is the sound produced
at the top of the tube by the impact of the mercury with the glass. If the sound is sharp and
metallic the vacuum is judged to be satisfactory; whereas, if the sound is characterised as soft,
dull, or leaden, it is generally inferred the vacuum is impaired in quality. The difference in
sound in the latter case is considered to arise from the fact that a bubble of air would act as a
buffer between the mercury and the glass at the top of the tube.

The so-called ‘metallic click’ is best produced while the barometer is inclined about 45° and
occurs just as the mercury, moving up the tube, reaches the top and completely fills it. The
greatest caution, i.e., tilt slowly, should be exercised in producing the click as it is possible to
damage the tube.

A quantitative method is to measure the size of the air bubble at the top of the tube while the
instrument is lying in a horizontal position. Following steps (1) and (2) of section 3.8.1, lay the
barometer in a horizontal position on a flat surface. Next, the hanger ring, cover cap and any
inserts should be removed from the top of the instrument. The upper portion of the tube should
be inspected to see whether an air bubble is visible. If no bubble is visible, the barometer
should be restored to its normal vertical position, tapped gently, and again inclined and
brought to a horizontal position. The tube is again inspected for the presence of a bubble. It is
emphasised that the greatest care should be exercised in carrying out this procedure.

If an air bubble is present then its diameter should be measured by means of a scale. The effect
of the air bubble on the reading of the barometer can be calculated, as indicated below, using
the following dimensions that must be obtained or measured:

(1) the internal diameter of the tube, D;


(2) length of tube between the 101325 Pa scale position and the top of the tube, h and
(3) the diameter of the air bubble when the tube is horizontal, d.

The pressure of the air in the vacuum space (or the change in reading of the barometer due to
the air) is given by:

P = PannV/V (3-12)
where Patm is the pressure of one atmosphere in the appropriate units (101325 Pa, 1013.25
hPa, 760 mmHg);

54
DB BROWSE, REVISED KEEN

V is the volume of the bubble, assumed to be equal to half the volume of a sphere of
diameter d‘,
V is the volume of the vacuum space of length h, the curved end of the tube is ignored.

Thus, P = Panud'/SD-h. (3.13)

If the barometer has been calibrated with a small amount of air in the vacuum space then its
presence is not important unless the temperature changes. The change in pressure due to the
change in the temperature of the barometer is:

AP = P[(273+Z2)/(273+r/)-1], (3.14)

where tj is the initial temperature or the temperature of calibration, and


?2 is the final temperature, both in °C.

These effects are summarised in Table 3.3 for a tube with a 100 mm long vacuum space.
Because of difficulties in measuring the tube diameter (unless it is known) and the bubble
diameter (it will be slightly wrong due to the lens effect of the end of the tube, and the volume
will not be exactly equal to a half sphere), the figures in Table 3.3, and equations (3.13) and
(3.14), should not be relied upon to better than ±30% in a practical situation.

3.6 Calibration of mercury barometers


Barometers and manometers are calibrated by comparison with a reference instrument,
preferably of the same type, as this eliminates corrections for temperature and gravity.
Manometers and Kew gauge barometers may be connected directly to the reference
instrument. However, Fortin and Kew station barometers must be placed in a pressure chamber
if they are to be calibrated over their full range. The reference barometer (usually a Fortin) is
also mounted in the chamber. This has the advantage of ensuring that both instruments are at
the same temperature, and if the reference and test instruments are of the same type,
temperature corrections need not be applied.

3.6.1 Frequency of calibration

Once a mercury barometer has been calibrated and mounted on the wall or installed on the
bench there is little that can happen to it, except outgassing of the tube, which, in a well-
prepared barometer should be minimal. The mercury surface in the cistern will eventually
become dirty to a greater or lesser extent, but except possibly for a gauge barometer, this does
not usually affect the accuracy significantly, unless, for a Fortin barometer, the pointer image
cannot be seen. When this happens, the barometer requires servicing and recalibrating. Some
Fortin barometers have a trap in the cistern for removing dirt from the surface. If dirt on the
surface is troublesome, then the adjusting screw should be wound out to its extremity and then
back again. This usually provides a cleaner surface.

55
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

Table 3.3. Pressure due to air in the vacuum space of a barometer. The vacuum space is
assumed to be 100 mm long. The second row of values (shaded), are the changes
in pressure, due to the air, if the barometer temperature is changed 10 °C.
Bubble Internal Diameter of the Tube - (mm)
Diameter
(mm) 4 6 8 10 12 15 20
Pressure (Pa)
0.5 2.6 1.2 0.7 0.4 0.3 0.2 0.1
OJ 0 0 0 0 0

1 21.1 9.4 5.3 3.4 2.3 1.5 O.S


az 0.3 0.2 0.1 0.1 0 0

1.5 71.2 31.7 17.8 11.4 1.9 5.1 2.8


2,4 U 0.6 0.4 0,3 0.2 O.l

2 169 75.1 42.2 21.0 18.8 12.0 6.^^


5.3 2.6 1.4 0.9 0.6 0.4 0.2

2.5 330 147 82.5 52.^ 36.6 23.5 13.2


11.3 5.0 2.3 1.8 1,3 0.8 0.5

3 570 253 143 91.2 63.3 40.5 22.S


19.5 3.6 4.9 3.1 2.2 L4 0.8

4 1351 600 2i6 150 96.1 54.0


46.1 20.5 11.5 7.4 5,1 3,3

5 - 1173 660 422 293 188 106


- 40.0 22.5 14.4 10.0 6.4 3.6

The calibration of a barometer only needs to be checked at one point after two years and then,
if there is no change, every five years. Provided the barometer is not shifted and the mercury in
the cistern is not too dirty there is no need for a more regular calibration. Because transporting
barometers is difficult and risky a number of alternatives to recalibration at a calibration
laboratory have been proposed that involve checking the barometer at one point in-situ.

3.6.2 Comparison with a remote barometer

One method of calibrating barometers in-situ has been to compare the barometer reading with
that of a reference barometer situated some short distance away, often done by telephone. This
can be a difficult and not very precise operation. Some criteria for knowing the value of the
pressure within the vicinity of a reference barometer are given in ASTM D 3631-99 [4].
These are as follows:

56
DB BROWSE, REVISED KEEN

Clause 1.2 In the absence of abnormal perturbations, atmospheric pressure measured by these
methods at a point is valid everywhere within a horizontal distance of 100 m and
a vertical distance of 0.5 m of the point.
Clause 1.3 Atmospheric pressure decreases with increasing height and varies with horizontal
distance by 1 Pa/100 m or less except in the event of catastrophic phenomena (for
example, tornadoes). Therefore, extension of a known barometric pressure to
another site beyond the spatial limits stated in 1.2 can be accomplished by
correction for height differences if the following criteria are met:
1.3.1 The new site is within 2000 m laterally and 500 m vertically.
1.3.2 The change of pressure during the previous 10 min has been less than 20 Pa.

Clause 10.2 The atmospheric pressure Pi, height hi, and atmospheric temperature Ti at some
Site 1 and the height h2 at a Site 2 are known then the atmospheric pressure P2 at
Site 2 can be calculated from the following equation (c.f. equation (3.10)):

„ O.O68332(/2, - hJ
P, = P, exp-----------------!------ —, (3.15)
T^+T^
where: Pj = pressure at Site 1, Pa,
P2 = pressure at Site 2, Pa,
hl = height above mean sea level of Site 1, m,
Z?? = height above mean sea level of Site 2, m,
Ti = atmospheric temperature at Site 1, K, and
T2 = atmospheric temperature at Site 2, K.

While application of the above method may seem attractive, the telephone comparison of
barometers is considered to be neither necessary nor capable of better accuracy than about 200
Pa (1.5 mmHg) for the following reasons:

(1) Unless something catastrophic occurs, a Fortin barometer will almost certainly not change
its calibration by 200 Pa over most of its life. Hence the comparison is not necessary.
(2) Not all laboratories know their height above sea level to sufficient accuracy.
(3) Some laboratories are air-conditioned and can have a pressure different from ambient. A
typical air-conditioned laboratory can have a pressure of 40 Pa above ambient.
(4) The formula given in section 10.2 of ASTM D 3631 (equation (3.15) above) requires a
knowledge of the atmospheric temperatures at the two sites. These are not the same as the
temperatures measured inside the laboratories.
(5) When the pressure is obtained from another laboratory, it is essential to find out if it is
merely an instrumental reading (which is perfectly adequate when comparing two similar
instruments in the same room) or whether it has been corrected to standard conditions and
allowance made for height above sea level. The Bureau of Meteorology, for example,
quotes the pressure at mean sea level.

Aneroid barometers (mechanical and electrical) are in general not as stable or as accurate as
mercury barometers, but since they are more readily transported than mercury barometers they
can be taken to a reference barometer. However, if a laboratory uses an aneroid barometer

57
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

regularly in its work, then it should possess a mercury barometer to check the aneroid at one
point every six months, or more frequently if it is moved regularly.

3.6.3 Checking mercury barometers in-situ

There are a number of mechanical and electrical aneroid barometers that have sufficient
resolution and stability for them to be used as transfer standards. After calibration against a
standard barometer or pressure balance at several points around atmospheric pressure, these
instruments could be taken by hand to check the mercury barometer at one point (atmospheric
pressure) in-situ. The transfer standard should be rechecked on completion. The correction to
the mercury barometer is calculated by taking the difference between the reading of the
mercury barometer and the average of the before and after readings of the calibrated transfer
standard. The reading of the mercury barometer is corrected for standard conditions (0 °C),
from equation (3.4).

{p - a}t
Pc =P,rc„ -PcSl’’, (3.16)

where Pc is the correction to the Fortin barometer, in Pa,


P
* Iran is the average of calibrated pressure readings of the transfer standard, in Pa,
is the average barometer reading at t°C, taken in mmHg, but expressed in m,
A is the density of mercury at 0°C, in kg/rn^
gL is the local gravity in-situ, in m/s^
a is the coefficient of linear expansion of the brass scale, 0.000 018 4/°C, and
p is the coefficient of dilation of mercury, 0.000 181 8/°C.

An example of the uncertainty calculation for the calibration of a Fortin barometer is discussed
in the following section. The calibration need only be carried out every five years.

3.7 Uncertainties in calibrating a Fortin barometer in>situ


As transportation of a mercury barometer is risky and causes safety concerns, sending a Fortin
barometer for calibration is not recommended. The barometer can be calibrated in-situ by a
calibration laboratory, e.g., CSIRO National Measurement Laboratory. The user can also use
this method to check their own Fortin barometer, if they have a suitable calibrated transfer
standard. An example calculation of uncertainty for this kind of calibration is described below.

Example

A digital barometer (transfer standard) was calibrated using a piston gauge (reference
standard) in the range of 800 to 1100 hPa. The digital barometer was then used to check a
Fortin barometer (test instrument) at one point at atmospheric pressure -1010 hPa. Six sets of
measurements comparing the digital barometer and the Fortin barometer were done. The
digital barometer was placed at the same height as the pointer above the cistern of the Fortin
barometer. The digital barometer was then rechecked using a piston gauge at the standard

58
DB BROWSE, REVISED KEEN

laboratory at about this pressure (-1010 hPa). A shift of +2 Pa was recorded for the digital
barometer compared with the first calibration.

3.7.1 The model (of the measurement)

The correction of the Fortin barometer is obtained by taking the average of the six pressure
differences between the digital barometer and the pressure reading of the Fortin barometer.
The ‘model’ for the measurement is expressed using equation (3.16). The uncertainty in the
correction Pc is determined by considering all uncertainty contributions to all the values used in
its calculation, i.e., in Ptran, Ps, gh >'i- t, ck, of equation (3.16).

The first step, as explained in [6], is the calculation of sensitivity coefficients c, for all these
parameters. They are needed to convert their uncertainties into components of uncertainty for
Pc. The coefficients are obtained by differentiating equation (3.16).

Here, we use the values: Ps = 1.36 x 10"^ kg/m\ 9.80 m/s^, r, = 0.760 m, t =20 °C and
with <7, as given above.

For Pjran, the coefficient Ci

1
for A = —1.^2 Pa/kg/m^ =-0.101 Pa/ppm,

where the term in the brackets [ ] is equal to 0.9967.

For gL, 9gt ^"L l + /» . = —10302 Pa/m/s\

for r, = -1.328x10'' Pa/m = -132.8 Pa/mm,

for t. 16.4 Pa/°C .


It
The effects of error in a and P are insignificant, so c, for these quantities is not required.

For each error source, the standard uncertainty is given the symbol u(Xi), evaluated as Vi / Jfc,-,
and its effect on the measurand Pc is Ic,- u(Xi)l. The symbols and terminology follow [5, 6].

The next step is to list all components and to assign values of uncertainty, reducing factors and
numbers of degrees of freedom. The components are described below and their assigned
values are presented in Table 3.4.

59
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

3.7.2 The components of uncertainty

(1) Calibration uncertainty of the digital barometer

From the calibration report of the digital barometer, when calibrated against a piston gauge
reference standard, the expanded uncertainty is 4 Pa and the coverage factor is equal to 2. The
degrees of freedom can then be taken as 50. Therefore, f/, = 4, Pa, ki=2 and v, = 50.

(2) Instability of the digital barometer

The digital barometer was calibrated twice, once before the calibration of the Fortin barometer
and once afterwards, and indicated a drift of 2 Pa. The correct reading of the digital barometer
during the test of the Fortin barometer was assumed to lie between these two calibrations. We
take the uncertainty due to drift in the digital barometer to be the semi-range, 1 Pa, with a
rectangular distribution. Therefore f/, = 1 Pa, L = V3 and v, = 1000.

(3) Repeatability of measurements (type A)

Six pressure differences between the digital barometer and the Fortin barometer were obtained
within an hour. The correction was the mean value of these measurements. The standard
deviation is 5 Pa. The repeatability uncertainty is then equal to the standard deviation of the
mean of these 6 pressure differences, which is -7^= 2.041, with 5 degrees of freedom. U, =
<6
2.041 Pa, kj = 1 and v, = 5.

(4) Uncertainty of temperature measurement

The thermometer used was a mercury-in-glass thermometer mounted on the body of the Fortin
barometer with a resolution of 0.1 °C. As the thermometer reading did not change more than
±0.2 °C and the thermometer was always read first, the uncertainty in the Hg temperature was
considered to be 0.2 °C, as a -95% estimate. So, U, = 0.2°C, L = 2, c, = 16.4 Pa/°C, from
above, and since U, was a ‘reasonable’ estimate [6], v, = 8.

(5) Uncertainty of gravity measurement

The gravity of the laboratory where the Fortin barometer is located can be calculated from
equation (3.7) if the latitude is known. Alternatively, the local gravity value may be found
from the National Geoscience data sets. In this case, we find the g value and the latitude of the
closest station maintained under the Australian Fundamental Gravity Network (AFGN) via the
Geoscience Australia web site http://www.ga.gov.au.

The standard uncertainty of g estimated in this way is, conservatively, 0.0005 m/s"' a
reasonable estimate. The degrees of freedom can be taken as 8. Therefore, U, = 0.0005 m/s“, L
= 1, c, = 10302 Pa/ m/s^, from above, and v, = 8. The pressure uncertainty due to this
component should be about 5 Pa.

60
DB BROWSE, REVISED KEEN

Table 3.4 Uncertainty analysis table for an in-situ calibration of a Fortin barometer

Component Values
Ic/ u(Xi)\
Component (/ = 1, 2,... 8) Ui ki Ic/I (Pa) Vi Ic,- u(Xi)\^ Ic/ u(Xi)\‘^/Vi

1. Calibration of barometer (Pa) 4 2 1 2.000 50 4.000 0.32


2. Instability of std (Pa) 1 1.73 1 0.578 1000 0.334 1.1E-04
3. Repeatabiltiy (Pa) 2.041 1 1 2.041 5 4.17 3.47
4. Temperature measurement (°C) 0.2 2 16.4 1.64 8 2.69 0.90
5. Gravity (m/s^) 0.0005 1 10302 5.15 8 26.5 88.0
6. Purity of mercury (ppm) 5 1 0.101 0.505 8 0.255 8.1E-03
7. Resolution of the scale (mm) 0 1.73 132.8 0 1000 0 0
8. Rounding of final result (Pa) 0.5 1.73 1 0.289 1000 0.084 7.0E-6

Combined standard uncert. Uc (Pa) 6.167


Degrees of freedom Veff 15.6
Coverage factor k 2.13

Expanded uncertainty U (Pa) 13.14

(6) Purity of mercury

If the mercury in the Fortin barometer was clean and freshly distilled, the density is 13595.08
kg/m'^ at 0 °C (standard condition) with a standard uncertainty of 3 ppm. In this example, it is
estimated the purity of the mercury is such that the standard uncertainty is ~5 ppm. Therefore,
Ui = 5 ppm, ki= 1, £?, = 0.101 Pa/ppm and v, = 8.

(7) Resolution of the scale

The Fortin barometer can be resolved to 0.1 mm. The resolution uncertainty is taken as half of
the resolution (i.e., ±0.05 mm) with a rectangular distribution. As the correction is expressed in
Pa, the sensitivity coefficient is equal to 132.8 Pa/mm and the degrees of freedom may be
taken as 1000. Therefore, (7, = 0.05 mm, E = , c, = 132.8 Pa/mm and v, = 1000.

However, since more than one reading was taken and scatter was observed, the effect of
resolution errors has already been included in the type A estimation, item 3 above.

(8) Rounding uncertainty

The final correction will be rounded to the nearest 1 Pa with a rounding uncertainty of ztO.5 Pa.
Thus, we put Ui = 0.5 Pa, C = VJ being rectangular, c, = 1 and v, =1000.

61
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

3.8 Care and handling Fortin and Kew barometers


3.8.1 Preparation for transport.

Mercury barometers are delicate instruments and must be handled with care. If the barometer
is to be moved a very short distance to another position on the same level in the same building,
then it can be carefully carried by hand in an upright position. Care must be taken that the
barometer is not tipped so that the mercury surges up the tube, or that the mercury in the
cistern becomes so agitated that the bottom of the tube is exposed, allowing air into the tube.

For longer journeys, Fortin and Kew station barometers must be carefully prepared by means
of the following steps.

(1) The adjusting screw of the Fortin barometer (Figure 3.3) is screwed up until it lightly
touches the end of the barometer tube. It is then backed off a half to one turn, this is to
allow mercury to flow in or out of the tube should the temperature change. The
barometer mounting board is then removed from the wall with the barometer attached.
Leaving it mounted on the board is important for the protection of the barometer.
(2) Both Kew and Fortin barometers are then carefully and slowly tilted over until the
mercury rises up and fills the tube. If there is no air in the vacuum space, a sharp
‘metallic click’ may be heard as the mercury reaches the top of the tube. The barometer
is then laid horizontal or tipped upside down. The barometer can be safely carried,
preferably by hand, in this position. It can also be safely transported long distances by
rail or road while in this position, providing it is packed as for any fragile instrument.
During the whole time of transport the top of the barometer tube must never be allowed
to rise higher than the cistern.
(3) The barometer is restored to operation by reversing the above procedure. It should be
raised to a vertical position very carefully while watching the column for any signs of
air bubbles.

It is the tilting at the beginning and the end that is the delicate operation, as this is the time
when air can enter the tube. Once the barometer is horizontal or upside down it is quite safe.

Kew gauge and other long range barometers are prepared for transport by evacuation of the
cistern, which is then sealed off, either with a valve or by other means such as clamping a tube.
The barometer is then carefully carried in an upright position. Many modern barometers have a
facility for evacuating the top of the tube. For these instruments the mercury should be
removed from the barometer prior to any transport, which cannot be done by hand.

Over a period of time, mercury tends to leave a mark where the surface comes into contact
with a glass tube and this causes difficulty in reading the barometer. To prevent this the
mercury should always be lowered from the position of usual reading. This is particularly
important for a Fortin barometer where the mercury should never be left in contact with the
index point. Lowering of the mercury is not possible for a Kew station barometer.

62
DB BROWSE, REVISED KEEN

3.8.2 Preparation for measurement

After the barometer has been brought upright and attached to the wall (or placed on a bench) it
is necessary to set it vertical. It is important to note that it is the scale that is to be vertical and
not the tube. For most Kew barometers, setting with a plumb line is adequate. Gauge
barometers have levels attached to the cistern and the barometer is correctly set up when these
are level. This is because the barometer is calibrated in this condition and any error due to
misalignment becomes part of the calibration.

The Fortin barometer usually has ferrules at the points of support enabling it to be rotated. The
mercury is brought into contact with the pointer, the barometer rotated 180° and the verticality
adjusted until the mercury is again in contact with the pointer. This is repeated until contact
between the pointer and the mercury surface is maintained for as full a rotation as possible.
All adjusting screws should then be tightened so that the barometer cannot move laterally,
although, it should still be free to rotate. Any lack of squareness between the scale and the
plate to which the pointer is attached contributes to the instrumental correction obtained at
calibration.

3.9 Aneroid barometers


The most common instrument consists of a corrugated chamber, formed by two thin metal
diaphragms, which has been evacuated. Its operation depends on the principle that a thin metal
disc or membrane responds elastically to a change of pressure on its faces. This movement is
magnified by mechanical linkages and translated into the rotary movement of a pointer on a
dial.

One sensitive instrument. Figure 3.8, measures the movement of a bellows with a micrometer
with the aid of an electrical contact indicator. The reading is made on the break so that the
micrometer is not exerting a force on the bellows at the time of reading. The instrument has a
resolution of 1 Pa.

, SWtTCW

Figure 3.8: Precision aneroid barometer showing the operation of the mechanical digital
micrometer and bellows system.

63
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

There are a number of other barometric pressure measuring devices that measure the
movement of a diaphragm capacitively, inductively, with strain gauges or the change in
frequency of an oscillating device. Such instruments usually have an electronic digital readout.

It is stressed that aneroid barometers are not fundamental instruments and so require
calibration with a standard barometer. With the advent of transducers with digital output, there
is a tendency for the “old fashioned” mercury instruments to be disregarded. This is a mistake
as they are the only convenient instruments suitable for use as standards in the barometric
range.

In reading these instruments, the temperature and gravity corrections for mercury barometers
obviously do not apply.

3.10 Manometers
Manometers are used to measure differential pressure. To measure pressure close to vacuum,
mercury is usually used as the manometer fluid. If it is to measure pressure relative to
atmospheric pressure, water or oil is used. Manometers can be divided approximately into two
types: micromanometers and inclined-tube manometers.

3.10.1 Micromanometers

When a single scale, vernier and sighting ring are used to measure the column height of a
liquid the practical limit of sensitivity is about 0.05 mm. or 0.5 Pa for water. This is inadequate
for the precision measurement of small pressures, hence the development of micromanometers.
They are used to measure gas pressure, either absolute or differential, in the range of 0.1 Pa to
around 7 kPa.

The advent of sensitive transducers, mainly of the capacitance type, has to a large extent
eliminated the need for micromanometers to measure pressure. They are, however, needed to
calibrate the transducers.

A common micromanometer employs distilled water with a hook gauge, which is an index
point fixed below the surface. The contact of the point with the surface is normally viewed
with a telescope from below the surface at an angle such that total internal reflection occurs at
the liquid surface (i.e., the surface acts as a mirror). Adjustment is made until the index point
and its image just touch.

There are a number of variations of this instrument where one or both arms of a U-tube are
read, and where the index points are micrometer adjustable from below the surface (Figure
3.9) or where the index points are fixed in the cistern and the cistern raised or lowered. The
latter system has the disadvantage that cistern movement disturbs both arms and so adjustment
requires care and patience. If two cisterns, with fixed or moveable index points, are connected
via flexible tubing, then the range of the instrument is limited only by the ability to measure
the height difference. With these instruments it is possible to set on the meniscus with an
accuracy of 0.01 to 0.02 mm, which is equivalent to a pressure of 0.1 to 0.2 Pa.

64
DB BROWSE, REVISED KEEN

Figure 3.9: Schematic diagram of a hook-gauge water manometer.

Various null reading micromanometers have been devised with different methods of setting the
null. The following gauges are described as examples.

(1) Chattock gauge.

Figure 3.10 shows a Chattock gauge. This instrument is a tilting manometer, developed by
Chattock in 1901, which balances the pressure to be measured against the pressure head caused
by the tilt. The null indication depends upon the observation of the interface between two
immiscible liquids of nearly equal density. The interfaces are arranged to be in tubes differing
in diameter by a factor of four. The sensitivity is enhanced by observing the interface in the
small diameter tube. A sensitivity exceeding 0.001 mm of water (0.01 Pa) is claimed.

Figure 3.10: Schematic diagram of a Chattock gauge micromanometer

65
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

(2) Bradshaw gauge.

Figure 3.11 shows a Bradshaw gauge. A sighting tube, inclined at a very small angle (usually
about 2°), connects to a reservoir through a rigid connecting tube at right angles to the sighting
tube. The pressure is measured by tilting the reservoir and connecting tube while observing
the meniscus in the sighting tube through a low power microscope. The effect of tilt is
compensated for mechanically by tilting about an axis passing through the centre of the sight­
tube meniscus and supporting the reservoir on the flat top of the jacking screw by means of a
part cylindrical foot. The centre of curvature of the foot is adjusted to coincide with the surface
of the liquid in the reservoir, so that the height of the liquid surface above the top of the
jacking screw remains constant and equal to the radius of curvature as the liquid reservoir is
tilted. The inclination of the reservoir does not affect the liquid level provided that its cross
section is symmetrical about an axis passing through the centre of the cylinder and parallel to
the axis of tilt. Since the position of the sight-tube meniscus is kept constant, the difference in
liquid level between the two arms is equal to the distance through which the jacking screw has
been moved. The instrument has a sensitivity of about 0.01 Pa.

Figure 3.11: Bradshaw gauge micromanometer

(3) Point-contact micrometer manometer

This manometer can be used to measure pressure at rough vacuum. The manometer fluid of
this type of manometer is mercury and micrometer driven points are employed to locate the

66
DB BROWSE, REVISED KEEN

liquid surface and measure the column heights. The usual range of such instruments is 50 or
100 mm depending upon the micrometers. A schematic diagram of such a mercury manometer
is given in Figure 3.12. In this version, large-bore glass tubes are clamped between stainless
steel end plates and sealed with Viton O-rings. Vacuum ports are provided in the top plate.
Sensitive levels are mounted on the platform to avoid sine and cosine errors. O-rings seal the
micrometer spindles. The temperature is measured as temperature change leads to a change in
density and hence the pressure measured. The micrometers are initially set to the liquid
surfaces with zero differential pressure on the columns. The projection of the spindle beyond
the micrometer thimble can cause an error due to temperature effect resulting from heating of
the micrometers as a consequence of manipulation by hand. This effect can be calculated and
allowed for, but it requires temperature measurement of the spindles and the micrometers. It is
better to reduce the effect by insulating the thimbles should this be a problem.

Figure 3.12: Schematic diagram of a precision point-contact U-tube micrometer


manometer

The position of the liquid surface may be detected by a number of techniques;

(1) Formation of a dimple caused by the contact of the point with the surface. The dimple is
observed by the distortion of a geometrical pattern, reflected in the liquid surface as
viewed with a magnifying lens from above with mercury and from the surface level in the
case of oil.

(2) The point and its image as reflected in the surface just touched.

(3) With mercury it is possible to make an electrical contact indicator, but a slight scum on
the surface can cause erratic readings making the technique unreliable.

The different methods give different readings and so should never be mixed in a series of
measurements. The measurement should always be made with the point making contact with
the surface, as, on breaking the surface, surface tension causes pulling of the surface.

67
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

A precision micrometer can be set and read to about 1 jim for mercury and other liquids. Two
readings are required on each column for a pressure measurement, and the micrometer setting
and reading introduce the largest uncertainty in the use of these instruments. Moderate care
must be taken to ensure that the micrometer axes are vertical to eliminate the cosine error.
However, any change in level that occurs during a measurement introduces a sine error which
is equal to the product of the distance between the micrometer axes and the sine of the change
in angle.

3.10.2 Inclined-tube manometers

Inclined tube manometers, with a low density liquid, are the most usual instruments of this
form (Figure 3.13) and are used particularly for the industrial measurement of airflow. Fixed
cistern instruments are the most common, but because a pressure differential is measured the
temperature difference is small and the uncertainty associated with temperature measurement
may be negligible. Also a contracted scale is not necessary as this is equivalent to changing
the tilt of the tube.

The sensitivity of the inclined tube manometer can be calculated as follows:

sin d

3/ 1
Differentiating with respect to A.- (3.18)
3/? sin 6

where / is the length of liquid inside the tube along the scale
h is the vertical height of the liquid inside the tube
3 is the angle of the inclined tube against its horizontal

DRAIN PLUG

Figure 3.13: Inclined-tube manometer

From equation (3.18), when 3 is small, the tube is close to the horizontal position and the
sensitivity of the manometer may be high. Unless special care is taken, these instruments can
have a large sensitivity without the corresponding accuracy. Poor accuracy can be due to
surface tension effects, errors in the density of the liquid, drainage of the liquid down the tube,
tilt of the instrument, evaporation of the liquid, etc.

68
DB BROWSE, REVISED KEEN

Because the liquid used in these manometers is often not a pure substance, the density could
change from sample to sample. It is therefore important to use only liquid from the same
sample with which the manometer was calibrated.

3.10.3 Gas manometer

Although not an actual ‘instrument’, the gas manometer is a device for producing very small
high precision pressure differences for the calibration of differential pressure measuring
devices. It consists of two tubes open to the atmosphere and filled to overflowing with
different gases. The bottom end of the tubes are applied, one to each side, to a differential
pressure transducer. For example, if nitrogen and argon are used in tubes 2 m high then from
equation (3.11):
SP = P ^(M -M ■ )
RT nitrogen''

= 101325 x 2 x 9.8(0.040 - 0.028)7(8.314x293)


= 9.78 Pa.

Very few precautions are required for the pressure difference to have an uncertainty of less
than 1%, regardless of which gases are used. For another example of this technique see [7].

3.11 Properties of mercury


Because mercury is widely used, in mercury barometers, its properties have been studied
thoroughly. Density, compressibility and vapour pressure are the three main properties of
concern in pressure measurement.
(1) Density

The density of mercury over the range 0 to 40°C and 101.325 kPa may be calculated by means
of the following formula [8]:

d = 13595.08/(l+(l8150.36 t + 0.70209 + 2.8655x10'’ + 2.621x10’^ ?)xl0’^) (3.19)

where d is in kg/m’ and t is in °C. At 20 °C the accepted value is 13545.844 kg/m'\ while at
0 °C the value is 13595.080 kg/m^. The uncertainty in the density values given by equation
(3.19) is 3 parts in 10^. The accepted value of the density at 0 °C for defining the relationship
of pressure units is 13595.10 kg/m’.

(2) Compressibility

The compressibility factor, k, (i.e., the change in volume per unit volume) of mercury with
change in pressure is given by:

k = 4.039 X 10'^ P - 0.8 x 10 + 0.2 x 10'’* P' (3.20)

69
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

where P is the absolute pressure in bars (100 kPa). A change in pressure of one atmosphere
changes the density of mercury by 4 parts in 10^, and so can be ignored except for the most
precise work.

(3) Vapour Pressure

Mercury has a vapour pressure, P, in Pa given by [9]

log P = 10.1621 -3204/r for 285 < r< 326 K (3.21)

giving a value of 0.169 Pa (0.00126 mm Hg) at 293 K.

The vapour pressure of mercury causes it to spread throughout any system connected to a
mercury manometer and sometimes this causes undesirable contamination of the equipment. If
this is avoided by the use of an isolation differential pressure transducer between the two
systems then a correction for the vapour pressure may be needed.

Also, if the vapour pressure is significant compared to the pressure being measured, such as
measurement of absolute pressure at close to zero pressure using a mercury manometer, then a
correction should be applied.

3.12 Properties of water


Because the appropriate properties of water are known it is widely used as a fluid in
manometers. The main problem, however, is the surface tension, which is very variable unless
very clean apparatus is used.

(1) Density

The density of Standard Mean Ocean Water (SMOW) or pure air-free water over the
temperature range 0 °C to 40 °C at standard atmospheric pressure 101325 Pa is given in [10]:

. (t - 3.983035V(r + 301.797)
d = 999.97495 - (3.22)
522528.9(z +69.34881)

where d is the density in kg/m^, and t is in °C. The uncertainty in the density is ±0.00084
kg/m^up to 35 °C, increasing to ±0.0088 kg/m^ at 40 °C.

Dissolved air decreases the density of water from 0 °C to 25 °C by [ 11 ]:

Ad = (-4.612 + 0.106 r)xl0*-’ (3.23)

Most users use distilled tap water as a density standard, therefore, a correction, which depends
on the isotopic composition of water, may be applied to equation (3.22) [12]. However, if there

70
Erratum

Please change equation (3.22) on page 70 of Monograph 7 to:

(z - 3.983035)^ (f+ 301.797)


J =999.97495 (3.22)
522528.9(z +69.34881)
DB BROWSE, REVISED KEEN

is no measurement of the isotopic composition of the water sample, a value of 999.972 kg/m^
may be assumed for the first term in equation (3.22) for water distilled from tap water [12].

(2) Compressibility

The change in the density of water for a change in pressure is given by [10]

zk/ = (50.74- 0.3261 + 0.00416 r) X 10 “(P - 101325) (3.24)

where P is the pressure in pascals. This equation shows that a change of 2 kPa produces a
density change of 1 part in 10^ and thus for most purposes can be ignored at atmospheric
pressure.
(3) Surface tension
The surface tension of pure water is 72 mN/m. This, however, is rarely obtained and the value
for distilled water is closer to 55 mN/m. If a drop of detergent is added to the surface the value
drops to around 30 mN/m and this produces a more consistent value.

3.13 Other manometric liquids


Manometer fluids are usually chosen to have low vapour pressure. Coloured non-corrosive oils
with differing operating temperatures, flashing points and densities, ranging from 830 to 3000
kg/m^, are available from the manufacturers.

71
CHAPTER 3. BAROMETRY, MANOMETRY & COLUMNS

3.14 References
[ 1 ] BS2520:1983. British Standard Specification for Barometer conventions and tables, their
application and use. British Standards Institution 1983.

[2] F. Pavese, G. Molinar. Modern Gas-Based Temperature and Pressure Measurements


Chapter 7. The International Cryogenics Monograph Series Plenum, New York 1992

[3] Gould, F.A. and Vickers, T. (1952). J. Sci. Instr., 29, 85.

[4] ASTM D 3631 -99 Standard Methods for Measuring Surface Atmospheric Pressure.
American Society for Testing and Materials, Philadelphia, 1993.

[5] ISO Guide to the Expression of Uncertainty in Measurement. International Organization


for Standardization, Geneva, 1995. ISBN 92-67-10188-9.

[6] R.E. Bentley Uncertainty in Measurement'. The ISO guide. CSIRO, Sydney, Monogr. 1:
NML Technology Transfer Series, Edition 6, 2003.

[7] N Bignell Radial tensioning of diaphragms for micromanometers. J. Phys. E, 9 730, 1976

[8] A Cook. Precision Measurements of the Density of Mercury at 2O‘’C Philos. Trans. A
254: 125, 1959

[9] F.M. Ernsberger and H.W. Pitman, Rev. Sci. Instrum. 26: 584, 1955

[10] M. Tanaka, G. Girard, R. Davis, A.Peuto, N. Bignell. Recommended table for the density
of water between 0°C and 40°C based on recent experimental reports. Metrologia 38:
301-309, 2001

[11] N Bignell Metrologia \9: 51-59,

[12] P Chappuis, Trav. Mem. Bur. Int. Poids et Mesures, 1907 13, DI

Further reading
The following publications provide information on the use of barometers and manometers,
although not all of them will be readily available.

Guide to the Measurement of Pressure and Vacuum. National Physical Laboratory, Institute of
Measurement & Control, London, 1998.

Measurement of Atmospheric Pressure. Handbook of Meterological Instruments (2""^ Edition),


London, H.M.S.O., 1980.

72
Chapter 4

The gas-operated piston gauge and its use


John KL Man

4.1 Introduction
Piston gauges have been widely used as a fundamental method of pressure measurement
because they are robust, portable, convenient to use and are able to measure pressure over a
wide range. Basically, a piston gauge consists of a piston mounted vertically in a close-fitting
cylinder filled with a gas or hydraulic pressure fluid (Figure 4.1).

Figure 4.1: Basic features of a piston gauge

The piston is generally equipped with a weight-loading table on which weights of known mass
values can be placed. The pressure to be measured is applied to the base of the piston
generating an upward vertical force. This force is balanced by the downward gravitational
force generated by known masses acting on the effective area of the piston-cylinder
combination. The applied pressure is calculated by the definition of pressure as force per unit
area.

F _ mg
(4.1)
\ 4

73
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

where P is the applied pressure measured at the base of the piston (Pa), is the effective area
of the piston-cylinder combination (m^), m is the mass of the piston and associated masses (kg)
and g is the local value of gravitational acceleration (ms’^).
The value of pressure, P is actually the difference between the internal pressure at the base of
the piston and the reference pressure acting on the top of the piston and the weights. In
absolute pressure measurements, the reference pressure is the vacuum pressure and in gauge
pressure measurements, the reference pressure is atmospheric pressure. In some cases such as
engine manifold pressure measurements, gauge pressure is below atmospheric pressure. This is
sometimes known as negative gauge pressure.

It should be noted that equation (4.1), as it stands, does not take into account other factors due
to environmental and operating conditions which may affect the values of the quantities force
and area. This will be discussed in more detailed in section (4.3).

4.2 Gas-operated piston gauges


The pressure fluid used in these instruments is a gas, usually nitrogen. Piston gauges generally
consist of a base and several interchangeable piston and cylinder combinations, each covering
a part of the total range, and one or more sets of weights. Some makers supply instruments
with automated pressure generator/controllers or fitted with hand pumps to generate and
control the pressure. Additional output pressure ports may also be supplied for calibration of
test instruments such as pressure gauges or pressure transducers. A complete instrument of this
industrial type is often known as a dead-weight tester. A basic construction of a gas-operated
dead-weight tester for calibration of bourdon gauges is shown in Figure 4.2.

(Volume Changer)

Figure 4.2: Bourdon tube gauge calibration using a dead-weight tester

74
JOHN MAN

There are a variety of commercial gas-operated piston gauges available covering the pressure
ranges from about 2 kPa up to 11 MPa. which is sufficient to calibrate most of the gas-
operated pressure instruments used for industrial applications. Some piston gauges have the
ability to measure both gauge and absolute pressure. These instruments are equipped with a
bell-jar to form an enclosure surrounding the piston and weights (Figure 4.3). The air inside
the bell-jar can be evacuated to allow absolute measurements (Figure 4.4). Some of these
systems have a remote weight changing mechanism such that the vacuum may be maintained
throughout the calibration.
Bell Jar (optional)

—- Pressure Fluid
Figure 4.3: Gauge pressure measurement

Bell Jar

Figure 4.4: Absolute pressure measurement

The uncertainties associated with gas-operated piston gauges are typically about 10 to 50 ppm
for the best of these instruments operating in a suitable environment. For lower precision
industrial instruments, the uncertainties are in the order of 100 to 200 ppm.

Rotating either the piston or the cylinder generates a relative circumferential motion and drags
gas between the surfaces, thus avoiding contact between the piston and cylinder and
eliminating indeterminate forces. Most modern instruments have a mechanical drive with a
motor to initiate and maintain rotation but such devices may introduce extraneous vertical

75
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

components of the driving force. This effect is more significant at the low end of the range of
the balance. In addition, the heat generated by the motor may also affect the measurement. For
these reasons, it is preferable to initiate the rotation manually. However, it is not practical to do
this when the piston gauge is mounted in a vacuum chamber for absolute pressure
measurement. In this case, it is best to use the mechanical drive to establish the rotation and to
disconnect it before the measurement.

In negative gauge pressure measurements, the piston is mounted in such a way that the weight­
loading table is located at the bottom of the piston (Figure 4.5) whilst the piston and cylinder
combination is positioned at the top. The pressure applied to the piston and cylinder
combination is less than atmospheric pressure and so an upward force is generated. This
upward force is balanced by the downward gravitational force on the masses.

Pressure Fluid

Figure 4.5: Negative pressure measurement

It is a common practice to calibrate high pressure gas instruments using hydraulically operated
piston gauges with oil/gas separators. The use of oil/gas separators will introduce uncertainty
to the measurement and may contaminate the instrument under test. Special designs of high
pressure gas-operated, gas-lubricated piston gauges with the capacity to operate up to 100 MPa
are available to calibrate instruments, where cleanness is paramount for safety purposes such
as for oxygen service.

The recent advance of technology in the fabrication of large diameter piston and cylinder units
with high precision geometries has made gas-operated piston gauges the alternative to mercury
manometers as primary pressure standards. For example, the establishment of a 50 mm
diameter piston-cylinder unit as a primary pressure standard at about atmospheric pressure has
been reported recently in the literature [1].

4.2.1 Variant designs of gas-operated piston gauges

(1) Ball and nozzle gauges

The ball and nozzle gauges are used to generate a wide range of pressures typically from 1 kPa
to( 10 MPa. These instruments have an inlet for connection to a pressure supply source, an

76
JOHN MAN

outlet and a vertical nozzle into which a ceramic ball fits (Figure 4.6). A weight carrier sits on
top of the ball. The inlet is equipped with a flow regulator that introduces pressure underneath
the ball, lifting the ball and the weight carrier slightly until equilibrium is reached. At this
balanced point, the ball and the weights are supported by the pressure, P, below the ball. The
pressure P is calculated from the downward gravitational forces generated by the weights
acting on the known effective area of the ball-nozzle combination using the form of Equation
(4.1).

Figure 4.6: Ball and nozzle gauges

The vented gas passes through a small hole under the ball and flows through a coil of tubes for
stabilising the flow. The output pressure, which is the same as the pressure P, is connected to
an instrument under test. During operation the ball is floating and is supported by a dynamic
film of gas, it does not need to be rotated to eliminate contact between the ball and the nozzle.
This type of instrument is often used in applications where a constant pressure needs to be
maintained for an extended period unattended. The ball and nozzle gauge is strictly not a
piston gauge. The effective area and its performance depend on the flow rate, which is much
higher than a normal gas-operated piston gauge. Because of the narrow gap between the ball
and the nozzle that controls the flow and thus the pressure developed. They are more
susceptible to vertical bumps than piston gauges.

(2) Digital piston manometers

This type of pressure instrument measures the force produced by the piston using an electronic
balance (functioning as a dynamometer) with a digital readout enabling a range of pressures to
be measured. The basic features of this force-balance piston gauge are shown in Figure 4.7.

It consists of a measuring block mounted onto a dynamometer. The measuring block has two
main components, a piston-cylinder assembly and a cap with a port for the connection of
pressure tube. The measuring block is interchangeable giving a wide working pressure range
from a few Pa up to 25 MPa. For higher pressure operation, the measuring block is designed to
be used with gas or oil. A motor drive is supplied to rotate the piston. The mounting block has

77
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

an inserted temperature probe to measure the temperature of the piston and cylinder assembly.
The piston and cylinder assembly has a coupling head with a ball bearing to transmit the force
generated by the applied pressure to the dynamometer with minimum friction. A set of masses
is provided to set the span of the dynamometer using a built-in calibration routine, and to
check the linearity of the readings of the dynamometer. The resolution of this type of
instrument is stated as 1 or 2 ppm of the full scale value with typical uncertainty value of 50
ppm to 100 ppm.

Figure 4.7: Digital piston manometer

(3) Taper weight gauges

These instruments use taper weights instead of a piston for generating relatively small gauge
and differential pressures, typically between 20 Pa to 16 kPa for gauge pressure and between 5
Pa to 16 kPa for differential pressure. Typical uncertainty associated with this type of
instrument, as claimed by the manufacturer, is 0.1 Pa below 500 Pa and 0.02% of reading from
500 Pa up to 16 kPa. Figure 4.8 shows a taper weight fitted into a corresponding cylinder. The
pressure supply source is connected to a cascade of flow regulators and a volume for
controlling and stabilising the inlet gas flow. The pressure introduced underneath the taper
weight through an inlet aperture lifts the taper weight slightly. At equilibrium, the pressure, P,
supports the taper weight. The effective area of the taper weight-cylinder combination is
calculated using the general form of Equation (4.1). The vented gas passes through an outlet
aperture underneath the taper weight and stabilised by means of a volume. The output
pressure, P is connected to the test instrument. The operating principle of the taper weight
gauge is similar to the ball-nozzle gauge. Strictly speaking, these instruments are not piston
gauges because their effective area is not so clearly defined and can change with flow rate. The
taper weights are interchangeable each having a unique mass for generating a specific pressure
over the operating range.

(4) Large area non-rotating piston gauges

Another design uses a large diameter piston and cylinder unit, made of Invar, with an effective
area of 100 cm" to generate small gauge and differential pressures, typically 1 Pa to 3 kPa. A
schematic diagram of the instrument is shown in Figure 4.9.

78
JOHN MAN

Figure 4.8: Taper weight gauge

Figure 4.9: Large area non-rotating piston gauge

There is a mechanism using two levers, one at each end of the piston. One end of each lever is
connected to the piston and the other to the cylinder, each by a flexible hinge. This holds the
piston centrally in the cylinder with a small gap to avoid friction between the piston and
cylinder without the necessity of rotation of the piston, while allowing axial movement. A
steady gas flow from a flow controller is introduced to the inlet port applying a pressure to the
top of the piston, P. The piston-cylinder assembly is rigidly mounted on an electronic balance,
which gives a direct read out of the downward force generated by the pressure, P. Two annular
grooves are machined in the cylinder wall, one near the top and one near the base which has
holes for venting the flow to the exit port (reference pressure Pre/)- This allows the gas to flow

79
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

between the two grooves with an initial flow to the top of the piston until a steady pressure is
established. The effective area of the piston and cylinder can be calculated by dimensional data
of the piston and cylinder or by calibration against a pressure standard. Standard weights are
used to calibrate the balance.

4.3 Evaluation of pressure


Equation 4.1 is a simplified mathematical model for the evaluation of the pressure measured
by a piston gauge. In practice, it is required to take into account correction factors for the
downward force and the effective area to achieve measurements of high accuracy. In addition,
further corrections may be required to the final values of the pressure.

4.3.1 Force

Mass and air buoyancy effects

The total mass m in equation (4.1) is the summation of the mass of the piston, the mass of the
weights on the piston and any trim weights which may have been used to achieve the pressure.
In mathematical notation, it is written as^m, . The weights are normally in disc form
supplied by the maker of the instrument. For high quality instruments, the weights are made of
non-magnetic stainless steel. Some instruments have a conversion ring or weights to convert
the pressure values of the instrument from imperial to SI units. The weights are usually
stamped with their nominal pressure values, based on a given value of gravity.

In gauge pressure measurements, all parts making up considered to be


submerged in air and therefore displace a volume of air equal to their own volume. By
Archimedes’ Principle, this produces an upthrust which is proportional to the mass of the
volume of air displaced. It can be shown that the upthrust is

717
(4.2)
' P mi

where: = ambient air density at the time of pressure measurement, kgm '^
= density of each weight that made up the total mass, kgm"\

The net downward force, F, on the total mass m. is calculated as

—)• (4.3)
' P mi

p
The term (1----- —) is often referred to as the air buoyancy correction factor, C/,.
Pmi

80
JOHN MAN

(a) Conventional mass

The mass of each of the weights of a weight set, and the mass of the piston are generally
calibrated by weighing in air against standard weights of known density and mass. When a test
weight, nij is balanced by a standard weight, the equation relating the equilibrium forces is

'Wr (1 - —)g = (4.4)


Pm Ps

where p^ is the density of the standard weight and p^' is the air density at the time of the mass
measurement. As it is not always practical to measure the densities required for the buoyancy
correction factors, a convention is widely adopted throughout the world to assign the mass
value of a weight in terms of , which is based on the values of p^ and p^' to be 8000
kgm’^ and 1.2 kgm'^ respectively. Using this convention, the mass value assigned to the test
weight is called the conventional mass, M. The relative error, e, in using the conventional mass
instead of m-j-, referred to as the true mass, can be derived from equation (4.4) as

(4.5)

It is seen that the relative error is zero when the density of the test weight is equal to that of the
standard weight. In this case, the true mass is equal to the conventional mass.

In gauge pressure measurements, where the accuracy requirement is not high, it is adequate to
use the conventional mass values with the assumptions that the density of the weights, is
8000 kgm '^ and air density at the time of the pressure measurements, p^^ is 1.2 kgm'^. Using
these values, the buoyancy correction factor, Q is equal to 0.99985. The relative error arising
in the calculation of the downward force F using the conventional mass values with the
assumed values of air density of 1.2 kgm’^ and density of weights of 8000 kgm'^ instead of
using the actual values p^^ and p^^, can be shown (to a good approximation) to be

= — (A,-1-2) (4.6)
Pm

A guide to the magnitude of the relative errors calculated using equation (4.6) for typical
materials used in piston gauges is presented in Table 4.1.

When the air density is measured but the density of the weights is assumed to be 8000
kgm'\ the relative error incurred in the calculation of the downward force using the
conventional mass values when the air density ' during the mass measurement was taken as
1.2 kgm '^ is
^ = (/7,-1.2)(-------------- (4.7)
A. 8000

81
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Table 4.1 Relative error (in ppm) in calculating the downward force using the
conventional mass and assuming = 8000 kgm'^ and = 1.2 kgm'^

Actual Actual Air density, p^ kgm’^


Materials
1.1 1.15 1.2 1.25 1.3 0
Aluminium
2800 -36 -18 0 18 36 -429
alloys
7000 -14 -7 0 7 14^^"^^ -171
Cast iron/
7500 -13 -7 0 7 13 -160
Stainless
7800 -13 -6 0 6 i
steel alloys
8000 -13 -6 0 6 13 -150
Brass 8400 -12 -6 ® 0 6 ililBIII -143
Tungsten
14000 -7 -4 0 4 7 -86
carbide

The largest e will arise when a piston gauge operates in vacuum, where the air density is
very close to zero. When the actual value of p^ is equal to 1.2 kgm'^ or p^^ is equal to 8000
kgm''\ the error is zero. A guide to show the magnitude of the relative errors calculated using
equation (4.7) for some typical materials used in piston gauges is presented in Table 4.2.

Table 4.2 Relative error (ppm) in calculating the downward force using the conventional
mass and assuming = 8000 kgm’^

Actual Actual Air density, p^^ kgm’^


Materials /7,„(kgm-’) 1.1 1.15 1.2 1.25 1.3 0
Aluminium
2800 -23 -12 0 12 23 -279
alloys
7000 -2 -1 0 2 ...-21
Cast iron/
7500 -1 -0.4 0 0.4 1 -10
Stainless
steel alloys
7800 -0.3 -0.2 0 0.2»b 0.3 i -4
8000 0 0.0 0 0 0 0
Brass 8400 J 0 0
Tungsten
14000 5 3 0 -3 -5 64
carbide

In general, the relative errors shown in Table 4.2 are smaller than those indicated by Table 4.1.
As a result, the accuracy of gauge pressure measurements can be improved if the variation of
the air pressure density is allowed for during pressure measurements. The density of ambient
air depends on the ambient air pressure, temperature and relative humidity as well as its
composition. Values of these parameters can be used to estimate the air density using the
expression recommended by the International Bureau of Weights and Measures [2].

82
JOHN MAN

(b) True mass

It is clearly shown from the aforementioned error analysis that actual values of the densities of
the component parts of the load would have to be used for very high precision work, specially
in the case of absolute pressure measurement, where weights are to be used in vacuum. In
these cases, it is necessary to use the true mass instead of conventional mass to minimise the
errors in the calculation of the load.

The true mass value m of a weight is related to the conventional mass value M, based on values
of air density of 1.2 kgm'^ and density of the weight of 8000 kgm’^ as follows

1
m = M 1 + 1.2 (4.8)
8000^

It should be noted that the treatment of the air buoyancy effect as mentioned in this section
applies to all parts of the piston. In fact, the upper part of the piston is immersed in ambient air
and the lower part immersed in the pressure fluid. For convenience, it is general practice to
apply the air buoyancy effect to all parts of the piston and then make corrections for the
buoyancy effect acting on the part of piston that is submerged in the pressure fluid. However,
for a piston with simple cylindrical form, it is not necessary to take account of these additional
buoyancy corrections [3]. For pistons with irregular shape, the piston mass will require other
buoyancy corrections, as discussed in the next section.

Fluid buoyancy Effects

In some designs the piston is of irregular shape, e.g., in gas-operated instruments, the piston
may have a hollow section of w, volume (Fig. 4.10) to reduce the mass of the piston so that a
low operating pressure can be achieved. In this case, the volume of the pressure fluid displaced
by the hollow piston is Dj smaller than the volume displaced by a simple cylindrical piston.

The fluid volume will cause a downward force , which is equivalent to the weight of the
pressure fluid

F, (4.9)

where is the density of the pressure fluid. If the pressure fluid is a gas, then will change
with pressure and temperature (see section 4.3.3 fluid head correction).

As the whole of the piston was included in the air buoyancy correction as shown in the
previous section, the air buoyancy acting on the volume y,, now subjected to fluid buoyancy
has to be subtracted. The net fluid buoyancy correction term, becomes;

83
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Equivalent
Reference Level
for a
Simple Piston

Figure 4.10: A hollow piston

Fc -Pa")8 (4.10)

In some other designs the piston may have protuberances of volume as shown in Figure
4.11, in which case the fluid buoyancy correction will be the additional upward force
equivalent to the weight of the volume, 1)2 of the pressure fluid outside the volume defined by
the simple piston.

Equivalent
Reference Level
for a
Simple Piston

Figure 4.11: A piston with protuberances

The fluid buoyancy correction term is

Fc =-01(P, - Pa}8 (4.11)

Combining equations (4.10) and (4.11) for a general case of pistons with irregular shape

- Pa )g (4.12)

84
JOHN MAN

Adding the fluid buoyancy correction term to equation (4.4), the downward force acting on the
piston of an irregular shape is

F= )+ -pjg (4.13)
' P mi

Gravitational acceleration

As seen in equation (4.13), the total downward force from known masses acting on the
effective area of the piston-cylinder combination is calculated by multiplying the total mass of
the load, corrected for air and fluid buoyancy effects, by the local value of the acceleration due
to gravity, g. The local value of g may be derived by a number of methods, the choice of which
depends on the accuracy required. For normal industrial operations, an accuracy in the pressure
determination of ± 0.05% is expected. That means it is probably sufficient for g to be known to
within ± 0.01%. The local value of g depends on the geographical locality and altitude and
may be calculated to generally better than ± 0.005% [4] from

S - Se( 1 + P\ sin2 (p-P2 sin 2 2^)- 3.088v 10"^// (4.14)

where: g = value of the acceleration due to gravity for latitude (/> and height H metres above
sea level, ms’^
ge = value of the acceleration due to gravity at the equator and at sea level, ms';

The values of the constants ge, and recommended by the International Union of Geodesy
and Geophysics are:

ge = m-s'*
P^ = 0.0053024
P2 = 0.0000058.

If a higher accuracy of pressure measurement is required, then a better estimate of the value of
g than those obtained from equation (4.14) can be calculated based on a published value of g at
a site as near as possible to the instrument. The local values of acceleration due to gravity at
sites throughout Australia may be obtained from the Geoscience Australia (web site address
www.ga.gov.au/map).

For the highest possible accuracy, the local value of g is determined by measurements using
gravity meters. The gravity value at the pressure laboratory, NML has been measured using an
absolute gravimeter based on timing a falling object. The uncertainty associated with this
gravity value is estimated to be ±0.3 ppm.

In some cases, the calibration results of a piston gauge (or a pressure instrument that uses
weights to generate the downward force) are reported in terms of pressure values, generated by
the piston gauge at a site with a particular value of gravity gi. If the piston gauge is to be used
at a site with a different value of gravity, g2 then the reported pressure values have to be
corrected by multiplying them with the factor {g2lgp-

85
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Force due to surface tension

The surface tension of a liquid acting on the piston at the point where it emerges from the
liquid will generate a downward force. Correction for force due to surface tension applies only
to hydraulically-operated piston gauges. (Refer to Chapter 5 for details of the correction term
for surface tension.)

Effective mass, M

There are some other sources of error involved in the operation of piston gauges which may be
difficult to assess accurately. These arise from the fluid buoyancy correction, or, in the case of
hydraulic operations from oil that may collect on the floating element and add to its mass. Air
drafts and other aerodynamic forces and stray magnetic or electrostatic effects can change the
load, the speed and the direction of rotation can also alter the pressure produced. For these
reasons, NML introduces a correction term known as effective mass, Mwhich includes all
these effects and any other indeterminate vertical forces that may arise during operation. The
value of Mis established by experimental and statistical procedures. The effective mass
includes the air buoyancy effect. Adding Mto equation (4.13), the total force is

f= -pjg (4.15)
' P mi

4.3.2 Effective area

Effective area at reference pressure and temperature

The area over which the pressure is considered to act is usually known as the effective area, Ag
and has a value somewhere between the cross-sectional area of the piston and that of the
cylinder. The value of the effective area of piston-cylinder combination can be greatly
influenced by pressure and temperature. In practice, the effective area is referred to standard
conditions with respect to pressure and temperature. The standard conditions are taken as
atmospheric pressure and 20 °C (some laboratories choose 23 °C as their reference
temperature). The effective area under these conditions is denoted by the symbol A0.20 and its
value is normally determined by direct comparison with a pressure standard. It is also possible
to determine the value of Ao,2o by dimensional measurements a procedure used for
establishment of primary standards. The dimensional measurements include diameter,
roundness and straightness at intervals over the length of the piston and internal surface of the
cylinder. This information is then used to give the geometrical shape of the piston-cylinder
combination for calculation of Ao,2o [3]- Determination of Ao,2o by dimensional measurements is
beyond the scope of this discussion.

Changes in effective area due to pressure

As pressure applied to the piston-cylinder combination increases, the effective area changes
due to the elastic deformation of the piston and cylinder materials. At a given pressure, the

86
JOHN MAN

amount of elastic distortion mainly depends on the Young’s modulus, E of the piston and
cylinder materials. Tungsten carbide has a Young’s modulus of about three times that of
stainless steel. This means that the influence of the distortion of the piston-cylinder
combination is reduced when tungsten carbide is used instead of stainless steel. This is one of
the main reasons that most laboratories use pistons and cylinders made of tungsten carbide for
high precision pressure measurements.

The amount of distortion also depends on the design of the piston and cylinder. In the case of a
simple piston-cylinder design, the distortion increases as the pressure increases, resulting in
excessive fluid leakage past the piston at high pressure. One design to minimise this problem is
the re-entrant type in which the pressure fluid acts on the external surface of the cylinder as
well as the base of the piston and along the gap between the piston and cylinder. This external
pressure may cause the internal diameter of the cylinder to close up around the piston with
increasing pressure resulting in a negligible increase or even a decrease in the effective area.
Another design is to apply an auxiliary pressure to the cylinder’s outer surface to control its
distortion. This is a controlled clearance piston gauge and can be used to very high pressures of
up to 1.4 GPa. Types of piston and cylinder units are shown in Figure 4.12.

Re-entrant Controlled Clearance


Figure 4.12: Types of piston and cylinder units

The effect of pressure on the effective area is characterised by a distortion coefficient, A. It is


generally assumed the effective area varies linearly with the applied pressure for simple and
highly regular piston-cylinder units [3]. The equation for calculating the effective area at an
applied pressure P and at the reference temperature of 20 °C, is
Ap.2o = A).2oO + >^^) (4.16)

where A = pressure distortion coefficient. Pa '.

The value of A is normally determined experimentally by direct comparison with a pressure


standard. There are several analytical and numerical models, based on elastic theories and an
assumed knowledge of the pressure profile in the clearance between the piston and cylinder,
for calculating A [5]. However, these models have limitations as some of the basic
assumptions from which the models are derived are questionable.

87
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Changes in effective area due to temperature

As the temperature of the piston-cylinder unit deviates from the reference temperature usually
taken as 20 °C, the effective area will change with a rate depending on the linear thermal
expansion coefficient of the piston and cylinder materials. The equation to calculate the
effective area at a temperature t and at the reference pressure of atmospheric pressure is

^0., = >^0.20 [1 + («p + )(? - 20)] (4.17)

where: 6^, = linear thermal expansion coefficient of the piston, °C *


= linear thermal expansion coefficient of the cylinder, °C '
t = temperature of the piston-cylinder assembly during its operation, °C.

The values of 6^, and a, are often supplied by the manufacturers. As an example, the linear
thermal expansion coefficient of tungsten carbide (with 6% cobalt) and of stainless steel is
typically quoted as 4.5 x 10'^ °C*’ and 11 x 10'^ °C*’ respectively. Measurement of temperature
r of a piston-cylinder unit should be carried out by placing a miniature platinum resistance
thermometer close to the piston-cylinder unit. For gas-operated piston gauges, it is advisable to
install temperature sensors in the pressurising gas to detect temperature changes due to
adiabatic expansion and compression of the gas during operation. This will give an indication
of the time required for the piston-cylinder unit to reach thermal equilibrium.

Effective area as a function of pressure and temperature

The expression for the effective area, taking into account the corrections due to pressure and
temperature is given by combining equations (4.16) and (4.17) to give

- ^0,20 “ 20)] (4.18)

where 4^, , is the effective area of the piston-cylinder combination at the applied pressure P
and at temperature t.

4.3.3 Fluid head corrections

In the forgoing, the pressure P is calculated at the base of the piston. This then becomes the
reference level. If the pressure is required to be known at a different level, the head of pressure
due to the difference in levels, P/, has to be taken into account. Its value is:

Ph = ^Pf - (4.19)

where: P,^ = pressure due to fluid head. Pa


p f = the density of the fluid at pressure P, kgm'^

88
JOHN MAN

h = height difference between the base of the piston and the selected level, to be
taken as positive if the selected level is below the reference level of the piston
gauge and taken as negative if above, m.

In the case of an hydraulic instrument, the value of may be considered constant for a given
temperature. The value of for a gas-operated instrument will vary significantly with
pressure and also with temperature. As a good approximation, the density of most gases at
pressure, P and temperature, t may be found from

__2Z_ (4.20)
(273.15 + 0

where = 3.3694 x 10'^ kg m’'^ KTPa for nitrogen gas and 3.4837 x 10'^ for air, when t is in
°C and P is the absolute pressure in Pa, giving the density in kg m "^.

4.3.4 Final expression for the pressure

Taking all the corrections in equations (4.15) and (4.18) into account, a general model of
pressure measured by a gas-operated piston gauge in gauge mode at the base of the piston is
given by

£/71.g(l )-h -t-fu, - Pjg


P = ------------- ------------------------------------------- (4.21)
Aq.20 (1 + '^)[1 + (CCp + - 20)]
where:
P = gauge pressure at the base of the piston. Pa
nij = total true mass value of the load, including the floating element with its attachment, kg
i
M = effective mass, kg
Pa = ambient air density during operation, kgm’^
p,ni = density of each weight, kgm'^
g = local gravitational acceleration, ms '
^0,20 = effective area of the piston/cylinder unit at atmospheric pressure and at the reference
temperature of 20 °C, m"
ex,, = linear thermal expansion coefficient of the piston, °C '
ttc = linear thermal expansion coefficient of the cylinder, °C '
t = temperature of the piston-cylinder assembly during its operation, °C.

For most operations in gauge-mode, where highest accuracy is not required, it is normal to use
the conventional mass values taking the density of the weights as 8000 kgm'‘\ Furthermore, it
is not necessary to take extra effort to measure the mass of the piston and to calculate the
buoyancy effects on the piston. The mass of the piston, its buoyancy effects and the fluid

89
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

buoyancy correction term may be incorporated into the term M^. Equation (4.21) can be
simplified to

i 8000 (4.22)
^0,20 ~

where: = total conventional mass value of the load acting on the piston, kg.

M = effective mass of the piston, kg


Pa = ambient air density during operation, taken as 1.2 kgm’ .

For gas-operated piston gauges in absolute mode, the air buoyancy effect on the masses is
negligible. The pressure applied at the base of the piston is expressed as

(4.23)
" ■'■.2. (1 + )[l + («,, + - 20)] “
where: Pq = the vacuum reference pressure. Pa
M' = effective mass without air buoyancy, kg.

The reference pressure in the bell jar is generally measured by Pirani or capacitance diaphragm
gauges. It is desirable to operate the absolute measurement at a very low so that the
uncertainty of Pq has an insignificant contribution to the measurement.

Incorporating the true mass of the piston and fluid buoyancy effects into A/Equation (4.23)
can be simplified to

(4.24)
/lo.2,(l + /iP„j,)[l + (a^+aJ(z-20)] “

where M is the effective mass of the piston (in vacuum), kg.

Note that equations (4.21) to (4.24) contain P on both sides so that a strict evaluation requires
an iterative calculation, that is starting with a rough value of P on the RHS a better value is
obtained form the equation. This can then put in the RHS and the calculation repeated. In
practice this process will very rapidly converge (or stop making any difference).

The pressure at the selected level with a height difference of h from the base of the piston, ,
is calculated by applying the fluid head correction given in equation 4.19 to the pressure
measured at the base of the piston.

90
JOHN MAN

PsL=P + <P,f-Pa^gh (4.25)

4.4 Calibration of gas-operated piston gauges


The purpose of the calibration of a piston gauge is to determine the values of the effective area
A^, the distortion coefficient, z and the effective mass of the piston as expressed in
equations (4.22) and (4.24).

The usual method used to determine these three parameters is the cross-float method. In this
method, the piston gauge under test is connected to a reference pressure standard with an
isolation valve, usually a constant volume valve, placed in the pressure line between the two
assemblies. The two instruments are pressurised, either using individual or a common pressure
generating system such as a screw press or compressed gas in the case of gas operated
instruments. With the isolation valve opened, both instruments are then brought to a balance
condition by adjusting the weights on either side with small trim weights of known mass
values. The applied pressure in the cross-connected system can be calculated from the known
constants of the reference standard. The downward force on the test instrument at the balance
condition is known from the mass values of the applied weights and the local value of g, and
hence the effective area of the instrument can be determined at that pressure. This process is
then repeated for pressure values covering the entire range of the instrument. The cross-float
method may be used for both hydraulic and gas-operated piston gauges including variant
designs such as ball and nozzle gauges and the digital piston manometers. There are two
methods commonly used to determine the balance conditions in the cross-float method. These
are briefly described below.

4.4.1 Rate of fall method

A typical arrangement for the cross-float by rate-of-fall method is shown in Figure 4.13.

With the constant volume valve (CV) closed, both piston gauges are loaded with approximate
weights for the desired nominal pressure. The gauges are then pressurised so that they float at
their respective reference levels and with the weights rotated. The pistons will fall due to
leakage of the pressure fluid through the annulus between the piston and cylinder. A position
sensor such as an inductive type is mounted close to the weights of the standard instrument.
The natural fall rate of the standard piston, due to leakage between piston and cylinder, is
measured by the time taken for the piston to fall through a given distance. The CV is then
opened, and the fall rate again measured. An increase in fall rate indicates that the pressure
developed by the standard is too high, or, the downward force on the test instrument is too
small and conversely. Adjustment of the weights on either side is made with small trim
weights until both fall rates return to their isolated values.

This method is of high inherent accuracy, but is time-consuming. It is usually employed where
the calibration of a laboratory standard is involved or for intercomparions between standards.

91
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Figure 4.13: Cross-float by rate of fall method

4.4.2 Differential pressure method

A differential pressure cell equipped with constant volume by-pass valves is an alternative
method to determine the balance condition when cross-floating two piston gauges as shown in
Figure 4.14.

Figure 4.14: Cross-float by differential method

This method is generally slightly inferior to that of the rate-of-fall method, it provides an easy,
much less time consuming way to calibrate piston gauges. The standard piston gauge and the
test (X) piston gauge are connected via two constant volume valves. Both instruments are
pressurised for nominally equal pressures with CVl closed and CV2 open. The differential
pressure cell will have a zero pressure difference across the diaphragm in this state, allowing
the indicator to be set to zero. CV2 is closed, and CVl opened. If the pressures due to the
standard and instrument under test are different, the indicator will not read zero. Adjustments
to the pressure of either are made with the aid of small trim weights until the measured

92
JOHN MAN

pressure difference is zero. Since the instrumental constants of the standard piston gauge are
known, together with the downward force, the calibration pressure, P, in the system can be
computed by equation (4.22) for gauge pressure measurements.

4.4.3 Data evaluation

Cross-floating at several different pressures over the range of interest will yield several pairs of
data. Each pair consisting of a known developed by the standard at the reference level of
the test instrument and M,, the mass acting on the test instrument. For each data pair the
following equilibrium equation may be written

(4.25)
A)0 + ^)C,
where:
C, = temperature correction term, [ 1 •+• («p + aA -20)]
Ch = air buoyancy correction factor, (1 - -^^).
Pnu

Equation (4.25) may be re-arranged as

+ A,C,^P^ - M^g = M,gC,

A least-squares solution yields Ao, Mg and /. for the test instrument as well as uncertainty
values associated with each of these parameters.

4.4.4 Uncertainty evaluation

The uncertainty associated with the pressure generated by gas-operated piston gauges can be
estimated using the appropriate mathematical model. As already discussed in section 4.3, the
pressure P to be measured can be expressed in different ways that depend on the measurement
mode (absolute or gauge). In the case of a gas pressure measurement in gauge mode, P may be
expressed by equation (4.22). The use of this equation implies that the uncertainty of the
pressure value is a combination of the uncertainties of each of the input quantities involved in
the pressure calculations. The uncertainty components used in the evaluation of the overall
uncertainty of P are due to:

(i) Fitted parameters from the calibration of the instrument: Me, Ao and 2

(ii) Instrument parameters: mass (AZ); thermal expansion coefficient of piston and cylinder
combination (a,,, aA density of piston, (/?,,)
(iii) Pressure transmitting fluid: density {pf)
(iv) Environmental quantity: air density (p„)
(v) Actual operational parameters: temperature (r); head correction (P/,); local gravity (g)

93
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

An example of an uncertainty estimation of the pressure generated by a gas-operated piston


gauge is given in section 4.7.1.

4.4.5 NML Reports

NML have different calibration procedures and reports for piston gauges, depending upon the
quality of the gauge under test and its application. There are basically two classes of
calibrations offered for gas-operated piston gauges. They are designed to cater for users who
require the high-precision calibration for laboratory instruments or first-line industrial
standards, and those who require an industrial instrument capable of calibrating test gauges.

(1) Laboratory instruments

Calibration of a piston gauge involves the calibration of the weights and the piston-cylinder
combination. Weights are generally calibrated by weighing in air against NML mass standards.
Calibration of the piston gauge involves the experimental determination of the instrumental
constants: effective area at atmospheric pressure and at reference temperature (A^), distortion
coefficient (z) and the effective mass of the piston assembly {Mg). This is achieved by cross­
floating the test instrument against a pressure standard of known instrumental constants over a
range of calibration pressures. A formula incorporating the instrumental constants and all the
corrections is supplied, enabling the user to calculate the pressures generated by the
instrument. Values of the three constants, the uncertainty value associated with each of these
constants and the expanded uncertainty of the pressure generated by the instrument are
presented in the calibration report. A sample calibration report on a laboratory instrument is
presented in the section 4.8.1.

(2) Industrial instruments

This procedure is intended for industrial piston gauges (dead-weight testers), which are used
for the calibration of test gauges and other pressure measurement instruments that require the
applied pressures with an expected uncertainty of no more than ±0.05%. The calibration
procedure is basically the same as the procedure for laboratory instruments, but uses fewer
calibration pressures (usually 8) to determine the instrumental constants. The report gives
values for the effective area over the pressure range of the instrument (A^.) and the effective
mass of the piston assembly (Mg) together with a table showing the pressure generated by each
weight at a particular stated value of g and the pressure range of the instrument where the
uncertainty associated with the applied pressure is less than or equal to ±0.05%. A simplified
equation, based on the equation (4.22) is presented which may be used to calculate the pressure
values of the weights for different values of the acceleration due to gravity, g. A typical NML
calibration report on an industrial piston gauge is presented in section 4.8.1.

4.5 Principal uses of piston gauges


Pistons gauges of the industrial type, often known as dead-weight testers are mostly used to
calibrate pressure gauges. The dead-weight testers can also be conveniently used to calibrate
pressure transducers with an electrical output signal. Precision gas-operated piston gauges can

94
JOHN MAN

be used to calibrate other higher accuracy pressure instruments, such as dead-weight testers,
ball-nozzle gauges and the digital piston gauges, as mentioned in section 4,2.

4.5.1 Calibration of pressure gauges

A pressure gauge is one of the most convenient instruments for measuring pressure and
vacuum. It is relatively robust and is easily installed. A very common type of pressure gauge
is the Bourdon gauge. It consists of a curved shaped metal tube of an elliptical cross section
called a Bourdon tube, which is closed and free at one end and connected to a pressure source
at the other. When pressure is applied, the elliptical cross section becomes more circular
causing a deflection of the closed, free end of the tube. The magnitude of the deflection is
proportional to the magnitude of the pressure difference between the inside and the outside of
the tube. A mechanical linkage is used to convert the tube displacement into a rotation of a
pointer on a dial with suitable gearing or levers. The range specified by the manufacturer is
usually the range in which the pressure is related linearly to the pointer rotation. Several
variations of Bourdon gauges exist, such as the C shaped tube, the twisted helical and spiral
tubes, depending on the required sensitivity. A variety of materials such as phosphor-bronze
and stainless steel may be used for the tube, depending upon the application for which the
gauge is intended. Low pressure gauges often use metal diaphragms, capsules or bellows.
Quartz elements are also used for high accuracy gauges.

Calibration procedures vary, depending on the regulating authority, and can be found in the
appropriate codes or standards. The calibration procedures given in AS 1349-1986 are widely
adopted in Australia for Bourdon tube gauges. In general, a standard piston gauge is used to
apply reference pressures to the pressure gauge at several values within the range of the gauge.
By determining any errors which may exist corrections can be assigned to the gauge readings.
The uncertainties associated with the corrections may be estimated as shown in section 4.8.2.

4.5.2 Calibration of pressure transducers

Various types of pressure transducers and their characteristics are presented in Chapter 6.
There is no Australian Standard available for calibration of pressure transducers but they may
be calibrated according to the procedure used for Bourdon tube test gauges. A calibration cycle
consisting of a series of increasing and then decreasing pressures is applied by a pressure
standard to the test instrument. The first point and final point are normally at the zero applied
pressure. The zero may be adjusted at the beginning of the test. The number of calibration
points should not be less than 10 and these pressures should distribute as uniformly as possible
over the calibration range. The number of cycles depends on the accuracy required. In general,
two cycles are performed. The transducer, indicator and any connecting wiring are treated as a
system, and substitution of any parts of the system will invalidate the calibration. If the
transducer is to be used with one of several indicators, each combination will require a
separate calibration. For pressure transducers and pressure transmitters with voltage and
current output, suitable electrical measuring instruments with valid calibration certificates
should be used.

95
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Calibration results can either be presented as a table of corrections at nominal values over the
calibration range, or as a calibration equation relating the electrical output signal, R and the
applied pressure P. A general form of the calibration equation is

R = Qq + a^P +

It should be noted that the value of R is normally taken as the difference between a reading of
the transducer output after the application of a pressure and the reading at zero applied
pressure. The calibration equation is obtained by applying the method of least squares to fit a
polynomial curve to the calibration data, using one of a number of curve-fitting computer
programs that are available.

An example illustrating how the uncertainty in calibration may be determined is given in


section 4.7.3.

4.5.3 Calibration of ball and nozzle gauges

Strictly speaking, the ball and nozzle gauges are not piston gauges in as much as their effective
area is not clearly defined and their performance is dependent on the proper operation of the
gas inlet regulator. Nevertheless, the expression of pressure generated by ball and nozzle
gauges is essentially the same as that for piston gauges. In this case, the parameters Me and
AO are referred to as the effective mass of the ball and the effective area of the ball and nozzle
combination. Calibration of a ball and nozzle gauge is usually carried out by cross-floating by
the differential pressure method against a reference piston gauge standard with known
instrument constants. Data reduction and uncertainty analysis of the measurements are treated
in the same manner as for the piston gauges (Section 4.4).

4.6 Practical notes on the use of piston gauges


Some of the practical and safety aspects to be observed when using a piston gauge are
discussed briefly below.

Safety

Safety is of paramount importance when using a gas-operated piston gauge. All fittings and
tubing must be of the required pressure rating and safety screening must be used in the case of
working in the high pressure range of the instrument. The volume of the interconnecting tubing
should be kept as small as possible in order to minimise the potential energy contained in the
system. This is very much greater than for oil at the same pressure.

Cleaning

All tubing, valves and other fittings should be kept clean and free of foreign matter to prevent
damage to the piston and cylinder unit. Filters are placed at the gas inlet to the piston and
cylinder unit. In the case of gas-operated pressure instruments, cleanness of the piston and
cylinder is especially important. The piston and cylinder should be cleaned periodically in

96
JOHN MAN

order to achieve good results. Cleaning procedures are usually provided by the manufacturer.
A common procedure is to use a solvent, followed by rinsing with distilled water, polishing
with a lint-free tissue and then blowing with a jet of clean dry gas to remove any residue of
lint. The working surfaces of the piston and cylinder should not be handled with bare hands.
After cleaning and assembling the piston and cylinder unit, it is a good practice to spin the
piston with little or no added weight and to observe the deceleration rate which should be low
if the components are clean.

Levelling

The instrument should be levelled so that the axis of the piston and cylinder is vertical. Most
piston gauges have adjustable feet with a bubble level mounted on the base. It is advisable to
float the piston and place a level on the weight carrier when levelling the piston gauge. The
level is then checked over one complete revolution of the piston. If the piston axis is inclined
at an angle smaller than 1 minute of arc to the vertical, the error incurred is less than 1 ppm of
the gravitational force which may be ignored for most industrial purposes.

Environmental considerations

The instrument should be mounted on a rigid bench, which is able to support the weights and
the instrument without changing the level of the piston gauges. The instrument should be
shielded as fas as possible from draughts generated by air-conditioning units.

Magnetic and electrostatic effects

The spurious force due to magnetic and electrostatic effects will incur errors in the evaluation
of loading forces. Non-magnetic weights should therefore be used to minimise the magnetic
effects. For high precision measurements, the surface magnetisation of the weights and of the
piston-cylinder assembly should also be measured using a gauss meter. If the surface
magnetisation is found to be higher than 5 Gauss, it is advisable to demagnetise the
components by placing them inside a coil. The use of a bell-jar that made of Perspex materials
may generate electrostatic charges on its surface due to the rotation of the floating piston. One
way to minimise the electrostatic effect is to ensure the piston-cylinder combination is properly
earthed.

Aerodynamic effect

In gauge mode pressure measurements, a potential source of error is the effect of aerodynamic
forces acting on the rotational weights. These forces are generated by the airflow resulting
from the rotation of the weights. Such forces can be significant because gas-operated piston
gauges are often operated at high rotational frequency. It has been shown that the magnitude of
this effect depends on the rotational frequency and on whether a bell-jar is on place or not [6].
In order to minimise this effect, piston gauges should be operated in as low a rotational
frequency as is practicable, and where appropriate, with the bell-jar in place.

97
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Recalibration intervals

The recalibration intervals recommended by NATA for gas-operated piston gauges are given
as follows:
Instrument with uncertainty > 0.01% 5 years
<0.01% 3 years.

Refer to Chapter 5 for a discussion of the recalibration intervals of piston gauges.

4.7 Calculation of uncertainty


4.7.1 Uncertainty in pressure generated by a gas-operated piston gauge

The evaluation of the uncertainty associated with the pressure generated by a piston gauge can
be estimated using the mathematical model given below. The procedure and terminology used
in the evaluation are described in the ISO Guide to the Expression of Uncertainty [7, 8].

Mathematical Model

Equation (4.22), for the pressure measured by a gas-operated piston gauge in gauge mode, may
be modified to include a correction P/„ if pressure is not measured at the base of the piston, and
written as:

Au»C,(l + -V>) '


which has been simplified by using M to express the total conventional mass of the
load, A/, , and by combining terms into a buoyancy correction factor, C/, and a
temperature correction term. C,, where
G= 1--^,
8000
c, = 1 + (a,, + fife) (t -20)
= 1+2 a(t-20) if fifp = fife = (x,
and from equation (4.19):
Pi,=(Pt -Pa'^gh-
In applying the ISO Guide, and assuming all the parameters in the above expression are not
correlated, we calculate (equation (3.2) in [8]):

I
where Uc(P) is the combined standard uncertainty of the pressure, P
Ci is the sensitivity coefficient of the i* parameter (given below) and
ii(Xi) is the standard uncertainty of the i**’ parameter.

Each value of m(x,) is obtained from the initial, raw estimate of uncertainty, L/„ using an
appropriate reducing factor, via u(x,)= U,//:,. Also needed is a measure of the quality of

98
JOHN MAN

each estimate (/„ its number of' degrees of freedom, t<, as explained in [8]. For example, if the
estimate is judged ‘reasonable’ or ‘good’, Vi will be assigned the value 8 or 50, respectively.

Sensitivity Coefficients

The sensitivity coefficient, c„ for each of the input parameters in equation (4.27) is required.
They are: M, /?„, g, Aq^.o, a, t, 2,, Pf and h. Each value of c, is determined by taking the
partial derivative of P with resj)ect to that parameter.

Notice that P appears in the clenominator of this equation, in the correction term (7 + AP).
However, because this term is <close to unity (within a few ppm) any uncertainty in P will have
no significant affect on the va lue of (7 + AP). So, in differentiating equation (4.27) we may
regard P as a constant in (7 + /IP). The coefficients are:

• Mass, M C -
Ao,j(,C,(l + /lP)

• Air Density,
^Cl, ' <^Pa
----- 1_^ _
Ao 2oC,(1 + AP) V 8000)

• Effective Mass, Me
Ao2oC,(I + /IP)
MCi,+M,
• Local Gravity, g C,, =-------- -------- ---- + (P f - Pa
>■ /lo.2oC,(l + /lP) '
_ -p
• Effective Area, Ao,2o ^A).2O 4
^0,20

3P 3C.
• Expansion Coefficients, a =----- .------
dC, da

= - — .2(t-20)
Cr

• Temperature, t
~ dc, ■ dt
= -^.2a
c,

• Distortion Coefficient, z c
” (1 + XP}

• Fluid Density, Pf Cp, = gh

• Height of Head, h Ch = {Pf - Pa

99
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Parameter values

In this example, the measured parameter P, the input quantities and the corrections Ch, C, and
Ph have the following values:
Symbol Parameter Value
P measured pressure lE+06 Pa
M total mass of load 0.844658 kg
Pa density of air 1.21 kg.m'^
Me effective mass 0.0117943 kg
S local value of gravity 9.7963 m.s'2
^0.20 effective area at 1 atmosphere and 20 C 8.38882E-06 m^
a expansion coefficient of piston and cylinder 4.55E-06 °C ’
t temperature 19.5 °C
A distortion coefficient 1.042E-12 Pa’
Pf density of fluid 12.66 kg.m'^
h height of head 0.016 m
Ch buoyancy correction 0.999849
C, temperature correction 0.999995
p, pressure head correction 1.79 Pa

Components of uncertainty

• Conventional mass, M
As an example, a set of stainless weights of a piston gauge were calibrated against standards of
known mass and density at NML. The weighings were made in air. The 95% uncertainty
associated with the conventional masses of the weights was ±1.5 ppm (of 0.844658 kg). So,
U, = 121 X 10’^ kg, k,= 2.0 and v, = 50.

• Air density, p,,


The density of ambient air depends on the ambient air pressure, temperature and relative
humidity as well as its composition. In this example, the air pressure, temperature and relative
humidity in the pressure laboratory have been measured during the measurements. The air
density was calculated using the expression recommended by the International Bureau of
Weights and Measures. The air density was reported as 1.21 kg.m’^ ± 0.025 kg.m'‘\ The
uncertainty will be taken as a 95% value from a Gaussian distribution, so k,= 2.0 and t< = 8.

• Effective mass of the piston,


The value of the effective mass of the piston and its uncertainty is obtained from the
calibration report, as 11.7943 ± 0.0057 g with a coverage factor of 2.26. Thus, kt = 2.26,
corresponding to 9 degrees of freedom, from Student’s tables [8].

• Local value of acceleration due to gravity g

100
JOHN MAN

The value of local gravitational acceleration can be estimated using the formula recommended
by the International Association of Geodesy. The formula gives the best simple method of
calculating g at a place where it has not been measured. It is likely to give results within 100
ppm. In this example, the value of the local gravity was obtained from the published literature
given by Geoscience Australia. The uncertainty was estimated to have an uncertainty of ±25
ppm, which will be taken as a 95% value. A Gaussian distribution is assumed, so ki = 2.0.

• Effective Area, Ao,2o


Value of the effective area of the piston-cylinder combination at atmospheric pressure and
20 °C is also obtained from the calibration report, as 8.38882 ± 0.00020 mm^ with a coverage
factor of 2.26. Thus, (7, = 2.0 x 10 ’° m\ ki = 2.26, and v, = 9.

• Thermal expansion coefficient of piston-cylinder combination, (X,,, (Xc


The thermal expansion coefficient of the piston-cylinder combination is given by the
manufacturer as o^, = ex, = 4.55 x 10'^ °C'' with a maximum ‘accuracy’ of ±10 %. Taking the
notation of a to represent both and , and assuming a Gaussian distribution for the
uncertainty, we take f/, = 4.55 x 10'^ °C ', ki = 2.0, and v, = 8.

• Temperature of the piston and cylinder combination, t


The temperature of the piston and cylinder combination has been measured during operation. It
was reported as 19.5 °C ± 0.1 °C. Thus, Ui = 0.1 °C, C = 2.0, and v; = 8.

• Distortion Coefficient, 7.
The value of 2 given in the calibration report for the piston/cylinder is 1.042 x 10'^ kPa ’ with
an uncertainty of ±4.1 x 10*'° kPa ’ and a coverage factor of 2.26. So, k,■ - 2.26 and = 9.

• Density of fluid, pf
It is seen from equation (4.20) that pf is a function of P. Most of the operating pressure range
is less than 10 MPa, and the uncertainty of pf calculated from equation (4.20) should be within
±5% of the calculated value, in this case, 12.66 kg.m'\ Taking this estimate of uncertainty as a
95% Gaussian value and reasonable, we have JJ-, = 0.633 kg.m'\ A'/= 2.0, and v, = 8.

• Height of pressure head, h


The height of the fluid head was measured to the nearest mm. So, we put (7, = 1 x 10'm,
ki = 2.0, and v, = 8.

Uncertainty analysis

The above components of uncertainty and their characteristic values are presented in Table 4.3.
Included are the calculated values leading to an expanded uncertainty of 36 Pa, which is based
on a coverage factor of 2.11 and a coverage probability of 95%. Note that the uncertainty
contribution due to gravity is the most significant component.

101
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Table 4.3 Estimation of uncertainty associated with a pressure of 1000 kPa generated
by a gas-operated piston gauge
Component Units Ui ki V/ 4r,) Ci lc/.af x,)l ICi.u( X,)lVv,

Total mass, kg 1.27E-06 2 50 6.35E-07 1.17E+06 0.743 0.552 0.01


M
air density, kg.m’ 0.025 2 8 0.012 -123.5 1.54 2.38 0.71
Pa
EfTective mass. kg 5.66E-06 2.26 9 2.50E-06 1.17E+06 2.93 8.59 8.19
Me

Gravity, m.s‘ 2.50E-04 2 8 1.25E-04 1.02E-I-05 12.75 162.6 3303

Effective Area, m* 2.00E-10 2.26 9 8.85E-11 -1.19E+11 10.53 110.9 1367


4 0,20

Expansion Pa' 4.55E-7 2 8 2.28E-7 1 .OOE-i-6 0.23 0.05 0.00


coefficient, (X
Temperature, °C 0.1 2 8 0.05 -9.1 0.46 0.21 0.01
t
Distortion Pa' 4.10E-I3 2.26 9 1.81E-13 -l.OOE-i-12 0.18 0.03 0.00
coefficient, 2
Fluid density, kg.m’ 0.633 2 8 0.316 0.157 0.05 0.002 0.00
Pf
Head height, m 0.001 2 8 0.0005 112 0.06 0.003 0.00
h
Sums 285.28 4679
Combined Standard uncertainty, Uc 16.89 Pa
Effective degrees of freedom, Veg 17.4
Coverage factor, k=Student's t for v>,yand CL 95% 2.11
Expanded uncertainty, U=kuc 35.7 Pa

4.7.2 Uncertainty in calibration for a Bourdon pressure gauge

This example presents an evaluation of the uncertainty for the calibration of a bourdon tube
test gauge that has an operating range from 0 to 100 kPa. The gauge dial diameter is
approximately 165 mm and has major scale markings at 10 kPa intervals and minor scale
markings at 0.5 kPa intervals.

The gauge is calibrated in accordance with the AS 1349-1986. A pneumatic dead-weight tester
is used to apply pressures to the test gauge in a series of increasing and decreasing pressures.
This is repeated. The pressure fluid is nitrogen and the calibration temperature is 19.7 °C.

The model

The measurand is the correction to be applied to the test gauge reading at a given pressure. The
model is:

102
JOHN MAN

C — Pa Ph— Pi
and from equation (4.19):
Ph =^Pf -

where: C is the pressure correction,


Ph is the pressure head correction.
Pa is the pressure applied by the dead-weight tester and
Pi is the pressure reading of the test gauge.

As explained in section 7.4.1, each value of «(%,), the standard uncertainty of the i*’’ parameter
in the model, is obtained from the initial, raw estimate of uncertainty, Z7„ using an appropriate
reducing factor, ki, using m(x,) = (/, / A:,. Also needed is a measure of the quality of each
estimate its number of degrees of freedom, vi, as explained in [8]. For example, if the
estimate is judged ‘reasonable’ or ‘good’, w will be assigned the value 8 or 50, resp.

Sensitivity Coefficients

The sensitivity coefficient, c„ for each of the input parameters in the model is required. They
are Pa, Pf, Pa, g, h and P,. Each value of c, is determined by taking the partial derivative of C
with respect to that parameter. They are given below together with the values assumed for this
example calculation.

Parameter Value Sensitivity coefficient, c,


Pa lE+05 Pa 1
Pf 3.45 kg.m’^ gh
Pa 1.21 kg.m'^ -gh
g 9.1963 m.s'2 (Pf - Pa)h
h 0.100 m (Pf-Pa)g
P, lE+05 Pa -1

Components of uncertainty

As mentioned above, the test gauge that has an operating range from 0 to 100 kPa. Its
calibration involved two runs with various pressures applied while the pressure was increasing
and then decreasing. For this example, the uncertainty will be determined for 100 kPa only.

• Applied pressure, Pa

The expanded uncertainty of the pressure applied by a dead-weight tester is taken from the
relevant calibration report. In this example, the uncertainty of the applied pressure at the base
of the piston is given by the calibration certificate as ±0.025% (i.e., ±25 Pa at 100 kPa) with a
coverage factor k = 2. This factor may have been rounded, so assume 20 degrees of freedom
[8]. Drift of the dead-weight tester is assumed to be insignificant compared to the uncertainty
of the applied pressure.

103
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

• Head correction. Ph

As there was a height difference between the reference level of the DWT and the test gauge,
the uncertainty in the pressure head correction is required. The dominant component is that due
to the measurement of the height difference, h, which was 100 ±10 mm. We take this
uncertainty to be a 95% value and ‘reasonable’, i.e., U, = 0.010 m, = 2 and = 8.

The values of g and Pa and their uncertainty was taken as the same as given in section 4.7.1.
The value of pj is different (see above), although its uncertainty is 5%, as before.

• Pressure gauge reading, P,

There are three components of uncertainty for P,:


(i) The uncertainty due to the readability of the pressure gauge, which was taken as 1/5 of
minor scale marking, that is, a semi-range of 1/5 of 0.5 kPa, which is 0.1 kPa. A rectangular
distribution is assumed (I; = 1.73 ) with 50 degrees of freedom.
(ii) The uncertainty due to the repeatability of the pressure gauge readings. The maximum
difference between the readings in the two runs, either increasing or decreasing (see
Table 4.4), was taken as the estimate of uncertainty. It was 0.2 kPa or ±0.1 kPa, i.e., U, = 100
Pa. This will be assumed a 95% Gaussian estimate (A:, = 2) with = 8.
(iii) The uncertainty arising from hysteresis. This is found from the maximum difference
between decreasing and increasing readings observed in the two runs. At a gauge reading of
100.3 kPa, the differences between the increasing and decreasing readings are 0 and 0.1 kPa
for Run 1 and Run 2 respectively. From this we take U, = 50 Pa, kj = 2 and = 8.

Table 4.4 Calibration results of the 100 kPa test gauge


Applied Readings (kPa) Corrections
Pressure Run 1 Run 2
(kPa) Increasing Decreasing Increasing Decreasing Increasing Decreasing Mean
0.00 0.0 0.1 0.1 0.1 -0.03 -0.05 -0.04
10.07 9.9 10.0 10.0 10.0 0.11 0.06 0.09
20.16 20.0 20.1 20.1 20.1 0.16 0.08 0.12
30.16 30.0 30.1 30.1 30.1 0.16 0.08 0.12
40.26 40.0 40.1 40.0 40.1 0.25 0.15 0.20
50.21 50.0 50.1 50.0 50.1 0.21 0.11 0.16
60.15 60.0 60.1 60.0 60.1 0.15 0.05 0.10
70.25 70.1 70.2 70.1 70.2 0.15 0.05 0.10
80.25 80.1 80.2 80.1 80.2 0.15 0.05 0.10
90.35 90.2 90.3 90.2 90.3 0.14 0.04 0.09
100.30 100.1 100.1 100.2 100.3 0.15 0.10 0.12

104
JOHN MAN

Uncertainty analysis

The above components of uncertainty and their characteristic values are presented in Table 4.5,

Table 4.5 Uncertainty analysis the calibration of a Pressure Gauge at 100 kPa

Component Ui ki u(Xi) Ci Vi 1 Ci u(Xi)\

Applied pressure. Pa 25 Pa 2 12.5 1 20 12.5


Fluid density, Pf 0.173 kg.m’^ 1 8.65E-02 0.98 8 0.08
Air density, Pa 0.025 kg.m-^ 1 1.25E-02 -0.98 8 0.01
Local gravity, g 0.00025 m.s'2 1. 1.25E-04 0.224 8 0.00
Head height, h 0.010 m 2 5.OOE-O3 21.94 8 0.11
Readability for P, 100 Pa 1.73 57.8 -1 50 57.8
Repeatability in P, 100 Pa 2 50 -1 8 50.0
Hysteresis in P, 50 Pa 2 25 -1 8 25.0

Since the uncertainty components are uncorrelated, we have for the combined standard
uncertainty [8] in C:

III =^lc.w(x,.)|2,
i
Uc = 81.38 Pa

and Veft =---- --------- -r- = 41.6.

I',
So, k = 2.02 and U = k.Uc = 164.4 Pa or 0.16 kPa when rounded to two significant places.

Uncertainty associated with the uses of pressure gauges

When an operator uses a calibrated pressure gauge, the operator takes a gauge pressure
reading, P„ and then applies the correction at that given pressure to give the actual applied
pressure. The model is

Pa = Pi+ C

In this case, the uncertainty associated with P,, is calculated by combining the uncertainty of P,
and C. The uncertainty of C is given in the calibration report. The uncertainty of P, depends on
the working environments and conditions. These may be quite different from the laboratory
conditions in which the pressure gauge is calibrated.

105
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

For example, a calibrated pressure gauge was used to measure a single point pressure
nominally at 100 kPa in a pressure line. It is observed that the pressure reading fluctuates
between 100 and 100.2 kPa. A mean value of 100.1 kPa is taken as the gauge reading.
Conservatively, we may assume a rectangular distribution for the uncertainty, with a semi­
range of 0.1 kPa and put U, = 0.1 kPa, ki= 1.73 and v, = 8 (being reasonable).

From the calibration report on a pressure gauge, it is found that the correction is 0.12
±0.16 kPa with a coverage factor of k = 2.02, equivalent to about 40 degrees of freedom. That
is, we have: (7, = 0.16 kPa, ki = 2.02 and i< = 40.

Since the sensitivity coefficients for P, and C are both unity, iic = 0.0981, = 38.9, k = 2.03
and U = 0.199 kPa, which rounds to 0.20 kPa.

Thus, for this example, the corrected pressure reading is 100.22 ± 0.20 kPa.

4.7.3 Uncertainty in calibration for a pressure transducer

The calibration of pressure transducers is discussed in sections 4.5.2 and more generally in
Chapter 6. In this example, a transducer of capacity 2000 kPa gauge pressure in combination
with a digital multimeter was calibrated against a gas-operated dead-weight tester. A series of
twelve increasing and twelve decreasing pressures was applied. This was repeated, and the
voltage measured on the multimeter was recorded for each calibration pressure. The pressure
transmitting fluid was nitrogen and the ambient temperature was 21 °C.

The model

The ‘model’ is the calibration equation given in section 4.5.2, which in this case is the least­
squares fit:
R = 0.000189 + 4.99675 X 10'^ P +7.1998 X 10’^/^,

where R is the voltage displayed on the multimeter, in volts, and P is the pressure in kPa.

The sensitivity coefficient, c„ for P = dP/ oR = 4.99675 x 10'^+ 1.43996 x 10’^ P


= 0.00500 V/kPa at 150 kPa
= 5.00 mV/kPa

Uncertainty components for P

(1) The calibration uncertainty for the applied pressure. From the report for the dead-weight
tester, we have that U, = 0.00452 kPa at 150.55 kPa, which has a stated coverage factor of
2.02. This value becomes the reducing factor, K, for U, and is equivalent, = 40 [8]. Drift in
calibration of the dead-weight tester was considered insignificant.

(2) Based on the calibration history of the transducer, the drift of the transducer during a
recalibration period of 1 year is estimated to be no more than 0.001%. That is, at 150.5 kPa,

106
JOHN MAN

Ui = 0.00151 kPa. Assuming the drift is Gaussian, put ki = 2 and take Vi = 8, since the estimate
is ‘reasonable’ [8].

Uncertainty components for R

(3) The type A uncertainty for each ‘reading’, which was effectively the average of about 20
values, as judged from the DVM display. The range of values was estimated to be 3 in the last
digit at 150 kPa, i.e., 0.3 mV. For this approximate method [8] of evaluating type A
components, take f/, = 0.3 with a reducing factor of 15. For 20 readings v, = 19.

(4) The resolution of the DVM is 0.1 mV. However, since resolution errors contributed to the
observed scatter, its effect is included in the type A evaluation, above.

(5) Possible changes in ambient temperature during calibration, considered to be about ±1 °C.
Assume the temperature coefficient for the transducer is 0.002%/°C, from the manufacturer’s
specification, and that it is a 95% value. Thus, we have (/, = 0.015 mV at 0.75 V (for 150 kPa)
and k, = 2. Since this is a reasonable estimate v, = 8.

(6) The uncertainty in the fit. The standard deviation for the residuals (data - fit) of Figure 4.15
is expected to be about 0.3 mV, judged by eye (±0.3 covers about 67% of the data) and that
given by the fitting routine is 0.316 mV with 45 degrees of freedom. So, the standard deviation
for the fit, effectively a mean with 45 dof, is 0.316/^45 = 0.047 mV. Here, ki= 1 and v, = 45.

This component covers only the quality of the fit assuming the function (quadratic) chosen for
the fit is correct (see item 7).

(7) Quality of the model — how well does the fit represent the true behaviour, i.e., what is the
estimate for any systematic error not covered by the fit? This was judged to be no more than
about ±0.1 mV, since the residuals in Figure 4.15 show no underlying dependence on P, just
scatter, i.e., take U,- = 0.1, ki= 2 and v, = 8 (reasonable).

(8) Hysteresis in the fit: on examination of Figure 4.15, any hysteresis would be about
±0.2 mV at 150 kPa, increasing to ±0.4 mV at 2000 kPa. So, put U, = 0.2, k, = 2 and v, = 8
(reasonable).

(9) Rounding of the final result (to the nearest 0.1 mV). Put (7, = 0.05 mV and v, = 1000 [8].
Being rectangular, put ki= 1.73.

The above components appear in Table 4.6, which contains u(Xi), the standard uncertainty
equivalent to f/, and obtained as f/,7 A:,. From this data we have (using [7, 8]):

Me = 0.1270 mV and
Vejf = 19.4.
So, k = 2.096 and
(7 = 0.2662 or
= 0.27 mV when rounded.

107
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

As an aid to the user, U may also be expressed in pressure units, using the sensitivity
coefficient given above. That is:
(7 = 0.2662/5.0 = 0.05324
= 0.053 kPa, when rounded.

Some transducers may have their own linearisation and scaling function built-in and give an
output reading in terms of pressure units. In these cases, the selected pressure units of the
output readings must be reported. If the applied pressure and the output reading have the same
units in pressure, the sensitivity coefficient, c is equal to 1.

Table 4.6 Uncertainty analysis for a transducer at 150 kPa

Component Ui ki M(Xf) Ci Vi Ic,m(x,)I


1 Applied pressure 0.00452 kPa 2.02 0.00224 5.00 mV/kPa 40 0.0112
2 Drift 0.00151 2 0.00076 5.00 mV/kPa 8 0.0038
3 Type A 0.3 mV 15 0.02 1 19 0.02
4 Resolution 0 1.73 0 1 1000 0
5 Ambient temp. 0.015 2 0.0075 1 8 0.0075
6 Fit 0.047 1 0.047 1 45 0.047
7 Quality of model 0.10 2 0.05 1 8 0.05
8 Hysteresis 0.2 2 0.1 1 8 0.1
9 Final rounding 0.05 1.73 0.0289 1 1000 0.0289

Figure 4.15: Calibration data for a pressure transducer over two runs: 1 (up and down
arrows) and 2 (arrows to the right, for going ‘up’, and to the left, for ‘down’).

108
JOHN MAN

4.8 Appendix. Sample measurement reports

4.8.1 Report of calibration for a piston/cyUnder and weight set

National Measurement Laboratory


Bradfield Road, West LindHeld, NSW 2070, Australia
Phone: +61 2 9413 7000 Fax: + 61 2 9413 7202

Measurement Report on
A Piston/CyUnder Unit
and a
Weight Set

Serial Number: XXXX

The tests, calibrations and measurements reported in this document have been
performed in accordance with NATA requirements, which include the requirements of
ISO/IEC Guide 25:1990. Accreditation Number 1.

For: CSIRO, National Measurement Laboratory


Bradfield Road
West Lindfield NSW 2070

Reference: Internal request

Description: A pneumatic piston gauge comprising a base, detachable


piston-cylinder unit and a set of 14 stainless steel weights.
The piston-cylinder unit has a nominal area of 8.4 mm' and
an operating pressure range from 14 to 7000 kPa.

Manufacturer: XXXX

Model: 1200

Instrument markings: Piston and cylinder unit: XXXX


Weight set: 50666
Weight: Marking 1 to 14

Date of test: July 2002

109
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Calibration of the weights

The mass values of the set of 14 weights were determined by weighing in air against standards
of density 8000 kg m'^. The uncertainty of these values is ±1.5 ppm. The coverage factor for
the uncertainties is k = 2.0.

To obtain the results, which are given in Table 1, it was assumed that the densities of the
weights lie within the required range for weights of OIML class Fp The set of weights was
calibrated in air of density 1.2 ± 0.034 kg.m'^.

The mass value assigned to each weight represents, to within the corresponding uncertainty,
the mass of a hypothetical object of density 8000 kg.m"^ which, in air of density 1.2 kg.m’^,
would balance that weight.

Table 1 Weight values

Weight Mass Uncertainty (±)


identification (g) (g)
1 500.0247 0.00075
2 1000.0282 0.0015
3 1000.0320 0.0015
4 1000.0026 0.0015
5 1000.0272 0.0015
6 1000.0247 0.0015
7 500.0348 0.00075
8 300.0003 0.0003
9 200.0070 0.0003
10 100.00803 0.00015
11 50.01075 0.0001
12 30.00834 0.00009
13 20.00366 0.00008
14 10.00769 0.00006

Pressure evaluation

The pressures associated with a given combination of weights on the piston assembly may be
calculated by:
ZM 1-^
xg
I Pm )
X„(l + aP'Xl + («„+aJr-20))

no
JOHN MAN

where: P = pressure at the instrument reference level, kPa


2M = sum of the mass values of weights on the piston assembly, g
Pa = density of air , kg m'^
Pm = density of weights, kg m’^
Me = effective mass of the piston, g
g = gravitational acceleration, m s'^
Ao = effective area of the piston-cylinder unit (PCU) at atmospheric
pressure and 20 °C, mm^
A = pressure distortion coefficient of the PCU, kPa ’
6^, = linear thermal expansion coefficient of the piston, °C ’
a, = linear thermal expansion coefficient of the cylinder, °C*’
t = temperature of the PCU during the pressure determination, °C

P' is the approximate pressure calculated by:

Instrumental constants

The instrumental constants Aq, A, and Me yNQVQ determined experimentally in air by comparison
with a pressure standard of known instrumental constants. Twelve data pairs of force and
pressure were obtained from the calibration process. The pressure medium was nitrogen. The
ambient temperature during the test was in the range 19.5 to 19.6 °C. The constants for the
piston/cylinder unit (PCU) were determined over the pressure range 200 to 7000 kPa.

The effective ms of the piston includes the true mass of the piston and any small, indeterminate
vertical forces which may be acting on the piston when in a balanced condition.

Values of the instrumental constants together with the density value of the piston are given in
Table 2. The instrumental constants are determined at a reference temperature of 20.0 °C.

Table 2 Instrumental constants at 20.0 °C

Ao(mm^) 8.38882
at 20 °C ±0.00020

1.042 X 10*’
Z (kPa*’)
±4.1 X 10*’°

Me(g) 11.79434
±0.0057

111
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Reference level

Calibration pressures have been calculated with reference to the base of the piston.

Calculation of pressure

Knowing the instrumental constants, the pressure generated by a given load acting on the
piston and cylinder unit at the actual operating conditions of the instrument can be calculated
by using the equation given in this report.

The quantities specific to the actual operating conditions of the instrument are head height
correction, local value of gravity, operating temperature and air density.

If the reference temperature, tref, is not 20°C, the effective area Ao at tref is found by:

= A)’2O ref “20))

Uncertainty
The uncertainties associated with the pressure values generated by the instrument are given in
Table 3.
Table 3 Uncertainties associated with pressure values

Pressure Uncertainty ± Pressure Uncertainty ±


(kPa) (ppm) (kPa) (ppm)

200 47 1000 37
300 40 1500 37
400 40 2000 37
500 38 3000 36
600 37 4000 36
700 37 5000 36
800 38 6000 36
900 37 7000 36

The uncertainties of the pressure values given in table 3 have been estimated from the
calibration results and include uncertainty contributions from the following:

Mass Set ±1.5 ppm


Buoyancy ±3 ppm
Height difference ±0.001 m
Local Gravity ±25 ppm
Linear thermal expansion coefficient ±10%
Temperature measurement ±0.1 °C

112
JOHN MAN

Notes

The uncertainties associated with the pressure values given in Table 3 are those associated
with the calibration and the calibration environment. These values may not apply at the site
where the instrument is used and the uncertainty budget may need to be re-evaluated to take
account of local environmental conditions and changes in the instrumental constants with time.

A value for the linear thermal expansion coefficient of the piston/cylinder unit of
a^, =a^ = 4.55x10"^ has been used in the calculations.

The calibration of this instrument was carried out according to the Test Method and Work
Instruction detailed in the NML Project operations manual EAFA - Mass and Related
Quantities. The relevant Test Method for this calibration is EAFA 8.2.15 “Calibration of
Pressure Balances” and the relevant work Instruction is EAFA 9.2.21 “Pneumatic Pressure
Procedures”.

The uncertainty has been calculated in accordance with the principles in the ISO ‘Guide to the
Expression of Uncertainty in Measurement’, and gives an interval that is estimated to contain the
measurand with 95% probability. The coverage factor for the interval is k = 2.0.

The stated uncertainty applies at the time of test only and takes no account of any drift or other
effects that may apply afterwards.

J. Man N. Bignell
for Dr B. D. Inglis NAT A Approved Signatory
Director
National Measurement Laboratory

113
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

4.8.2 Report of calibration for a ball and nozzle piston gauge

National Measurement Laboratory


BradOeld Road, West LindOeld, NSW 2070, Australia
Phone: +61 2 9413 7000 Fax: + 61 2 9413 7202

Measurement Report on

Ball and Nozzle


Piston Gauge

Type: XXXXX
Serial Number: XXXXX

The tests, calibrations and measurements reported in this document have been
performed in accordance with NATA requirements, which include the requirements of
ISO/IEC Guide 25:1990. Accreditation Number 1.

For: A Pressure Laboratory


Bradfield Road
West Lindfield NSW 2070

Reference: Quotation No. 123

Description: The instrument is a ball and nozzle type pneumatic


pressure balance which is provided with a set of fourteen
kPa weights and weight carrier. It has an operating range
from 50 to 7000 kPa.

Manufacturer: XXXX

Model: XXXXX

Instrument markings: Nozzle: PM54321


Weights: 12345, 1 to 14 (nominal pressure values)

Date of test: July 2002

114
JOHN MAN

Calibration of the weights

The mass values of the set of 14 weights were determined by weighing in air against standards
of density 8000 kg m''\ The uncertainty of these values is ±5 ppm. The coverage factor for the
uncertainties is k = 2.0.

To obtain the results, which are given in Table 2, it was assumed that the densities of the
weights lie within the required range for weights of OIML class F2. The set of weights was
calibrated in air of density 1.2 ± 0.034 kg.m '^.

The mass value assigned to each weight represents, to within the corresponding uncertainty,
the mass of a hypothetical object of density 8000 kg.m’^ which, in air of density 1.2 kg.m'‘\
would balance that weight.

Pressure evaluation

The pressures associated with a given combination of weights on the piston assembly may be
calculated by:

P= ---------------- L---

where: P pressure at the instrument reference level, kPa


=
XMi sum of the mass values of weights on the weight carrier, g
=
Ch air buoyancy correction factor, taken as 0.99985
=
Me effective mass of the ball and weight carrier, g
=
g Gravitational acceleration, m s “
=
Ae effective area of the ball and nozzle combination, mm‘
=
area thermal expansion coefficient of the ball and nozzle
=
combination, °C'’
t = temperature of the ball and nozzle during the pressure
determination, °C

Instrumental constants

The effective mass of the ball and weight carrier was determined experimentally, and it
includes buoyancy effects, as well as any small, indeterminate vertical forces that may be
acting on the ball when in a balanced condition.

The instrumental constants Me and Ag were determined by comparison with a standard pressure
balance of known instrumental constants. The pressure fluid was nitrogen and the temperature
during the test was in the range 19.1 to 19.5 °C.

115
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

Calibration results

The values of the instrumental constants Me and Ag are listed in Table 1

Table 1 Instrumental constants


Constant Value Uncertainty
Effective Area (Ae)
28.52624 mm^ ±0.00049 mm^
At 20°C
Effective Mass (Me) 146.154 g +0.21 g

The mass values of the weights and the effective mass of the ball and weight carrier
combination are given in Table 2 together with the values of the pressure they develop for a
gravitational acceleration of 9.79637 ms'^.

Table 2 Mass and pressure values


Mass Nominal Actual
Weight Mass Uncertainty Pressure Pressure
Number (g) (+ g) (kPa) (kPa)
1 2766.624 0.014 950 949.9608
2 2912.120 0.015 1000 999.9189
3 2911.897 0.015 1000 999.8424
4 2912.214 0.015 1000 999.9512
5 2912.074 0.015 1000 999.9031
6 2912.169 0.015 1000 999.9358
7 1456.125 0.007 500 499.9818
8 582.390 0.003 200 199.9721
9 582.387 0.003 200 199.9711
10 291.2152 0.0015 100 99.9930
11 145.5969 0.0007 50 49.9928
12 58.2421 0.0003 20 19.9983
13 58.2405 0.0003 20 19.9977
14 29.1205 0.0003 10 9.9989
Me
Ball & weight carrier
146.1539 0.0007 50 50.1916

Uncertainty

An estimation of the uncertainties associated with the output pressures generated by the
instrument is given in Table 3 and include NML Type B uncertainties and the Type A
uncertainties associated with pressure and mass calibrations.
The uncertainties of the output pressures given in Table 3 are those associated with the
calibration and the calibration environment. These values may not apply at the site where the

116
JOHN MAN

instrument is used and the uncertainty budget may need to be re-evaluated to take account of
local environmental conditions and changes in the instrumental constants with time.

Table 3 Uncertainties
Pressure Uncertainty
(kPa) (±%)
350.0 0.050
500.0 0.033
1000.0 0.016
2000.0 0.0080
3000.0 0.0070
>4000.0 0.0060

Notes

1. If the gravitational acceleration at the site where the instrument is used varies from the
value quoted above (9.79637 m s'^) then, the tabulated pressure values given in Table 2
may be adjusted by multiplying them by the site value of g and dividing by the gravity
value quoted above.

2. The pressures listed in Table 2 apply to the nozzle outlet level. For other levels, fluid head
corrections may be necessary.

3. The calibration of this instrument was carried out according to the Test Method and Work
Instruction detailed in the NML Project operations manual EAFA - Mass and Related
Quantities. The relevant Test Method for this calibration is EAFA 8.2.15 ‘‘Calibration of
Pressure Balances” and the relevant work Instruction is EAFA 9.2.21 “Pneumatic
Pressure Procedures”.

4. The uncertainty has been calculated in accordance with the principles in the ISO ‘Guide to
the Expression of Uncertainty in Measurement’, and gives an interval that is estimated to
contain the measurand with 95% probability. The coverage factor for the interval is k = 2.29.

5. The stated uncertainty applies at the time of test only and takes no account of any drift or
other effects that may apply afterwards.

J. Man N. Bignell
for Dr B. D. Inglis NAT A Approved Signatory
Director
National Measurement Laboratory

117
CHAPTER 4. THE GAS-OPERATED PISTON GAUGE

4.9 References

[1] J W Schmidt el al. “A primary pressure standard at 100 kPa”, Metrologia, 1999, 36,
pp 25-529.

[2] R S Davis, “Equation for the Determination of the Density of Moist Air (1981/91)”,
Metrologia 1992, Volume 29, pp 67-70.

[3] RS Dadson, S L Lewis and G N Peggs. The Pressure Balance: Theory and Practice.
National Physical Laboratory, London: Her Majesty’s Stationery Office

[4] G W Kaye and T H Laby. Tables of Physical and Chemical Constants, 16* Edition.

[5] F Pavese and G Molinar. Modern Gas-Based Temperature and Pressure Measurements
Plenum Publishing Corporation.

[6] D B Prowse and D J Hatt. “The effect of rotation on a gas-operated free-piston pressure
gauge”. Journal of Phys. E: Scientific Instruments, vol 10, pp 450-461, 1977.

[7] ISO Guide to the Expression of Uncertainty in Measurement. International Organization


for Standardization, Geneva, 1995. ISBN 92-67-10188-9.

[8] R.E. Bentley Uncertainty in Measurement'. The ISO guide. CSIRO, Sydney, Monogr. 1:
NML Technology Transfer Series, Edition 6, 2003.

Further reading

The Pressure Balance: A practical guide to its use. Sylvia Lewis and Graham Peggs, National
Physical Laboratory, UK, HMSO, 1982.

The Calibration of Pressure Instruments using a Pressure Balance. CSIRO, Division of


Applied Physics, Technical Memorandum, No. 19, 1992.

CCM Second International Seminar, Pressure Metrology from 1 kPa to 1 GPa. Metrologia,
Volume 30, Number 6, April 1994.

118
Chapter 5

The hydraulic piston gauge and its use


Walter J Giardini

5.1 Introduction
In Australia most industrial hydraulic pressure measurements occur in the 1-100 MPa range.
Some users (e.g., explosives and munitions manufacturers) use higher pressures up to 500-700
MPa. In the industrial pressure range, the piston gauge is used to measure pressure accurately.
A screw press is used to generate pressure in a hydraulic fluid circuit and at one point in the
circuit, the fluid acts from below on a vertically aligned piston-cylinder assembly on which
extra masses can be loaded to increase the effective weight of the piston. When the hydraulic
pressure in the fluid is just sufficient to lift the weight of the loaded piston, then it is balanced
against the hydraulic pressure of the fluid. The piston is rotated to relieve friction, generating a
thin layer of oil between it and the cylinder. This system is also called a pressure balance,
because its operating principle is that a hydraulic pressure is balanced against a mechanical
load. It is also called a dead weight tester because the mechanical load is generated by the
action of the gravitational field of the earth acting downwards on the "passive" or "dead" mass
of the piston and extra masses. When a port is introduced into the pressure circuit, then the
pressure at that port is calculable from the known pressure at the datum level of the piston (that
is the level at which the reference pressure is calculated), combined with the extra head of
pressure due to the difference in the vertical level of the port exit and the piston datum.

Pressure is defined as the ratio of force acting over a surface. Since a piston-cylinder has a
calculable area, the ratio of the gravitational and other loads bearing down onto the cross
sectional area of the piston cylinder is a mechanical realization of the pressure. Since the
mechanically-generated pressure is balanced against the hydraulic pressure generated by the
fluid acting on that same cross-sectional area, the two values of pressure can be equated. The
value of the pressure obtained by the mechanical realization of the piston-cylinder loads and
areas can be calculated, and is said to be equal to the hydraulic fluid pressure in that plane.

The load applied to the piston depends on the intrinsic instrument parameters, such as the mass
of the components and their densities, and also on operational and environmental parameters
such as the surface tension of the hydraulic fluid acting on the emergent stem of the piston, air

119
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

buoyancy acting on the weights and the value of the gravitational constant g. The area of the
piston-cylinder, usually referred to as the effective area, must account for such things as the
difference in areas of the piston and the cylinder, the intervening annular oil filled gap and the
temperature of the assembly. Most importantly the geometry of the piston-cylinder unit
undergoes distortion under the action of the applied pressure and this must be taken into
account, particularly at higher pressure.

The uncertainty in measurement of the value of the pressure generated therefore is in part a
function of parameters that are intrinsic to the instrument, and in part a function of
environmental and operational parameters that operate at the time of use.

The most common application of a pressure balance is its use to generate and measure a
pressure at its output port, which is then connected to a pressure measuring device for
calibration. Sometimes the pressure balance is connected to a pressure system to directly
determine the operating pressure of the system, or its change with time, as for example in
pipeline testing, which makes use of the relatively high accuracy, stability and sensitivity of
the pressure balance. When one pressure balance (the reference standard) is used to calibrate
another (the test unit), the two units are connected or "cross-floated", and knowing the
parameters (primarily the area as a function of pressure), and hence the true pressure generated
by the standard, the equivalent parameters of the test unit can be calculated.

It is convenient for many users to associate each weight directly with a nominal value of the
pressure it generates when applied to its associated piston. This leads to some complications in
that when the area of the piston-cylinder is pressure dependent, the same weight will generate
a different pressure increment depending on the pressure at which it is applied to the
instrument. Furthermore, some of the significant inputs to the calculation of pressure and its
uncertainty are operational, such as local gravity value and temperature, and so pre-tabulated
values of pressure generated can only refer to some specified standard conditions and
associated accuracies. This means that even for work of moderate accuracy, it is essential that
the pressure calculated and its uncertainty, take into account the value of the actual operational
parameters and their uncertainties at the time of use.

The pressure balance is a deceptively simple device, whose accurate usage is dependent on the
characteristics of many individual components that introduce considerable complexity and
inevitably require a reasonably complex analytical treatment. This can and does lead to a
further complication when calibrating a piston gauge since, as for any mathematical model,
there is a degree of choice in how to represent, process and report the critical measurands, and
even which parameters to include in the model, and which order of regression to accept. The
choices are often made on the basis of experience and subjective judgement rather than
systematic statistical tests. In a very open market, where users seek the most economical mix
of instrument configuration and type, supplier, and source of traceable calibrations for the
various components of the instrument, this can mean that the different modelling assumptions
cause some confusion and even cause systematic biases introduced by mixing models, both in
the value of the pressure, and in the value of the uncertainty calculated.

120
WALTER GIARDINI

In the calibration of one pressure balance with reference to another known reference
instrument, even when the same mathematical model is used, the convenient and widespread
use of mathematical software tools such as linear regression to calculate solutions can
introduce systematic differences when applied to real experimental data. This is a result of
implicit differences in weighting regimes within the regression models, a factor of which
calibration staff can sometimes be unaware. Fortunately these systematic biases are not critical
at the industrial and the general user level, but they certainly should be of concern to
calibration laboratories.

The field of "pressure" calibrations suffers from a somewhat loose terminology, because a
calibration can mean different things depending on the context. Pressure calibration can refer
simply to the pressure associated with a particular weight or set of weights on a particular
piston, under specified conditions. Even with this definition, there is some ambiguity, since the
complex calculation can be simplified in different ways, as for example in tabular form, or
using a simplified pressure summation table with an equation, or using the full equation giving
pressure as a function of all the input parameters. On the other hand, at the highest level,
pressure calibration means in practice determining the geometric and pressure dependent
characteristics of the piston-cylinder. One particular element of the calibration model which is
discussed in some detail below is the total force acting on the piston, which is only
determinable to within a small but non-zero fraction of a gram.

We must therefore clearly specify what the parameters of the "pressure" calibration are and
what the model of the pressure calibration is. Using the terminology of the ISO "Guide to the
expression of uncertainty in measurement" (GUM) [1], we must provide clear specification of
the measurands and, for pressure metrology, the measurands are as follows:

(1) the pressure output by a pressure balance at a defined level under operating conditions
which are specified,
(2) the corrections of a pressure measuring device under test, being calibrated using the
specified pressure balance, or
(3) the parameters A0.20 (effective area at zero applied gauge pressure and 20 °C and
generally referred to simply as Ao) and L (pressure distortion coefficient of the area)
determined for the piston-cylinder under test and the taring mass parameter (called the
effective mass me in Australia) associated with that calibration.

Note that in (3), the measurands become the inputs to (1) when the piston under test is itself
used as a pressure standard further down the hierarchy of standards.

When a piston cylinder assembly is calibrated by comparison with another, its parameters are
determined over a range of pressures. The question often arises as to what is to be taken as the
pressure range over which the test unit is considered to be traceably calibrated [2]. A simple
approach would be to state that the traceable calibration extends only to the region within the
calibrated range. On the other hand there is strong support from a great deal of knowledge and
data in the field for the hypothesis that the piston gauge area generally changes elastically and
according to a linear relationship over its operating range. At this point, the characteristics of
the piston-cylinder as an artefact must be differentiated from its operational application in

121
CHAPTER 5. THE HYDRA ULIC PISTON GA UGE

producing a steady pressure when operated in a pressure balance. In the latter context, at very
low pressures (for example with just the piston acting) issues of operability become relevant
such as reduced and insufficient length of spin-time due to low rotational inertia, and relative
accuracy of the load. At very high pressures, the question of validity of the linear behaviour of
the piston is relevant. In this case however, we need to differentiate between extrapolation
beyond the maximum traceably calibrated point up to the maximum design pressure of the
instrument, and extrapolation beyond the maximum design pressure.

5.2 Basic Theory


5.2.1 Models of piston gauge operation

The pressure defined by a piston gauge is modelled mathematically as: Mg/A. However, the
model is then modified and its analytical form changed because the way it is used depends on
the application. The piston gauge can be characterised from first principles, in which case its
effective area derives from dimensional measurements and calculations of the pressure
distortion from stress/strain analysis. On the other hand, it can be characterised by comparison
with another piston gauge, in which case, the ratio of loads, the linearity of the relative
pressure distortion behaviours and the weights of data points used to generate regression
results becomes the analytical focus. Finally, the end user may wish to calculate the value of
generated pressure with best possible uncertainty, or she may seek various levels of
approximations which ensure a known accuracy under specified conditions of use, and then the
trade off between complexity, skill level, algorithmic complexity, uncertainty and accuracy
becomes the problem to be solved.

Fundamental model of piston gauge pressure

The elements of a piston gauge are illustrated in Figure 5.1, and the input parameters and their
symbols and units are described below.

Note that the direction of loads is positive downwards. This is consistent with the head height
"h" (see below) being positive when the test unit is below the standard unit.

Ps = Pressure referred to reference instrument datum level. Pa


P = Pressure generated at test level. Pa
Pr = Reading of the pressure gauge. Pa
m = Total mass including piston components and extra added weight for balance, kg
Pn = The approximate pressure calculated from Pa
Pa = density of air, kg/m’
Pm = density of weights from conventional weighing, 8,000 kg/m^
Cf, (= l-pc/pm) = buoyancy correction factor, (pure number)
aA ~(Xp = area thermal expansion coefficient of piston-cylinder material, approximately
equal to the sum of the coefficients of the piston and the cylinder, °C ‘
t = temperature of the piston-cylinder assembly, °C
C, (=7-f<7a(/‘-20)) = Piston-cylinder assembly temperature correction factor, (pure number)

122
WALTER GIARDINI

Ao = Effective area of piston-cylinder at zero applied gauge pressure and 20 °C, m“


= Pressure distortion coefficient of piston-cylinder unit, Pa ’
nie = effective residual equivalent mass bias, obtained by regression, kg
r(= nieg) = residual load component obtained from regression, N
circ = circumference of piston, m
(T= surface tension coefficient of the pressure fluid, N/m
£ (=circxa) = surface tension force by oil adhesion at the emergent piston stem, N
g = gravitational acceleration N/kg
Pf = density of pressure fluid, kg/m^
V = Vp - V„, is the non simple piston volume for fluid buoyancy correction. Volume beyond the
simple piston envelope Vp is positive and volume below the simple piston envelope V„ is
negative, m^,
h = height difference between datum levels of reference and test instrument, measured
from the level of the reference to the level of the test +ve downward, m
Eg = Error of the gauge at pressure P, Pa
eo = implicit error of the gauge due to resolution limit or digital display limit. Pa
eo = implicit error of the gauge due to random variability. Pa
Cg = Correction of the gauge at pressure P, Pa.

Figure 5.1: Elements of a piston gauge pressure balance (see text).

A(p,t)

where F is the total load acting on the effective area A of the piston-cylinder assembly. The
area is a function of both applied pressure and temperature.

123
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

Over the range of pressures for which piston-cylinders behave elastically, and within their
operating range, their effective area is generally assumed (and experimentally found) to vary in
a linear manner with pressure. They also have linear temperature dependence. The linear
model that describes the change of area with pressure is written:
Ap =/Io(1-fAP)
but it is also written in the following form, (particularly in literature from NPL in the UK):
Ap = 6!, +a->P

where = A^ and tZj = A^A . In Australian practice, the parameter b is sometimes used for
X, as in the expression Ap = Ao(l + bP)

The total force F is primarily due to the dead weight of the piston and masses acting under the
influence of gravity and buoyancy due to the ambient air, with smaller corrections applied for
the surface tension of the fluid acting on the emergent stem of the piston, and further small
buoyancy corrections to allow for those parts of the piston which are immersed in oil and
either extend beyond, or form cavities from, the hypothetical envelope enclosing the simple
cylindrical surface of the piston of effective area A and extending to the lower face of the real
piston.

In British practice, the correction to the "fluid" buoyancy component is generally applied as a
direct correction, so that the lower face of the piston remains the actual datum level to which
the piston-cylinder pressure is referred. In US practice on the other hand, the positive
(protrusions) and the negative (cavities) parts of the piston volume are summed and added to
the actual piston volume, assuming the cross sectional area remains uniform, so that a "virtual"
piston is defined which is uniform in cross sectional area, but which has a length such that its
volume is equal to the volume of the actual (non-simple) piston. This means that the datum
level to which the piston-cylinder is referred, is slightly different in the two cases. Since the
pressure generated by 1 mm of oil fluid head is around 10 Pa, this is the order of systematic
difference that would be caused by this effect.

If the piston is of truly simple form, then the fluid buoyancy corrections are not relevant, even
though part of the piston is indeed immersed in oil under normal operating conditions. The
reason for this is that the measurand "gauge pressure" is strictly defined as the difference
between the internal and external fluid pressures in the system at the same level. When the
normal ambient air buoyancy correction is applied to the total mass of the system (which
includes the weights and the piston itself), the small error introduced due to the partial
immersion of the piston in the fluid just exactly cancels with the equal but opposite sign of the
correction required to obtain the value of the ambient air pressure at the level of the bottom
face of the piston. This subtle but important simplification simplifies the mathematical model
considerably. This is explained in detail in Appendix A.

Expressing the load and effective area terms explicitly in terms of the parameters discussed
leads to the general form:

124
WALTER GIARDINI

m(\-—)g+circX(j-Vg{Pf -pj
Pm (5.1)
Ao(l + /lP)(l + (6Z^+«J(r-2O))

where Ps is the pressure generated at the level of the base (bottom face) of the reference piston.

In this expression, ‘cz'rc x (f is the force due to the surface tension of the hydraulic fluid acting
along the circumferential line of contact of the piston stem where it emerges from the fluid,
and -Vg{pf- Pa) is the correction to allow for those parts of the piston which extend beyond
the simple cylindrical geometry, or are in deficit from that same geometry. It is emphasised
that this is a different effect to the one described above (and in more detail in the Appendix)
relating to the fact that the immersion of the piston in fluid, instead of air, on which the air
buoyancy calculation is based, still gives the correct result. In this case, the piston has actual
extensions to its form (or cavities) that are in the hydraulic fluid, but were calculated as if they
were in air, and hence this further correction must be included.

When the reference piston is used to calibrate a second device (either a secondary standard or
another piston gauge), the pressure is required at the datum level of the test device (for
example, the pressure port, or the base of the piston gauge under test). In this case a correction
is applied for the pressure due to the head of fluid between the reference level and the test level
and we obtain:

(5.2)
/lo(l + /lP)(l + «^(r-20))
where P is the pressure at test level datum, and is referred to reference pressure plus the fluid
head term.

This is the first point at which different modelling alternatives may have to be considered. In
European practice, calibration reports on pressure balances generally give only the values of Aq
and Z The user then has to calculate the extra loads due to the surface tension of the hydraulic
fluid acting on the emergent stem, and fluid buoyancy correction for deviations of the piston
from simple cylindrical geometry, and then perform the calculation according to the above
equation. In some reports the actual values of circxa and - VgiPt - Pa) for the pressure balance
are given for convenience. On the other hand, in Australian practice, these terms are generally
not calculated, but rather they are replaced by a "hypothetical" mass term rrie, whose effect is
equivalent to these loads. This term is generated by fitting the mathematical model to the
calibration data. In Australian practice, therefore, the equation looks like this:
m(l )g +m^g
------------------------------- + (P / - Pa )gh (5.3)
Ao(l + /lP)(l+cr^(r-2O))
where the explicit load corrections have been replaced by rueg. Note that the effective mass
term is not corrected for buoyancy. This is because its value is generated by a fitting algorithm;
it is not a "real" mass and is merely used to generate the load value when multiplied by g. A
natural extension of this procedure is to not weigh the piston itself, as it can be a delicate and

125
CHAPTER 5. THE HYDRA ULIC PISTON GA UGE

time consuming operation, and does expose the piston-cylinder to a small risk of damage. In
this case, the undetermined mass of the piston is included in the fitted mass value together with
the other small load effects, and the piston is then said to have an "effective" mass that
includes those effects. Where the mass of the piston is not included in this way, the fitted value
of nie would be expected to be small and equal only to the sum of those load effects left
unaccounted for. Where all known load effects are included in the mathematical model, the
value of rrie generated would be expected to be zero. Some calibrating laboratories (e.g., NIST
in the US), include this term as a "tare" load and do not convert it to an equivalent effective
mass term.

It should be noted however, that there may be a number of other small load effects present,
which cannot be adequately captured in the standard mathematical model. These include
variable loads from air movement and ancillary devices, stray magnetic and electrostatic fields
and mechanical effects due to the interaction between the rotating piston and cylinder such as
corkscrewing, aerodynamic forces on the rotating weights, relative position of the region of
engagement between the piston and cylinder and hydraulic fluid adhering to floating parts of
the system or collecting above the piston-cylinder. The load will also almost certainly include
an error term for the applied mass value, and an error term for the pressure head correction.
The method of fitting a "tare" component will therefore include all such load bias effects as
well as the determinable load effects shown above.

5.3 Use of the pressure balance


The piston gauge is a primary realization of the SI unit of pressure, directly defining a force F
acting over an area A. The parameters of high quality pressure balances of suitable design are
characterised directly, and then subsequent pressure balances are calibrated with reference to
the fundamentally realised standard by comparison. These pressure balances in turn are used to
calibrate pressure gauges of all types. This includes Bourdon tube pressure gauges,
transducers, calibrators and digital indicating gauges as well as other pressure balances of
lower accuracy. Some relevant documentary standards used are:

AS 1349-1986, Bourdon tube pressure and vacuum gauges


EN837-1998 Parts 1, 2 and 3. Parts 1 and 3 cover the metrology and testing requirements for
Bourdon tube and diaphragm/capsule gauges.
ASME B40.1-1991, Gauges-Pressure indicating dial Type - Elastic Elements.

The most desirable features of a pressure balance are its high precision and stability. This is
because both area and load can be very precisely defined. A very low and relatively constant
proportional uncertainty on mass value can be maintained over a large range, and the
uncertainty on the area acts proportionally on the output pressure, so that again, a very low and
relatively constant proportional uncertainty can be maintained over a wide range. Unlike many
other types of pressure measuring devices therefore, the uncertainty of a pressure balance is
generally expressed as a percentage of operating pressure, rather than of full scale. On the
other hand, the disadvantages of a pressure balance are its complexity of operation and
metrological analysis and the fact that it is not easily portable. These factors make the pressure

126
WALTER GIARDINI

balance particularly suitable for use as a pressure standard, to which lower level, less accurate
and less stable instruments can be referred.

A specialised use of pressure balances is the testing of pipelines to measure pressure changes
and hence detect leaks. This takes advantage of the very stable pressure generating ability of
pressure balances. In one such instrument used for gas pipelines, the gas pressure is applied to
the top of a fluid tube, itself connected to the pressure balance. The pressure at the piston is
maintained constant, but the correction to the pressure in the pipeline is configured to include
the visible oil column which contributes a pressure head term. By adjustment, the pressure
changes can then be measured in terms of the changes in the height of the oil column [2].
Pressure balances for pipeline testing conform to the requirements of AS2885.5

The basic process used is that a secondary pressure device is connected to the output port of
the standard pressure balance, and the known pressure compared with the pressure indicated
by the device under test. To achieve the best performance from this test, (reliable and accurate
results), requires that the standard and all associated equipment and software tools which can
affect the accuracy of the test be traceably calibrated, validated, maintained in good order, used
in accordance with valid procedures, and used by skilled personnel. The evaluation of the
uncertainty of the test is a critical element of the calibration, since any decisions made about
the compliance and adequacy (or otherwise) of the measurement results is only meaningful
with reference to the uncertainty in the value of the results. A detailed discussion of the
significant metrological parameters, which must be well defined, controlled and processed,
follows below.

5.4 Loads
The pressure generated by a piston gauge is defined as the ratio of the total force acting to the
effective area of the piston-cylinder assembly. The effective area is a complex parameter that
depends on the 3D form geometry of the piston and the cylinder along the engagement length,
the physical conditions of the fluid in the annular gap between the rotating piston and the
cylinder, and the changes in those conditions with pressure. At the level of a few parts in 10^
the effective area is generally regarded as the "critical" parameter of the pressure balance,
difficult to calculate with high accuracy and considered to be the traceable measurand that
must be metrologically controlled and managed. Mass, on the other hand, is one of the oldest,
best understood and most accurate physical measurements defined at the level of a few parts in
10^. Possibly because of this distinction, the load on a pressure balance has tended to be
neglected, and its satisfactory control regarded as a matter of good metrological practice. The
weights applied to the pressure balance are easily manufactured from high quality austenitic
(non-magnetic) stainless steel, should be handled carefully with gloved hands (in a metrology
laboratory at least), and can be routinely weighed with an accuracy of around 1 part in 10^.

The load, however, depends not only on the mass but on a host of other factors, which can
easily introduce significant systematic biases into the simple physical model These factors can
generate some components which are very difficult, if not (for practical purposes) impossible,
to measure and control. The load in fact is the component that has generated the greatest
variability in calculation and calibration algorithms developed and adopted by the various

127
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

National Measurement Institutes, manufacturers, and calibration laboratories. Fortunately,


small load effects become relatively less significant (say less than 5 parts in 10^) as the total
mass applied to a pressure balance increases above a few kilograms. The extrapolation of
calibration parameters down to the minimum pressure possible (that is with just the piston
rotating by itself) however becomes problematic, not only because of poor rotational
characteristics due to low rotational inertia and relatively high fluid friction, but because the
relative influence of the small and imperfectly understood load parameters becomes critical in
this regime. For these reasons, piston-cylinders are frequently certified only for use only above
a certain pressure.

Gravity

The first and most important effect is the conversion factor between mass and force, the value
of the gravitational constant g. This can be measured with an accuracy of a few parts in 10^ by
the geological survey departments of the various State Governments. Its value depends on the
mass of the earth, which is a global parameter, and locally on the latitude and height above sea
level, modified by local mass density variations. Values of g at many locations are published in
the literature (e.g.,AS 1349-1986), and a value obtained in this way, in close proximity to the
location of the pressure balance (both in terms of latitude and height above sea level and
furthermore assuming no major geological anomalies between the locations) can be
determined with an accuracy of around 0.01 % (see Figure 5.2).
Gravitational Acceleration vs. Latitude at Om and 1000m atxjve sea le\«l
1 vertical division = 0.05%
9.83500
9.83000 ^••'***
9.82500 7
9.82000
/7
9.81500
9.81000 7
9.80500
E
9.80000
9.79500
--------- 0
9.79000
........... 1000
9.78500
--------- Darwin
9.78000
9.77500 ........... Hobart
--------------------- -
9.77000
0 10 20 30 40 50 60 70 80 90 100
Latitude (Nor S)

Figure 5.2: The value of gravitational acceleration varies with latitude and height above
sea level. In Australia, the value of g (at the same height above sea level)
spans a range of over 0.2% between Hobart and Darwin.

Other than the effects of large variations in the density of the earth locally, the variation of g
can be modelled relatively well using standard equations, which take into account the latitude
and the height above sea level. The latter parameter needs to distinguish between the "free-air"
gradient of the density which would apply if say we were measuring the variation of g with
height in a balloon, or a building, and that which would be obtained if we were rising in height

128
WALTER GIARDINI

with the landmass itself by going up hills and mountains. Clearly, in the latter case, the rise in
height above sea level is associated with an increasing volume and mass of earth below that
height, which modifies the gravitational value in a different way. The free air correction is
given below, and can be useful also to calculate an exact value of gravity within a building, or
at different heights above the floor inside a laboratory, when a measured value is available at
one point. The second correction (when height rises with land mass) is called the Bouger
correction and is also given below.

Fortunately all these effects have been calculated and accounted for in standard equations,
some of which are given below. The equations all calculate the local gravity, relative to the
datum value in the international gravity network, referred to Potsdam. As the geophysical field
is usually expressed in units of milligal, these units are referred to where relevant. (Note:
9.80 m/sec“ = 980 mgal.) These equations have been tested at several locations where the true
value of g was known and found to give a value of g within 20 xlO’^ of the known value.

1 mgal = 10 '*’ m/s^

Free air gradient approx = -0.3086 mgal/m = -3.086 x 10'^ m/s"/m


Bouger term approx = +0.1118 mgal/m = +1.118 x 10'^ m/s‘/m

Yo is sea level g referred to Potsdam datum. Free air and Bouger corrections then need to be
applied. Formula (4), from the British Standard, includes the free air correction, however, and
only the Bouger correction needs to be applied.

(j) = latitude
H = height above sea level
Yo = 978.049 (1 + 0.005 2884 sin2(0) - 0.000 005 9 sin2(2(|))) (5.4)
Yo = 978.037 3(1 + 0.005 289 1 sin2((|)) - 0.000 005 9 sin-(2(}))) (5.5)
Yo = 978.049 6(1 + 0.005 2934 sin-((l)) - 0.000 005 9 sin-(2({))) (5.6)
Yo = 9.806 16 (1 - 0.002 6373 cos(2(|)) + 0.000 005 9 cos-(2(» - 3.086x10’^H (5.7)
Yo = 9.780 318 4 (1 + 0.005 302 4 sin-((t)) - 0.000 005 9 sin2(2<|)) (5.8)

Note: mgal units for (5.4), (5.5) and (5.6) and m/s^ units for (5.7) and (5.8).

Sources:
Eq. (5.4): International Union of Geodesy and Geophysics, 1924
Eq. (5.5): H.Jeffreys Mon.Not.Roy.Astro.Soc.Geophys.suppl.5,219 (1948)
Eq. (5.6): U.A.Uotila, Pub 7 Ohio State Univ. Inst, of Geodesy Photogrammetry and
Cartography, Columbus (1957)
Eq. (5.7): British Standard on Barometers (World Meteorological Organization) BS2520
(1983)
Eq. (5.8): International Association of Geodesy, 1970 (Also the formula in Kaye and Laby)

The use of equation (5.7) to generate gravity values is recommended as this is the most recent
and agrees well with equation (5.8), which is from a reliable source and is the next most
recent.

129
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

Buoyancy

The weights on a hydraulic pressure balance are immersed in the ambient air. For low pressure
gas piston gauges, it is possible to enclose the piston and weights in an evacuated chamber, so
that the reference pressure is a vacuum, but this is not generally the case for hydraulic units.

The density of the ambient air is around 1.2 kg/m^ and it will generate an upward force we call
buoyancy. The upward force of buoyancy is numerically equal to the weight of air displaced
by the object being considered. For example, a typical steel weight of 5 kg mass and a density
of 8,000 kg/m'^ has a volume of (v = rn/pm) 0.625 L or 625 cm‘\ This volume of air has a mass
of (m = paXv) 0.75 g, and therefore introduces an upward force equivalent to 0.75 g. In relative
terms this is an upward force of around 150 parts in 10^ and is certainly significant.

In practice however, steel weights do not have a density of 8,000 kg/m^, but rather closer to
7,900 kg/m^. This introduces another complexity that is well known to mass metrologists.
When weights such as those used for pressure balances are calibrated, it is done directly by
comparison against reference weights. Since the reference weights and the test weights are not
normally of exactly the same density, the buoyancy forces on the two objects will not be the
same. Two objects that have the same true mass (amount of matter) but different densities will
therefore not balance when the weighing is performed immersed in air. A further refinement
required is that the buoyancy force depends not only on the density of the objects being
weighed but also the density of the air in which they are immersed, which in turn depends on
its temperature, pressure and relative humidity. The consequence of this is that the
determination of the true mass of any object requires it to be weighed in a vacuum, or else the
densities of both the reference and the test weight as well as the current value of air density
must be known.

Mass metrologists have long ago found a convenient and practical solution to this problem,
which relies in part on the fact that most masses are used as weight references, that is used to
calibrate, weigh and generate forces in The mass of an object is therefore generally
calibrated, not as the true mass value, but rather as the mass value of a hypothetical object
whose density is exactly 8,000 kg/m'^ and which in air of density exactly 1.2 kg/m^ would
balance the test object. This works, since we are interested only in the downward force
generated by the test weight. We know that the hypothetical object would balance our test
object, hence we know that by calculating the force generated by the hypothetical object, this
must be equal to the force generated by our test object. The final issue is that the above is
exactly correct in air of density 1.2 kg/m\ but in air of different density, a correction has to be
applied.
For example, a calibrated test weight with an assigned mass value of 8,224 g on the "weight­
in-air" basis means that in air of density 1.2 kg/m^ that test weight would be balanced exactly
by a hypothetical object whose true mass value is 8,224 g, and whose density is exactly
8,000 kg/m\ The downward force F generated by the test object under these conditions is then
exactly the same as the downward force that would be generated by the hypothetical object,
namely
F = ni'(l-A,/8,000)g

130
WALTER GIARDINI

where m' is the value assigned to the mass of the object on the "weight-in-air" basis.

For a temperature range of 15 °C to 25 °C, an ambient pressure range of 98 kPa to 102 kPa,
and a humidity range of 20% RH to 80% RH, the density of air has a range of 1.18 kg/m^ to
1.22 kg/m^. For weights of density between 7,000 kg/m'^ and 9,000 kg/m'^ this is equivalent to a
variation in load component of up to ±3 parts in 10^ from the value obtained using the standard
equation assuming Pa = 1-2 kg/m^ exactly. For most pressure work, this is clearly a negligible
term. However, the trend for increasingly accurate instrumentation and calibration
requirements in the past decade or so, requires an awareness that these limits are at least
beginning to be relevant to some types of pressure users or work environments. Note that for
pneumatic instruments working in the absolute mode, that is where the weights are operating
in vacuum, and where the density of the weights is significantly different from that of steel
(e.g., aluminium), this effect causes significant biases, which must be accounted for (see
Chapter 4).

Fluid buoyancy on non-simple elements of the piston form

We refer to the Figure 5.3, which shows a piston engaged in a cylinder. The simple cylindrical
form of the piston, which produces the effective area along the full length of the piston, is
modified by two regions where a volume Vp of the material extends beyond the simple
cylindrical form, and where there is a cavity of volume VOv • In the first instance, the air
buoyancy correction is applied to all parts of the applied load including the piston, even though
the latter is actually immersed in fluid. As shown above however, (and detailed in
Appendix 5A) the correction required to refer the gauge pressure to the same level as the lower
face of the piston is just equal and opposite to the bias introduced by the above calculation, and
so the result is as required. However, this influence is only corrected to the extent of the
immersed parts of the piston, which produce the effective cross sectional area of the piston.
The volumes Vp and Vyv require further corrections.

Figure 5.3: A schematic illustrating that a piston often has protrusions and cavities
causing its form to deviate from a simple, solid, cylindrical form.

The simplest way to think of this is that the cavity region of the piston is a region that has been
included in the calculation of the buoyancy of the piston, but ought not to have been included.
Therefore, the buoyancy effect in this region must now be subtracted from the load
component. However, recall that buoyancy acts upward, which is the negative direction for
load, so therefore subtracting in the negative direction produces an arithmetical addition. In a
similar manner the extended region of the piston, ought to have been included but was not.

131
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

therefore it is added to the buoyancy correction and again since buoyancy acts in the upward
(or negative) direction, the net result is an arithmetic subtraction. The net volume to be
corrected for then is the difference between the positive and negative volumes or V = Vp - Vyv,
which generates the load term -Vg(pf- Pa).

Surface tension

Where the piston emerges from the cylinder, the hydraulic fluid forms a line contact around its
circumference generating a downward force due to the surface tension of the fluid acting on
the surface of the piston material. For hydraulic fluids such as Sebacate, Tellus oils of VG 22,
and even high pressure mixtures of castor oil and brake fluid, measurements of the surface
tension are quite similar and around 0.031 N/m. This is multiplied by the length of the contact
line to determine the force downward. For example for a high pressure piston of nominal area
2 mm‘ and circumference 5 mm. the force is 0.031 x 0.005 = 0.000 150 N, equivalent to a
mass of approximately 15 mg. On the other hand a large piston for low pressure use, of say
100 mm^ area and circumference 36 mm, produces a force of 0.001 116 N, equivalent to a
mass of approximately 112 mg, which is rather more significant. It should be noted however,
that in terms of pressure equivalent, 15 mg on the small area piston with pressure factor
5 MPa/kg is equivalent to 75 Pa, which at 100 MPa (20 kg) represents 0.75 parts in 10^
whereas 112 mg on a large area piston with pressure factor 0.1 MPa/kg is equivalent to
11.2 Pa, which at 2 MPa (20 kg) represents 5 parts in 10^. Some care therefore needs to be
taken to interpret the significance of the surface tension effect as a function of the operating
pressure.

A more difficult issue to deal with, however, is just exactly where to consider that the piston is
emerging from the hydraulic fluid. A number of pressure balance designs incorporate an
auxiliary piston along which the force of the applied masses is transferred down to the
metrological piston. The diameter of the auxiliary piston can be different to that of the
metrological piston, and it is possible in some instrument designs for the oil to collect above
the cylinder for some way, so that it is important to determine exactly which part of the piston
is emerging from the hydraulic fluid.
Wetting of the piston by the hydraulic fluid

It should also be noted that the piston does not in fact rotate at a constant height, but rather it
falls and is raised as the volume is adjusted by the user, so that it might be expected to pick up
and retain some of the hydraulic fluid, therefore changing its effective mass by some small
amount. Tests of piston weight after dipping in fluid to various levels to test this hypothesis
show that the mass change is generally negligible but, again depending on the oil used and the
pressure range being considered, this factor may be relevant in some instances.

Build up of hydraulic fluid above piston


The oil that can accumulate above the cylinder exit also causes a small pressure head there,
which needs to be taken into account in the calculation. Most pressure balances have a relief
hole close to the cylinder to carry the oil away. However, in some piston-cylinder designs the
hole can be several millimetres above the cylinder exit. For some dual range units a diagonal

132
WALTER GIARDINl

hole to carry the oil away combines with the rising and falling piston to generate a pumping
action, which can generate a variable oil pressure head. The use of retaining nuts in some
pressure balance designs can also form a well that generates a short oil column.

Pressure head to calculate pressure at different levels

The pressure head to allow for the difference in height between the calibrated reference level
of the pressure balance (generally the lower face of the piston or the pressure port outlet), and
the height of the device being calibrated is calculated by the usual expression pgh. This is
influenced, however, by the variable working height of the piston over around ±1 to 2 mm.
Since each mm of oil generates a head of 10 Pa, this sets a lower limit to the accuracy of the
pressure. A practical solution is to both calibrate and then use the pressure balance at a very
accurately controlled height, say ±1 mm. One manufacturer uses the position and fall rate
sensor to determine the exact height at the time of pressure measurement in order to calculate a
more exact value for the pressure head. A systematic bias in the pressure head will generate a
fixed bias in terms of pressure, seeming like a mass bias.

Rotational effects - aerodynamic and mechanical

High speed rotation of the weights in air can generate aerodynamic lifting or downward forces
on the weights. This is only a significant issue for high speed rotation usually associated with
pneumatic instruments. Hydraulic instrument weights spin at rates of 30-40 revolutions per
minute and so this effect is negligible. However, if the piston and/or cylinder are scoured by
helical grooves, which can be caused by dirt particles in the hydraulic system, particularly in
industrial environments, this can generate lifting or downward forces. The presence of these
forces can be determined by noting the difference in output pressure, when the weights are
rotated in the clockwise and counter clockwise directions. At calibration this can and should be
determined by a sensitive crossfloat against a second pressure balance, but, in principle, a high
sensitivity pressure transducer could be used to do the same thing.

Levelling

The weights generate a force in the direction of the gravitational field, this being the true
vertical by definition. If, however, the piston and cylinder are aligned at an angle to the
vertical, then only the vertical component will act on the area. The actual load acting in the
direction of the area is then modified by the cosine of the angle to the vertical. An error of 0.5,
0.2 or 0.1 degrees generates a bias respectively of 39, 6 or 2 parts in 10^.

Magnetic and Electrostatic forces

These forces are quite difficult to characterise and the best strategy is to try to minimise them
by avoiding sources of strong magnetic fields, using non magnetic weight components
wherever possible, and avoiding the use of materials which can build up electrostatic charges.
The use of perspex sheets to create draft shielding enclosures for example needs to be applied
with care. The degree to which weights are sensitive to magnetic fields can be tested by
placing the weight on a sensitive balance so that it is well away from the balance mechanism

133
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

itself (for example by placing it on top of a long wooden column) and then using a magnet to
approach the weight and see if this changes its apparent weight. A gauss meter can also be
used to determine if the weights are magnetised.

Height indicating devices

Some pressure balance designs include finely balanced lever systems that move up and down
with the weight column via a low strength magnetic link, and indicate the height of the
column. Apart from the possible influences of the low magnetic field, unless the lever system
is perfectly balanced, it will inevitably generate a load effect. These types of devices while
convenient, are not suitable for the highest precision work.

Motorised rotation of the weights

This is generally achieved by a rotating wheel actuated via a drive belt on which a vertical pin
is placed that hits and drives a rolling wheel attached to the weight column. When the pins
make contact, there is inevitably a vertical frictional component to the impact imparted to the
weight column, which will disturb the load value dynamically and require a short time to settle
down again. Some systems operate by more or less continuously impacting, and these would
significantly degrade the accuracy of the instrument and the stability of its pressure output.
One system operates using an oval pulley for the drive belt, so the rotation of the driving wheel
is non uniform and therefore creates only isolated impact events with significant intervals of
time of free rotation. One such system was successfully modified by an Australian laboratory
so that the drive wheel could be driven directly at different speeds under computer control.

There is not general agreement on whether it is more desirable to use the driving motor for
high quality work. Certainly for lower precision work the driving motor is far more
convenient, but even for very high precision work, the effect of grabbing the weight stack by
hand and rotating it, even when done with great care can impart sideways forces which cause
the weight stack to swing slightly, and providing measurements are not taken very close to the
point of impact of the motor driving mechanism, these can leave the system in general less
disturbed. On the other hand, the use of a motor introduces a source of heat close to the
pressure balance which disturbs the thermal equilibrium. On some instruments this is
addressed in part by locating the motor well away from the drive column and using a long
drive belt.

Air draughts

This is another influence factor which is quite difficult to characterise, but which can at least
be readily assessed and abated. The pressure balance should not be operated in proximity to
sources of strong drafts, in particular air conditioning vents, windows and doors. If this cannot
be avoided, then some form of shielding should be used. Note, however, the comments above
on using plastic sheet, which may generate electrostatic forces, and also note that excessive
isolation and thermal barriers can have a negative impact on the thermal stability and control
of the pressure balance for best accuracy work.

134
WALTER GIARDINI

5.5 Operational considerations


Piston-cylinder materials

Most piston-cylinders are made of tool steel or tungsten carbide. The advantage of the latter is
that the linear thermal expansion is less than tool steel (4.5 x 10’^ per C° as against around
11x10 so that sensitivity to temperature is considerably less, and the stiffness is
considerably higher, so that the distortion of the unit with pressure is considerably less than
that of tool steel (for a simple piston design a change of area of around 1 part in 10^ as against
around 3 parts in 10^). Tungsten carbide on the other hand is more brittle than steel, and for
high pressure pistons which are quite slender (diameters down to around 2 mm), some
manufacturers use a combination of steel pistons and tungsten carbide cylinders.

Design

The problem of calculating the distortion of the piston-cylinder, and ensuring that it has a well
characterised value also influences the design of the piston. A very simple piston and cylinder
assembly, simply and repeatably attached to the hydraulic system, is desirable. Industrial level
units on the other hand, sometimes combine the metrological working cylinder with the
hydraulic housing element. In this case, the housing can be attached to the rest of the hydraulic
system with a variable torque, and this can affect the stress distribution of the system under
pressure, and hence the distortion characteristics.

A simple piston-cylinder will generally deform under pressure so that the piston is "squeezed"
by the fluid, and the cylinder is "pushed out" by the internal pressure, causing an increase in
effective area, or in other words a positive pressure distortion coefficient. If the cylinder seals
are placed high on the external wall of the cylinder so that the pressure fluid also acts on the
external walls of the cylinder as well as the internal bore (called a re-entrant design), then the
effective area changes by a smaller amount, and can even decrease with pressure.

Cleaning, the spin-down test

Whilst many users avoid removing the piston-cylinder from the pressure balance, to avoid
possible damage, it may be necessary to do this if the unit becomes dirty from contaminated
hydraulic fluid. The spin-down test is a simple and reasonably repeatable quantitative test,
which will indicate if the piston-cylinder has been damaged or is excessively dirty and needs to
be cleaned. This is performed by applying a relatively low amount of weight to the system to
generate perhaps around 5% of maximum pressure, and measuring how long it takes for the
weights to spin down to a full stop in free rotation, from an initial rotation rate of around 30-
40 revolutions per minute. A more qualitative yet very sensitive test is to carefully inspect the
character of the rotation at very low speed (the equivalent of 1 rotation per several minutes). If
the weights slow down to a full stop very gradually and uniformly, this indicates a very clean
and optimally working piston. On the other hand, any type of "grabbing" or non-uniform
deceleration indicates some attention may be required.

135
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

There is a technique to removing, cleaning and replacing a piston unit, and it is not
recommended that this process be undertaken without some training. The actual operation is
quite straightforward, but does require that the unit be handled as a very delicate metrological
device, rather than a piece of precision mechanics. We have found that washing under warm
water with mild soap, rinsing well, and finally rinsing with distilled water, and possibly
alcohol to remove any water and moisture, can be very effective. If washing with water
however, care must be taken with some types of piston-cylinders which have cavities to ensure
that water does not become entrapped in a poorly sealed cavity. This would not only affect the
mass of the piston, but also possibly contribute to corrosion in the long term.

Pressure factors

In Australia, piston-cylinders have traditionally been identified by their nominal effective area.
Thus we have had a 1/8'*^ sq. inch piston and so on. A more recent trend, particularly from
some European manufacturers has been to characterise pistons not only by their area, but also
in terms of the mass under which they operate. This suggests that a "pressure factor" term is a
useful way to talk about piston-cylinders. A piston of effective area 4.032 mm' then becomes a
piston of pressure factor 2.4 MPa/kg. This means that, on this particular piston, each kilogram
of mass applied (under nominal standard gravity) will generate 2.4 MPa of pressure. This is
particularly convenient, as 2.4 MPa per kg, would be the same as 2.4 kPa per gram, and 2.4 Pa
per milligram, by successive division by 1000. Some manufacturers make piston cylinders in
series that are numerically equal to simple fractions of the value of standard gravity (e.g.,9.80
mm\ 4.96 mm', 19.6 mm' and so on) so that whole number kilograms are nominally
equivalent to simple whole number and ratio pressure factors (e.g., 1 MPa/kg, 0.5 MPa/kg,
2 MPa/kg and so on).

Hydraulic fluids

Different hydraulic fluids are used depending on the pressure range and the characteristics of
the piston unit. Water as well as oil is used, the former particularly for use with oxygen
gauges, which must not be contaminated with oils as even small traces of oil may form an
explosive mixture with oxygen under pressure. For industrial instruments, in which the piston
cylinder gap is relatively large (a few micrometres), reasonably viscous fluids (VG15 and
VG22) have been traditionally used. For very high pressure use (up to 500 MPa) these oils are
not suitable, as they become very viscous. A mixture of castor oil and hydraulic brake fluid is
used with one common instrument for pressures up to 400 MPa. For instruments with very fine
piston-cylinder geometries and close gaps (of the order of 1 micrometre or less) sebacate and
similar (DOS, Dy-octyl sebacate) oils are used. These have relatively low viscosity at low
pressure, and retain good fluid properties to 500 MPa and beyond.

If instruments using different oils are calibrated an interface may be used. This can be a
loosely fitting membrane that separates the two fluids inside an interface housing, or a thin
plate used as a separator closely constrained in a housing with sensors and a fiducial indicating
device to detect pressure equality across the plate. Alternatively, a longer connecting tube can
be used with multiple valves and a bleed line, so that the two oils can be alternately released to
ensure that one side does not contaminate the other.

136
WALTER GIARDINI

Connectors

For high pressure work up to several tens of MPa, bonded seals (also called Doughty seals) can
be used, and up to several hundred MPa, coned fittings can be used to make connections. The
total volume of connecting tubing should be kept as small as possible. This is good practice
from a safety viewpoint, since the amount of energy in the system is minimised. It is also
important metrologically, since the system is less sensitive and will have longer time constants
for larger total volumes, and operationally, since the purging of entrapped air and cleaning of
the system will be facilitated by a small system volume.

Performance checks

As well as the spin-down time described above, the fall rate of the piston cylinder at maximum
pressure should be checked at regular intervals. This can be achieved with sufficient accuracy
simply using a stop watch and a height micrometer or even a good quality ruler. At maximum
pressure the weights are set spinning at normal rotation rate (around 30-40 revolutions per
minute) and the fall rate in mm/min should then be measured. A fall of around 2-3 mm should
be convenient to measure with sufficient accuracy. The piston gauge should fall at less than
1 mm/minute to ensure that it can be used to calibrate a pressure instrument under relatively
stable and well defined conditions. Any large changes may indicate wear of the piston. Of
course a large fall rate can also be caused by a small undetected leak in the system, and this
should be eliminated as a possible cause before investigating the possibility of a change to the
piston-cylinder itself.

Calibration intervals

In Australia, a group of experts from industry, NML and suppliers investigated available
histories of many pressure balances calibrated over long periods of time, and used in different
environments, and recommended to the National Association of Testing Authorities (NATA)
that a recalibration interval of 3 years would be appropriate for instruments whose uncertainty
of measure is 0.01% or better, and a recalibration interval of 5 years would be appropriate for
instruments whose uncertainty of measure is more than 0.01% and less than or equal to 0.05%.
These are the current recalibration intervals recommended by NATA, supported with regular
checks of fall rate and spin-down time.

5.6 Calibration of pressure measurement devices


In using a pressure balance, a pressure measurement device is connected to its output port and.
under operating conditions, the pressure generated at the port is compared to the device
indication. The main standard used for calibration of pressure gauges in Australia is AS 1349-
1986, Bourdon tube pressure and vacuum gauges. Whilst there is no formal standard
addressing the calibration of other types of transducers generically, EN837-1998 covers the
metrology and testing requirements for Bourdon tube and diaphragm/capsule gauges. Other
pressure measurement devices, including self-indicating transducers, transducers with current
and voltage output and calibrators, are calibrated wherever possible in a way which is
consistent with the requirements of AS 1349. Two pressure runs in increasing and decreasing

137
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

directions following a run to exercise the instrument would be recommended, and calibration
at 10 points over the range of the instrument, including the zero and maximum readings.

In the case of a Bourdon gauge or a self-indicating gauge such as a digital gauge, the
measurand is the correction determined at each pressure point according to the equation:
Pressure = Pressure reading -t- Correction
and the uncertainty of the calibration is the uncertainty in the values of the corrections so
determined. In the case of transducers with electrical output, the calibrated measurand can be a
correction, where the electrical output is scaled to the numerical value of the pressure, or can
be stated as a mathematical relationship between the signal level indicated by the instrument
and the true pressure acting at its port. In that case an equation relating the reading of the
transducer to the pressure can be developed. An uncertainty for this pressure generated by the
equation is based on the output of the transducer.

Gauges that are used for oxygen must not be calibrated with oil, and no oil of any sort must be
allowed to contaminate such gauges. This is because traces of oil and oxygen under pressure
can generate explosive mixtures. Such gauges can be calibrated using water as the hydraulic
fluid. Water operated pressure balances are used, or oil/water separators. Note that in the latter
case, the pressure drop introduced by the separator should be determined.

A full development of the uncertainty calculation for the reference pressure, and the
uncertainty in the value of the calibration results (the determined corrections of the device), is
given in section 5.8.

5.7 Cross float comparison calibration of piston gauges


Data reduction

When comparing two piston gauges, it is convenient to compare the ratio of loads to the ratio
of areas. The pressure head term in equation (5.2) can be converted into a "load equivalent"
term by multiplying by the nominal area S of the piston gauge and, thence, be placed in the
numerator, yielding the following form:
m(l - ^)g +CC7- Vg(pf - Pa) + {pf - Pa
P =------- ---------------------------------------------------- (5.9)
4(l + /lP)(l + (6z^,-F«J(r-20))

The pressure at the datum of the test instrument can also be described in terms of the test unit
parameters using the equivalent formulation:
m' (1 - ^}g -h circ'xc7'-V' gip^ - p^^)
Pt (referred to the test instrument) =-------- -------------------------------------- (5.10)
Ao’d + a’P)(l -h (cir/+cr/)(r-20))

138
WALTER GIARDINl

In this case the quantities are all primed (e.g., to indicate that they refer to the test unit.
Note that this assumes the same fluid acts throughout the system, and the same air density and
gravity applies to both instruments, hence these parameters are not primed. Note also that there
is no head correction term in this case, since the reference pressure calculation contains this
term to refer the reference pressure to the test unit datum.

When the two instruments are balanced against each other in a cross float experiment, the
pressures at the test datum level as expressed by the reference parameters and as expressed by
the test parameters are equated, hence:
Pt = P,
which leads to:
m' (1 - + £'-V' g{pf - ) m(l - - pJ + }ghS
--------------------------------------------------------------------------------------------------- (5.11)
A)' (1 + A’ P)(l + (a,, ' )(t'-20)) Ao (1 + /IP)(1 + - 20))

This is the fundamental theoretical equation describing the relationship between the parameters
of the reference and the test instrument when they are balanced and in equilibrium. There are
two unknown parameters A o and A,' of the instrument under test. Multiple experimental points
at which the equation holds (i.e., pressures are balanced) are determined and A'o, /i' are
obtained by least squares regression. This is described as the theoretical equation, because in
fact, it is not found to hold over all pressures. In practice, although unusual, the effective area
is sometimes found to vary in a non-linear manner, and a further term is included in the
expression for its functional dependence on pressure as follows:
A = Ao(l-h/t,P-h22^2)
More importantly and far more common, however, is the potential presence of small forces
which contribute to the load acting on the piston, but which are too difficult to characterise
precisely and incorporate formally into the theoretical model. These include possible
"corkscrewing" forces due helical scratches on the piston under rotation in the hydraulic fluid,
the effects of oil adhering to the piston’s upper portion as it alternately immerses in and
emerges from the oil, possible stray magnetic forces acting between the loads and the
environment (including parts of the pressure balance base itself), air drafts, oil buildup above
the piston not completely drained or with variable draining behaviour, the use of magnetically
operated follower levers to indicate the operating height of the piston, aerodynamic effects on
the large surfaces of fast spinning weights, and possibly others as yet undiscovered. Clearly,
some of these load bias effects may be systematic, and others variable, either randomly or in
some repeatable fashion. For example, the true value of the weights applied to the piston will
not be the value reported, and there must inevitably be a small error (expected to be less than
the uncertainty) present. Since each pressure value uses a different set of weights, the bias at
each point can therefore expected to be not constant, but certainly repeatable (given the same
set of weights is used).

In particular we should note that, as the head correction between the reference and the test
instrument can be considered an effective "load" term, an apparent bias in the load can also be
produced by a bias in the value of the constant pressure head. In practice however, when two

139
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

piston gauges are cross-floated, the two pistons have a range of movement of some few
millimetres (around ±2 mm from the mid position is typical),therefore the exact value of the
head, being the difference between the two lower faces of the pistons, can vary by about the
same degree and will cause a variability in the conditions of balance. As 1 mm of oil produces
a head of around 10 Pa, a variability of the order of 10-20 Pa might be expected from this
source alone.

Finally, a fixed systematic error in the reference pressure will also be expressed as a "load"
bias since it can be thought of as an equivalent fixed mass bias on the reference piston.

It is clear then, that the load bias is a parameter that is dependent not only on the characteristics
of the piston under test, but also on the calibration conditions and the characteristics of the
reference piston itself. In the past, the "taring bias" has been attributed solely to the unit under
test, in the belief that the bias could only be due to some feature of its operation, but we see
that, in general, it is more appropriate to treat the taring bias as a parameter of the calibration.
On the other hand, it is convenient and efficient to model the taring bias as a term associated
with the unit under test only, in the form of a force bias, or as is common in Australian practice
an "equivalent mass" bias. In the past this has been called the "effective mass" term rUe,
because of the common practice of not weighing the piston itself, but rather evaluating its mass
and the effects of any systematic load effects in combination. Hence this produced the
"effective mass" of the piston.

The recommended practice now is to weigh the piston separately, and to account as far as
possible for all known forces, including surface tension, non-simple piston volume fluid
buoyancies and so on. It might be thought that under these circumstances, there is no need to
include the "mass bias" term in the mathematical model, as there should now be no such bias,
and to the extent that any bias is present, it ought to be rigorously included in the uncertainty
component arising from the loads. However, it has been found by high precision cross float
experiments that, even when all known loads are accounted for, there remains a small but
statistically significant taring term, which at this time we cannot explain.

The current situation is that the mathematical model including the taring term appears to be
accurate and consistent with the characteristics of the experimental data obtained. By allowing
explicitly in the mathematical model for the possible effects of a force bias, we are able to
obtain estimates of the effective area, and the distortion coefficient to high degrees of
accuracy. In effect, we use the pressure balance cross-float to achieve a high accuracy
measurement of a geometric parameter (the effective area). Including the mass bias term
allows a much better measurement of the geometric parameters, since the results do not
contain the "hidden" effects of the mass bias (which is explicitly accounted for). We must be
aware however, that when that piston is used to generate pressure, notwithstanding its high
geometric accuracy, the mathematical model used to generate pressure must contain what
might be better called its "zero" load error which is associated with its calibration (and hence
includes not only its characteristics but also those of the calibration system at the time of
calibration, and the bias possibly present in the reference standard.)

Rearranging equation (5.11) we obtain:

140
WALTER GIARDINI

[/w’(l-^)g + £-'-V'g(/7. -/7j]


n (1-F(6Z +6fJ(t-20))
4'(1 + AT) =---------------------------------------------------------------------------------- —------ -.4(1 + /T)
r A, z AZ A zci (l + (cr '-Fcr’)(z'-20))
[w(l----- ^)g + £-vg(p^ -pJ + (Pf-pJghS] P c y
Pm
where the left hand term is just the effective area of the test unit at pressure P, A'p, and is
rewritten;
[m'd-— }g+£'-Vg(p, -p„)l
A,'=------------------------------------------------------- (l+(a„ +a,)(z 20))^^^^!
[m{\-^)g + e-Vg(p, - + - pJghS] O + 20))
Pm

The value of Ap' is plotted over the pressure range, and hence the functional dependence of the
area of the test instrument is determined experimentally. A straight line dependence of Ap’ on
pressure is expected, and assumed. This is made clear in the explicit form of the mathematical
relationship Ap'= Ao(J+-^P), as above.

However, in practice, apart from the expected scatter of experimental points, it is found that at
very low pressures (where small loads are applied to the instruments), the data points for Ap'
begin to curve either upwards or downwards. It is not expected that at very low pressures, the
behaviour of the area should be non-linear in this way, and it is instead assumed that the
behaviour is caused by a small systematic load offset. Note that the form of fitting equation is:
W’
A; = — .Ap

where the temperature correction terms have been incorporated into the loads W and W. The
value of Ap’ is clearly functionally dependent on the ratio of the loads, and will be relatively
insensitive to a small systematic bias in either load for large values of W and W, but will be
considerably more sensitive at small values of W' and W (i.e., small pressure values). It is also
known that, particularly for industrial pressure balances (i.e., with non-simple and more
complex piston designs), there may be small load terms that are difficult to specify completely,
as already described above.

Because of this potential small and systematic indeterminacy in the applied load (which is
usually however made clearly evident by a systematic non-linearity in experimental data) a
modified mathematical model is fitted in which a small systematic load term is included and
determined by regression. In Australia, this indeterminate load has historically been thought of
as an "effective mass" term "me", being an equivalent systematic mass bias which would
generate the experimentally determined load bias.

W'+m s
----- -^.A (5.13)

An alternative, but less desirable, way of dealing with this would be not to include an effective
load bias in the mathematical treatment, and directly fit a simple straight line to the

141
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

fundamental theoretical equation above. In this case, the low pressure experimental points
which deviate significantly from linearity would be discarded from the regression fit, and only
the higher pressure points included. Because a small systematic load bias has very little effect
at high pressures this would introduce only a very small bias in the calculated test unit area. As
an example of this, a typical systematic load bias of 250 mg at higher pressures where total
masses of say 50 kg are used is 5 ppm. This is negligible in all but the very best quality
calibrations.

On the other hand, if an rrie term is included in the analytical model, but the experimental
results indicate no discernible bias, then whilst a non zero value of mg will in general be
obtained due to the always present noise of the system, its statistical significance (null
hypothesis mg = 0) can be readily and rigorously determined.

There are contexts in which some of these automatic corrections are appropriate, useful and
relevant, such as relatively low level calibrations for near-end users of pressure balances. In
these cases, the inclusion of the "mg" term applies a partial "black-box" approach to the
calibration of the pressure balance. In effect, the use of the nig term in the pressure generating
equation is used to generate pressure numbers which are closer (with smaller residual errors
due to the fit) to the reference pressure as defined by the standard instrument at the time of
calibration. This allows the end user to obtain the best value for the pressure generated. If for
example, one of the instrument masses contains an undetermined systematic bias then, the
treatment of the experimental data including the nig parameter, will ensure that a value of nig is
generated which allows just the right amount of counter-bias to generate a best-fit value of the
pressure.

It would appear then that there is little to be gained other than convenience by ignoring a
possible systematic load bias term, and much to be gained by routinely including it in the
formal analytical model. This conclusion, however, must be tempered by a strong warning.
The inclusion of the "nig" term in the analytical treatment can be used to allow mathematically
for small load terms which are difficult to calculate, such as the fluid buoyancy terms for non­
simple piston geometries, or indeed to evaluate the whole mass of the piston "effectively" by
including it as the major part of the nig term. However, the fitted term will also automatically
"correct" for any errors in mass or load measurement. Thus in adopting this partial "black-box"
approach we are in danger of missing significant errors in loads.

Furthermore, the assumption that the systematic bias can only be due to, and is only a property
of, the test instrument is simply not true. A systematic error in the reference pressure such as
might be caused by an error in the measurement of the fluid head between the reference and
the test datum, or indeed a systematic mass bias in the loads applied to the reference side, will
all be expressed through the value of nig determined by fitting the model of equation (5.12) to
the data.
Thus it might be appropriate to associate me not just with the test piston, but rather treat it as a
property of the calibration itself. Certainly the major component of nig could be reasonably
assumed to arise from the test piston and its load components since its behaviour, quality,
complexity and accuracy of characterization is generally of a lesser level than the reference

142
WALTER GIARDINI

standard instruments, but we should be aware that this is not a necessary consequence of the
fitted model.

Insofar as the purpose of a calibration is to establish the traceability of a test instrument to the
national standard with a given level of uncertainty, it also follows that if indeed there is a
systematic bias in any component of the national standard which expresses itself through an nie
value assigned to the test unit, then this uncertainty will flow through to the end user, and
properly reflects the uncertainty intrinsic to the national standard of pressure.

This last point needs to be developed a little further. In calibrations at the highest level, the
parameters of a piston gauge which are determined are A'o and A, with their associated
uncertainties. Because there is a well understood relationship between load and pressure which
is assumed to hold, it is possible to determine the area parameter (effectively A'o , /I) by a best
fit equation to the data, such that the systematic load bias me can be adjusted to give the best fit
to that assumed model. Thus the values of A 'o and 2 for a particular piston can be determined
with relatively low uncertainty. However, due to the presence of undetermined small load
biases, the relatively poor behaviour of piston-cylinder assemblies at low pressure (due to low
rotational inertia, the relatively greater influence of friction and hence rapid deceleration), and
the greater proportional uncertainty in the loads applied, the accurate values of A 'o and do not
translate into accurate values of pressure generated for small values of loads. This is
effectively the lower limit of operation for piston gauges, and metrologically it translates into
increasing proportional uncertainties as the fixed uncertainty limit in absolute pressure units is
approached.

A'o describes the static condition of the piston, when it is under no operating pressure. At low
values of applied load, the term me (or rather its uncertainty) representing the basic
indeterminacy of applied load measurement under operational conditions limits the accuracy of
pressure measurement to the fixed amount (u(me) xg/A'o). At high pressures, the uncertainty
in Z translates into an equivalent (proportional) uncertainty in A'p that grows larger in absolute
pressure units.

Considering the critical piston calibration parameters A'o, 2 and nie generically, it is now clear
that they constitute a necessary triplet to describe the traceability of a piston-cylinder over its
operating range. The inclusion of nig in the model therefore is more than an attempt to force the
data to fit the theory, or to make up mathematically for a lack of experimental capability, but it
is actually the necessary input which conditions the uncertainty in output pressure due to load
indeterminacy.

It must also be remembered, that whilst the assumption is made that there is a potential
systematic and constant load bias throughout the pressure range, in fact if there are different
bias components acting at each pressure, or indeed if one pressure point only includes a
significant error, then the value of nig will be sensitive to these conditions.

Taking all this into account, and being aware of the limitations on interpretation of the
resulting parameters, the use of the fitted parameter trig is recommended in the analytical
modelling and processing of piston gauge calibrations.

143
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

For low accuracy or industrial instrument calibrations, the me term can be used to generate the
"effective" or fitted value of some of the load components which are inconvenient or very
difficult to characterise, or it can be used to partially "black box" the instrument so the pressure
value generated by the fitted equation is optimum with respect to the reference standardizing
instrument (in effect the link to the national standard). Care must be taken however, that the nie
value does not conceal significant load or pressure errors that the calibration experiment is
intended to address.

For best accuracy calibrations, it is recommended that ah known or determinable load effects
be calculated in full and the mg parameter used as a diagnostic to determine the level of
indeterminacy of the loads at the time of calibration. At this level, the uncertainties in A 'o and A
are quoted, but so must be the value and uncertainty in nig, which is a significant fixed input to
the uncertainty in the pressure generated, proportionally becoming dominant at the low end of
the pressure (load) range for the piston-cylinder in operation.

With the inclusion of the term me, the fundamental equation (5.11) becomes:

--^)g +£'-V'g{p f - +
-------- ---------------------------------------= p, which, on putting
Ao’(l + /l'P)(l-h(6r/-Fa/)(r'-2O))

C, = (1 + {ap '-F6Z/)(f-2O)) and

W= zn'(l --^)g -l-f’ - V'g{p. - pA becomes:


Pn,

W'= -m,g + A,’[c; P] + ^'A,'[C\ P^]

This equation optimises the fit of the applied load against pressure using a parabolic equation
in the general form Y = ao + ajXj +02X2. We introduce the term uncorrected load at this point,
to indicate that this is the determinable load, which is not corrected for the possible presence of
a systematic "taring" effective mass nig.

Where the dependent variable is:


Y = Total (uncorrected) Load = W = ni'( 1 - pApA) + P + Y'g(pf - pP

and the independent variables are:


X, =C\PwnAX2 =C',P^.

The regression fit parameters are:


Qo = -nigg hence mg= -aAg and standard deviation u(me) = -u(ao)/g
Cl! - Ao' hence Ao = cii and standard deviation u(Ao') = u(ai)
02 = AAo hence A' = a2/Ao - a2/ai and standard deviation u(b') = u(a2)/u(ai).

A graph of W against P would appear as very nearly a straight line (the squared term being
very small), which intercepts the load axis very nearly (but for me) at zero and whose slope is

144
WALTER GIARDINI

the area of the piston cylinder. As the load varies over the full range (say 1 kg to 100 kg), the
scatter of the data points (of the order of 100 mg) would be insignificant at this scale and
visual inspection would show nothing more than a set of points in an apparently perfectly
straight line.

For the above reason, an alternative form of this equation can be used and is obtained by
dividing through by the pressure to give:
W' -m,g
+ A'o+A'Ao'P.
C\ P C, P
Here, the dependent variable is
Y = W'/C't P
= Area of test piston using uncorrected (i.e., no nig term) load
= "Uncorrected Area"

and the independent variables are:


X, = l/(Ct P) and X. = P.
The regression fit parameters are:
ao = Aq , hence Ao = ao and standard deviation, u{Ao) - u(ao),
a I = -nigg , hence, me= -ai/g and standard deviation, u(me) = u(aj)/g
a2 = AAo', hence A' = a^/Ao - a^/a and standard deviation, u(A') = u(a2)/u(a]).

This equation plots the (uncorrected for rrie) area of the test unit against pressure. Technically,
this form of data presentation is very much more useful than the W' vs P equation and plot for
the parabolic equation. First of all, if mg= 0, then the uncorrected area W'/(C't P) is actually
equal to the true area A 'p and the values of A 'o and A are obtained simply from the intercept and
slope of the plotted line. Second, because the area varies by about the same order of magnitude
as the data residuals (total change in A'p over the full pressure range is typically 100 ppm, and
residuals are typically 5 ppm), then the data plot is visually meaningful and can be inspected to
assess the experimental performance. Third, if nig # 0 then the hyperbolic term becomes
significant at low pressure values, and the straight line curves strongly up or down depending
on the sign of nig. This becomes a critical diagnostic feature of this equation.

There are two effects that influence the load data values at low pressures. Firstly, there is the
influence of a systematic load bias characterised by nig, already discussed. This will tend to
introduce a strong hyperbolic curvature into the straight line at low pressures. However, due to
poorer spinning and greater difficulty of obtaining accurate balance points at low pressures
there is also increasing scatter of results. It is possible therefore only due to increased scatter to
obtain two or three points at low pressures with relatively larger residuals, which if they
happen to be on the same side of the fitted line could be interpreted as indicating a systematic
curvature of the line due to a non zero me value. It is important therefore to confirm, even at
the larger value of scatter, that the curvature is indeed a true feature of the calibration, and not
a random result. Formally, this would be tested by the null hypothesis me= 0 on the regression
parameter, which is explained later, but it is nevertheless important that visual inspection of
the data be used to confirm the results of mathematical analysis.

145
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

Because of the reasons explained above, the hyperbolic regression fit on uncorrected area
might be thought preferable to the parabolic regression fit on uncorrected load. It is important
to understand at this point that the two equations represent exactly the same calibration model.
If the data set were perfect, that is contained no statistical variability, then a linear regression
solution based on either model would yield exactly the same results for fitted parameters.
However, the two forms of the equation do implicitly apply a different weighting to the data
points, and for non-perfect data (that is data which contain statistical variability) regression
solutions to the two equations will yield slightly different results for fitted parameters in order
to minimise the squared residuals condition for that data set.

The parabolic load fit attempts to distribute by best fit the residuals in load equally over the
entire pressure range. That is to say, for example, of the order of 100-200 mg at all points.
This is consistent with the known parameters of the experiment, that balance resolution is
achieved over the whole pressure range at around this level or better (as good as 10 mg under
best conditions).

The hyperbolic fit, being the area parameter, instead attempts to distribute by best fit the
residuals in area (say of the order of 1x10'“^ mm^ for a 10 mm^ piston) over the entire pressure
range. This will give much more weight (in relative terms) than is justified to the very low
pressure points. The ratio of loads between the reference standard and the test unit, given a
relatively fixed load resolution, will be far better at high load (pressure) values than low. In
other words the hyperbolic fit to uncorrected area will be unfairly influenced by the low
pressure points, which are in fact the least experimentally well-characterised points.

In his 1983 Technical Memorandum [3], J Wilbur-Ham states exactly this in more
mathematical terms by stating that the hyperbolic equation can in fact be seen to be the
parabolic equation with weights (variance = 1/P“, standard deviation = 1/P) applied. Clearly
high pressure values correspond to low weight, and vice versa.

It is concluded therefore, that for experimental analysis, inspection and diagnosis of cross-float
experiments, the uncorrected area parameter W7(C',P) is far more useful than the uncorrected
load parameter W*. However the determination of the fit parameters should be determined on
the uncorrected load parameter W' parabola as it is a better fit over the whole range and
apportions weighting equally over the pressure range rather than more strongly to the worst
points at the bottom of the pressure range.

Note that the approach described in Dadson [4] by which the ratio W'/W is determined over
the pressure range to which a straight line is fitted representing (A'o/Ao)(l+(X,2 - X,|)P), in effect
is equivalent to the hyperbolic fit method described above.
Reporting the calibration results of a pressure balance

The pressure balance is used in a broad range of metrological environments by users who have
very different metrological and operational needs. For this reason, the way in which results are
reported and used is relevant, as is the associated uncertainty, which must be appropriate to the
level of the user.

146
WALTER GlARDINl

Calibration for area and load parameters

The most accurate method to characterise a pressure balance is to determine and report the
value of Ao and A, for its piston unit, and the value of me for the calibration event, with their
associated uncertainties. If the mass set is included to make a complete and self contained
instrument, the mass values (on the in-air basis usually) and their associated uncertainties are
also reported, together with the mass of the piston. Whilst the user could in principle perform
the calculations, it is also useful to report the values for the surface tension force, the non­
simple fluid buoyancy term and any other terms required in the report. For the user, however,
this requires the calculation of the pressure output from first principles using the model
equation, and most importantly, the associated uncertainty in the pressure using the same
model. If the user requires merely to know the pressure being measured, then the total mass
applied is a direct input, but if the user requires to generate a particular pressure, then there are
usually trimming weights required to achieve that pressure. With spreadsheet programs this is
fortunately not too difficult.

It is, generally speaking, more convenient for the user to associate each mass with a known
value of the pressure increment. This however comes at a cost as it introduces some
restrictions and some degradation of the uncertainty. The first and most important technical
issue is that the pressure generated by a weight depends on the area of the piston to which that
weight is applied. If the piston's area changes with the operating pressure, which is certainly
significant for most common designs of pistons, then clearly the pressure increment generated
by a particular weight will be different, depending on the current operating pressure. A
common device then, is to fix the order of the weights, and fix the order in which they are
applied to the pressure balance. This would mean for example that weight D is always applied
after weights A, B and C, at the same operating pressure, and will generate the same, well
defined pressure increment. This however does rely on the user ensuring that this loading
protocol is adhered to.

The usual formal approach to addressing this issue, is to report a simpler modified form of the
pressure equation in which only the pressure appears, together with the distortion coefficient as
follows:
P = 'LP.(\-ALP-) (5.14)
where the P, represent the values associated with each individual weight i.

Another approach, which introduces a much larger error, but considerably simplifies the
calculation, is to use a value for the effective area that is the average value that it actually has
in practice over the full pressure range. This will only be correct at the mid point of the
pressure range, and will introduce a linearly proportional error term at both higher and lower
pressures. This may be suitable where the accuracy need is low, or where the range of the
instrument is quite low (e.g., up to around 10 MPa) so that the influence of the distortion
coefficient is relatively low. For example, for a typical distortion coefficient of 3 parts in 10^
per MPa, the area changes by 48 parts in 10^ for a total pressure range of 12 MPa. If the mid
(mean) point is chosen as the value for the effective area over the whole range, then the
effective area will be in error by up to ±24 parts in 10^ at the extremes of the pressure range.
For some users this may be acceptable.

147
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

Note, however, that these approximations ought not to be treated as "uncertainty" parameters.
Generally speaking, the calculation of uncertainty according to the ISO "Guide to the
expression of uncertainty in measurement" (GUM) [1] requires that a proper uncertainty be
evaluated linked to the value of the measurand. Such approximations as those discussed above
should more properly be treated as working approximations. It is recommended that generally
speaking, where the area has a significant distortion coefficient, the linear relationship of area
to pressure be reported in full.

Beyond this, the actual pressure applied by a weight depends on the operating temperature and
the local value of gravity, and in particular the associated uncertainty depends on the
uncertainty in those components. In this case therefore, if the calibration laboratory reports the
uncertainty in the pressure generated by a particular weight, some assumptions must be made
about the uncertainty in the influence quantities at the time of use. If these assumptions are
made, then the user must ensure that to maintain the reported uncertainty, the stated operating
limits are adhered to. Alternately, if the limits are exceeded, the uncertainty calculation should
be updated to include the larger limits. It is also possible for the calibration laboratory to report
the uncertainty associated with a particular weight, on the assumption that there is zero
contribution from the influence quantities. In this case, the user must then add the effect of
these components at the time of use.

A third level of approximation is required when a user wishes to simply use the values of the
pressure as marked on the weights, added arithmetically in the case of multiple weights. In this
case, the user is interested in the maximum deviation of the true pressure from the nominal as
marked on the weights. Again, whilst the use of such values may be appropriate in the context
of "tolerance" limits for instruments used in an industrial context, this is formally not what is
considered to be the uncertainty of the measurand according to the ISO GUM. In this case,
from the user’s point of view, the figure of interest would be the uncertainty calculated for the
pressure generated by each combination of weights, arithmetically added to the maximum
deviation (error) found between the true pressure generated and the nominal pressure marked
on the weight. Particularly in this last case, it is very important to understand that this
approximation still does not take into account such operational factors as the local value of
gravity and the actual working temperature of the piston, and if these are neglected,
particularly the former, very large errors of as much as 0.07% or more can be introduced when
operating over a geographical area the size of Australia.

148
WALTER GIARDINI

5.8 Uncertainty of generated pressure and device calibration


The basic measurands for pressure metrology are;
1. The pressure generated by the reference instrument at the level of the test unit datum
2, The corrections of the unit under test.
And their uncertainties should be calculated using the procedures of the ISO ‘Guide to the
expression of uncertainty in measurement’ [1], which is explained in [6].

We refer again to Figure 5.1 and the fundamental pressure generating equation, Eq. (5.2):
+g-Vg(/?^ -pj + r I
X„(1 + AP„>C, * '■ (5.15)

where Ph = (pf- Pa)gh


Ch = 1- pjprn and
Cj =/+<%( t — 20)
and we have now made the substitution for both the effective mass load term, the buoyancy
and the temperature correction factors. The symbols are described in section 5.2.1.

As stated, some calibration reports do not present the values of £; and -Vg(pf- Pa} explicitly,
as they are small corrections to the main load terms. Instead, their values are included in the
"effective mass" value rrie. In these cases, the uncertainties associated with these terms are
ignored, as they are expected to be included in the uncertainty of me.

The model to this point, the four equations above, does not include drift corrections for some
parameters. In this case, the two critical parameters are the area of the piston cylinder and the
value of the applied masses. However, their drift is difficult to determine for pressure balances.
Clear trends may take many years to become evident, if at all, and factors that affect these
parameters are variable and may be one off, so they may not generate deterministic and time­
uniform drift components. Unless a clear drift is therefore determined, the best estimate is
generally zero. These drift values, however, are themselves associated with an uncertainty, and
so, whilst they do not affect the value of the parameters themselves (area and mass), they do
contribute to the uncertainty in the values of those parameters.
Sensitivity coefficient expressions
Required is the sensitivity coefficient c, for each input quantity, with respect to the measurand,
the output pressure P, where c, - cPIcx, and x, is the i''' input quantity in equation (5.15). These
are tabulated below, and are used in calculating the combined standard uncertainty Uc for P
using (assuming no correlations):
t/" (P) = ■ ir (Xj) or

ir(m) +

For an example of such a calculation, see section 4.7.1 (gas-operated piston gauge).

149
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

Component Sensitivity coefficient, c,


Mass m (and drift component) 3^ g
dM Aq
Taring mass ^P g

Fluid density Pf

Air density Pa dP ~ - P hg Vg
^Pa Pm Pm Pm^O
Density of weights
^Pm Pm J
Gravity g dP P
^8 g
Surface tension force, £
de Aq

Height of head h 3P

Effective Area (0,20°C) Ao (and drift dp -p


component) 3Ao Ag

Distortion coefficient A
dli
Piston cylinder assembly Area thermal dP
-^^-P(t-2())
expansion coefficient (Xa da^
Temperature t
— = -Pa
dt
Excess piston volume V
av

Pressure reading of pressure gauge, 1


correction and error.

150
WALTER GIARDINI

Uncertainty of calibration for a pressure gauge

Calibration of the pressure gauge requires the comparison of the known (reference or standard
pressure) with the indication (reading) of the gauge. The difference between the two is
expressed as the error (deviation of the device from true value), or the correction, (value to be
applied to reading to obtain the true value), these being the measurands of the calibration. For
convenience we work with the error in the following discussion, the correction being the
negative of the same value. The simple model of pressure gauge calibration is generally
expressed in the form:

^E^ = P,-P
When the uncertainty is evaluated, however, we need to account for the resolution of the gauge
(or the readability limit of the least significant digit) and the random scatter of results under
actual operating conditions, called the repeatability. For the sake of mathematical consistency,
it would be desirable to include these effects into the formal mathematical expression above,
and we can do this by writing:

where eo refers to the implicit error present in the reading, due to the resolution limit of the
analogue gauge, or else the least significant digit cut-off of the display. This of course cannot
be determined, and does not change the value of the reading, in fact its value is implicitly
assigned to be zero in all cases, since given that it is unknowable, our best estimate is that "on
the average" it is equal to zero. However, its uncertainty will certainly affect the uncertainty of
the determined error, and by including the parameter formally into the equation, the structure
of the uncertainty calculation algorithm is more straightforward.

5.9 Appendix: Buoyancy correction for immersed parts


In this section we show that the fluid buoyancy corrections to the piston just equal and cancel
the correction required to correctly refer the gauge pressure to the ambient air pressure at the
same level outside the hydraulic pressure system.
The total mass of a pressure balance is adjusted for buoyancy by multiplying by the term
(1 — ), but the piston itself is immersed in oil not air. The question is often asked - is
this an error in the calculation? The answer is no, for the following reason.
The pressure balance by definition measures the gauge pressure, that is, the difference between
the absolute pressure inside the hydraulic circuit and the absolute pressure of the ambient air at
the same level outside the hydraulic circuit. When the complete calculation of gauge pressure
is performed, the extra small head-of-air correction to the reference height level introduces an
extra term that is exactly the same size, but opposite in sign, to the small bias introduced by
assuming that the entire load component (masses and piston) are all immersed in ambient air.

151
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

To see this, first consider the piston-cylinder and all other masses as being composed of the
true piston, itself extended to the top of the weight stack to make a hypothetical extended
piston in the centre, plus all the extra masses, as in the following figure:

Reducing this to the essential elements, which is the piston (partially in fluid and partially in
air), plus the added masses (fully immersed in air) we obtain Figure 5.4.

Figure 5.4: A schematic illustrating a pressure balance and its essential elements.

Ml is the total mass of all components except the solid piston that is extended from the actual
piston right up to the top surface of the weight stack. M2 is the part of the extended piston that
is immersed in air, and M3 is the part of the extended piston that is immersed in oil. In this
simple case, we consider that all components are made of the same material and hence have
the same density, and that the piston is of simple cylindrical form.

At equilibrium the total sum of all forces acting on the complete piston + weights = 0 and so
we have:

152
WALTER GIARDINI

M,(\-p„lp„} + (M,g + AP^) + (M^g-AP} = Q Al


where
P‘ term = atmosphere acting on the upper and lower exposed surfaces of the weight stack
2"^ term = weight + pressure on top surface acting over area A and
3*^^ term = weight + pressure on lower surface acting over area A.

The gauge pressure at the level of the reference datum (the lower face of the piston) is P - P'a
(i.e., the atmospheric pressure at the same level), and P'a = Pa + Pagh, where the extended
height of the piston is h.

Re-arranging equation Al, subtracting P’a and substituting as shown above, we obtain:

Simplifying the above and using the following identities,


volume of extended piston = hA = mass/density = (M2+M3)/p,n we obtain:

P-P'A=f^xg('^-PalPj + f^2g-^l^ig-P.8(’^i+M^)l P„
= M,g(i-pJp„) + (M^+M,)g-p^g{M,+M,)l p„
= M,g(\-pj pJ + kM.+M.Mi-pj p„)
= (M^+M^+M,')g(A-Palp„}
= Mg(i-Palp„}

which shows that the gauge pressure at the level of the test datum is just exactly the value of
the total mass adjusted for buoyancy as if the whole thing was immersed in ambient air.

This calculation can also be performed in exactly the same way, (that is by calculation of each
separate component and then summation) in the more general case where the piston is made of
one material (e.g., tungsten carbide), the weight stack a different material (e.g., steel), and the
piston has an extended retaining nut at the bottom made of a different material again (e.g.,
aluminium). The result is the same.

153
CHAPTER 5. THE HYDRAULIC PISTON GAUGE

5.10 References
[ 1 ] ISO Guide to the Expression of Uncertainty in Measurement. International Organization
for Standardization, Geneva, 1995. ISBN 92-67-10188-9.

[2] Sandars, G. (2002). Traceable Measurements (Monograph 3: NML Technology Transfer


Series, CSIRO)

[3] Bignell N. (ed.). Course on Pressure Measurement, CSIRO Division of Applied


Physics, 1986.

[4] Wilbur-Ham J.L., Technical memorandum M04, CSIRO Division of Applied Physics,
Comparison of Two Programs for Pressure Testers, November 1983.
[5] Dadson R.S., Lewis S.L., Peggs G.N., The Pressure Balance - Theory and Practice,
NPL, Department of Industry, London, HMO, 1982

[6] R.E. Bentley. Uncertainty in Measurement'. The ISO guide. CSIRO, Sydney, Monogr. 1:
NML Technology Transfer Series, 2003.

154
Chapter 6

Pressure transducers
Noel Bignell

6.1 Introduction
A pressure transducer takes a pressure and converts it into some form of output quantity that is
usually electrical, a voltage or current, or another pressure. We measure the quantity that
comes out as a way of measuring the pressure input. Why bother to do this and not to measure
the pressure directly? Because the output may be more easily or conveniently measured or put
in a form for computer analysis or because it may be transferred to a remote place more
suitable for measurement processing.

The range of accuracy available in pressure transducers is exceeded only by their range of
prices. Very inexpensive pressure transducers are being used in motor vehicles increasingly, to
monitor and then control performance and some expensive ones perform measurement tasks
not possible any other way. This chapter will look at the types of pressure transducers
available but first there are some general principles that apply to all pressure transducers and
indeed to all transducers whatever the input quantity or measurand. These will be discussed for
an ideal transducer and then the various forms of non-ideal behaviour will be considered. This
leads to a discussion of the specification of a pressure transducer and how it may be calibrated.
Later various specific types of transducer are discussed and their characteristics. A typical, but
fictitious specification, is examined in some detail. Finally, an example of the uncertainty
calculation needed to assign an uncertainty to a pressure measurement performed by a pressure
transducer is presented.

6.2 The transducer as a black box


Sometimes it is possible to be confused by detail and to miss essential elements - not to see
woods for trees. The black box approach considers only those things that are necessary; the
internal workings of some complex arrangement are disregarded in order to concentrate on its
inputs and outputs. Thus we can characterize the transducer as the box in Figure 6.1 and not
worry, for the moment, about how it works inside the box. The pressure (or in general any
quantity) is the measurand, the quantity to be measured. The output may be a voltage, a
current, a frequency or another pressure. In order to function, and at this stage we will not ask

155
CHAPTER 6. PRESSURE TRANSDUCERS

why, the transducer will generally require some energy source, usually an electrical one,
known as excitation.

MEASURAND OUTPUT
TRANSDUCER
(pressure) (voltage)

EXCITATION
(power supply)

Figure 6.1: The black-box model of a transducer focuses attention on the inputs and
outputs and the relationships between them.

6.2.1 The ideal relationship between measurand and output

For a transducer to be of any use there must be a consistent relationship between measurand
and output and this is best shown on a graph where the measurand is given the X, or horizontal
axis, and the output the F, or vertical axis, as shown in Figure 6.2. Here we have made the
measurand and the output both go from 0 to 100, graduations that may be thought of as
percentages of their actual values. This relationship is called a calibration curve and is shown
in Figure 6.2, curve (a), as a straight line so that

Y = mX (6.1)

is the equation describing it, where m is a constant. The sensitivity is the slope of the straight
line, m, and is the change in output divided by the change in measurand. This would apply to a
pressure transducer using a simple diaphragm, but for other transducers the ideal calibration
curve might be
Y = aX^+bX (6.2)

where a and b are constants, in which case the curve might be represented in Figure 6.2 as
curve (b). In this case the sensitivity is not constant but depends on the value of the measurand
(pressure). Many other forms are also possible.
6.2.2 Dynamic response and frequency response
Not all the information is shown in Figure 6.2 because there is no time axis. If a pressure is
applied to a transducer then the output will not be able to respond immediately, even in an
ideal case, but will respond as shown in Figure 6.3. Here, for the time before zero, the output
was constant at F/ because the measurand had been constant at Xj for a long time. At t = 0 the
measurand immediately changed from Xj to X?- The output then started to change to F? but the
various ways it might do this are indicated in curves (a), (b) and (c), which give the transient or
dynamic response of the transducer to a step change in measurand.

156
NOEL BIGNELL

Figure 6.2: The relationship between the measurand and the output is shown
graphically, for two transducers, as curves (a) and (b). The values for the measurand
and for the output have been scaled to go from 0 to 100 in both cases.

The behaviour shown in curve (a) is known as underdamped. The output oscillates about its
final value before coming to rest. The ringing period is the time for which the output
oscillations exceed 10% of the final change in output value.

In curve (c) the transducer is overdamped and comes to rest without overshooting the final
value at all. In curve (b) the damping is between overdamped and underdamped without being
either and is said to be critically damped or, sometimes, aperiodic. To be critically damped is
generally considered a good thing for an instrument though under some circumstances
underdamping can be used to produce a more rapid response.

Figure 6.3: The response of a transducer to a step change in measurand can be


underdamped (a), overdamped (c) or critically damped (b).

The ideas of response time and rise time are attempts to quantify the transient response. The
rise time is the time taken for the output to go from (usually) 10% to 90% of its final value.

157
CHAPTER 6. PRESSURE TRANSDUCERS

The response time is the time to rise to within a specified percentage of the final value, and to
remain within it (to allow for underdamped instruments).

Figure 6.4: A general form of frequency response curve showing a flat section at low
frequency and a fall-off in response at high frequency.

The frequency response of a transducer may be represented by a graph, such as that in


Figure 6.4. It is a graph of the amplitude of the output of the transducer versus the frequency
of the measurand, for a measurand of sinusoidal form and of constant amplitude as shown in
Figure 6.5. The frequency response curve can take many forms but the one shown is the
simplest general form, showing a flat section at the low frequency end and a falling response at
high frequencies. The static or DC response corresponds to zero frequency.

It is assumed in Figure 6.5 that the input signal has been applied for a time long enough for all
transient responses to have died down. Though, as shown, both output and input are of
sinusoidal form they are of different phase. This can be seen by looking at the value of the
input at zero time when it has a value mid-way between its maximum and minimum whereas
the output at the same instant has its maximum value. This is an example of a phase shift, and
such shifts are associated with those parts of the frequency response curve that are not
constant. The phase shift will vary with frequency and this variation is, properly, part of the

Figure 6.5: A sinusoidal measurand and a sinusoidal output that is shifted in phase.

158
NOEL BIGNELL

The frequency response curve may also, in some part of the frequency range, rise to a
maximum and then fall away. This is called a resonance and will usually be associated with the
mechanical vibration of a part of the transducer such as the diaphragm. Reliable operation will
not be obtained in the vicinity of a resonance. The frequency response may be calculated from
the transient response or vice versa.

6.3 Real transducers


For an ideal transducer, a complete specification would simply be the calibration equation and
the rise time (or the curve in Figure 6.3), but in the real world much more is needed to specify
the magnitudes of the various deviations from ideal behaviour that apply to real transducers.
6.3.1 Zero error
If the output from the transducer is not zero for zero measurand this can be a zero error.
Sometimes the actual output is not zero by design but in this case in the processing of the
output an allowance will be made for this. An ordinary zero error is illustrated in Figure 6.6
and the equation corresponding to it is
Y = mX + C . (6.3)
There is usually an adjustment for making the zero error equal to zero, an adjustment that can
be made without a pressure standard by making the measurand zero. Zero error is a special
case of datum error, that is, the error when a particular value of measurand is chosen as a
datum for checking the instrument. For example a transducer meant to read gauge pressure -
100 kPa to +200 kPa should read zero when atmospheric pressure is applied. A transducer
meant to produce 4 to 20 mA of current when gauge pressures of 0 to 1 MPa are applied
should give 4 mA when atmospheric pressure is applied.

Figure 6.6: A zero error for a transducer having a linear relationship between
measurand and output can be represented by the equation Y = mX + C.

6.3.2 Linearity
Even in cases where the relationship between measurand and output is supposed to be a
straight line, i.e. linear, it often happens in practice that it is not, i.e., it is non-linear.
Obviously, there are many ways in which a calibration curve may not be straight and if we are

159
CHAPTER 6. PRESSURE TRANSDUCERS

not sure where the “proper” straight line should go then it is not possible to unambiguously
specify non-linearity errors with a simple number. There are several possible approaches to
specifying deviations from linear behaviour depending on this choice of the “proper” straight
line.

Least-squares straight line

The data, shown as “X” in Figure 6.7(a), are fitted to a straight line by the criterion of least
squares, i.e. the line is placed so that the sum of the squares of the deviations from it is as small
as possible. This calculation is simple with a small computer and, as the topic is always
covered in texts on statistics, it will not be further discussed here. However, precautions should
be taken when considering the uncertainty of fit (see section 4.2 in [16]).

To specify the linearity, the largest deviation (-i-ve or -ve) is given. This is called an
independent linearity, or a least-squares linearity specification.

Terminal Straight Line (TSL)

The instrument in Figure 6.7(a) has a zero error, and, in general, this could be adjusted out so
that the transducer read zero with zero measurand. If this is done then a straight line can be
drawn from zero to the point having the maximum measurand value allowed. This gives rise
to a terminal based linearity specification indicated in Figure 6.7(b). It will be written as for
example, 0.2% FS TSL.

Figure 6.7(a): A straight line may be fitted to the calibration points, shown as “x”, by the
method of lease squares. The worst deviation error is the largest difference
between this straight line and the data. A zero error is evident in this
example.

Least-squares-through-zero straight line

If we adjust the zero, but still fit a straight line by the least squares criterion, then we have a
situation, shown in Figure 6.7(c), called zero-based linearity.

160
NOEL BIGNELL

Figure 6.7(b): The terminal straight line is one joining the extreme values of the
calibration data that have been corrected for any zero error.

Best Straight Line (BSL)

This is a confusing term as the least squares straight line is often referred to as the best straight
line. Here the best straight line is that line mid-way between two parallel straight lines that
just enclose all the points, as shown in Figure 6.7(d). This is also sometimes called
independent linearity. It will be written as for example, 0.1% FS BSL.

Theoretical straight line

If there is some reason to believe that a particular straight line should apply, or a straight line
of a particular slope, then deviations from it may be used to specify the linearity.

Figure 6.7(c): The least squares through zero straight line is fitted after the calibration
data have been corrected for zero error.

6.3.2 Hysteresis
It is almost always happens that pressure transducers, like pressure gauges, do not read the
same for increasing pressure as for decreasing. This is shown in Figure 6.8 where it can be
seen that, although the transducer initially had no zero error, it did not return to zero after its

161
CHAPTER 6. PRESSURE TRANSDUCERS

cycle and that, for one value of measurand, there are two values of output. This is called
hysteresis.

Figure 6.7(d): The best straight line is mid-way between two parallel lines that just
enclose all the data.

In pressure gauges, hysteresis is caused principally by friction and is often minimized by


tapping the glass. For transducers the friction is of a more subtle kind and is often caused by
elastic materials not returning exactly to their former shape after deformation. A sophisticated
form of “tapping the glass” called dithering may sometimes be used to minimize hysteresis.

To specify it, it is necessary to state the cycle performed and the time taken since there will
often be a slow recovery. Maximum hysteresis (band) is the maximum difference between the
two output values for the same pressure in a cycle, and it may be expressed as a percentage of
full-scale output.

Figure 6.8: Hysteresis occurs when the output for increasing pressure is not the same as
that for decreasing pressure.

6.4 Change of characteristics - instability


The departures from non-ideal behaviour discussed in Section 6.3 are bad enough, but,
unfortunately, the output of pressure transducers is affected by quantities other than pressure.

162
NOEL BIGNELL

Time, temperature and level of excitation usually have significant effects and their influence
needs to be specified.
6.4.1 Aging
Transducers usually change their characteristics and their calibration as they age. For this
reason a calibration is only valid for a certain time. Good stability usually involves using
stress-free components, achieved by annealing, but the conditions of use, especially over
pressuring, can seriously alter the characteristics of pressure transducers. The usual
specification would be that the reading at room conditions remains within a certain percentage
of full scale of the calibration value for many months.
6.4.2 Temperature stability
Almost all measuring devices are seriously affected by temperature and pressure transducers
are no exception. The best accuracy is achieved by using them at the temperature of calibration
and keeping this constant, if possible, by using a good temperature controller. If this is not
possible, as is usually the case, then there are many techniques available to minimize the effect
of temperature and some of these are almost always used in pressure transducer design. Often
they are very effective and accurate temperature compensation is achieved over a range usually
specified, eg. 0 to 50 °C.

However, this applies only when the whole transducer is at the same temperature and often this
is not the case, i.e., temperature gradients exist. If the temperature compensating part of a
transducer is at one temperature and the measuring part at another then, clearly, the
compensation will be ineffective. They can be made more nearly coincident if the transducer is
very small. In general, it is best to avoid taking measurements after sudden temperature
changes.

If there are stresses in the pressure sensing parts of a transducer then higher temperatures will
release these more rapidly. This is annealing and the characteristics may be permanently
changed in what is a sort of accelerated aging.

Pressure transducers have been developed to operate from just a few degrees above absolute
zero to exhaust gas temperatures in engines. The temperature specification is an important one
for all pressure transducers and should include both the range of operation and the errors to be
expected when not operating at the calibration temperature. There will probably be a change
of zero, and a change in sensitivity and linearity are likely. The zero should, if possible, be
adjusted at the working temperature, or in the middle of the working range of temperatures.

6.4.3 Excitation

Changes in the (usually) electrical energy fed to the transducer to make it work can change its
characteristics. Sometimes there is a direct relationship between the sensitivity and the voltage
applied, e.g., for strain gauge types. This is normal and allowed for but there may be other
effects, due to heating for example, which will change the temperature compensation
characteristics. There should be a statement in the specification for stability with respect to
excitation.

163
CHAPTER 6. PRESSURE TRANSDUCERS

6.4.4 Other factors affecting stability

Differential pressure transducers usually undergo a zero shift when the reference pressure, or
line pressure, changes. This is the reference-pressure zero shift. There is, similarly, a
reference-pressure sensitivity shift. Both of these should be specified but it is unlikely that the
latter will be, even though errors of up to 1% may occur for line pressures of 3.5 MPa.

Humidity may change the characteristics but should not be a problem in a well-designed
transducer, many of which can operate under water.

Special conditions of operation often mean that it is important to know how magnetic fields,
ionising radiation, vibration or corrosive chemical environments affect the transducer and the
manufacturer of these special devices will provide the specification. It is often difficult to
check the validity of these claims. Nearly always a specification will nominate the “wetted
materials”. These are the substances that come in contact with the pressure fluid, and for
pressure measurements in a corrosive fluid this specification may be the most important.

6.4.5 Repeatability or instability

Even when all the above factors affecting stability do not seem to apply, the same measurand
may not always produce the same output. The repeatability is the maximum difference
between output readings when the measurand, the pressure, is applied in the same direction
and under the same conditions over a specified period of time. It is expressed as a percentage
of full scale (FS) for a specified period. An example from an actual manufacturer is: ±0.06%
of reading ±1 digit between 1 % and 100% FS. This figure includes 90 day stability.

6.5 Calibration and pressure standards


All pressure transducers must be calibrated. There is nothing particularly different involved in
the calibration of pressure transducers from the calibration of other pressure measuring
instruments. A standard is needed, which might be another pressure transducer, and various
standards are mentioned in this section but generally they are more fully described in other
chapters. Some fairly general procedures for various types of calibration are given in this
section.
6.5.1 Pressure transducers
These can be very convenient because of the ease of reading and the possibility of automating
the calibration process. Obviously they must have superior characteristics to the transducer
being calibrated. They will also have to be calibrated themselves so that one of the other
standards will have to be used, perhaps in another laboratory. Pressure calibrators is the name
given to instruments that contain a pressure transducer, usually with a digital output display,
and a small pump that enables the calibration pressures to be generated.

164
NOEL BIGNELL

Gas
Those transducers using gas will generally operate at pressures less than 10 MPa, but some
may go higher. It is advisable to calibrate these using oil as the pressure medium because of
the danger inherent in the use of high pressure gas.

Vacuum standards are dealt with in Chapter 2, but the term “pressure transducer’’ does not
usually include the high vacuum gauges, though, technically, they are transducers. Mercury
manometers are useful for both absolute and gauge pressures from 100 Pa up to 100 kPa. See
Chapter 3. Liquid columns using water or special oils are useful standards, generally for gauge
pressures of a few Pa up to 20 kPa. They are dealt with in Chapter 4. Piston gauges may be
absolute or gauge and operate over a wide range of pressures from about 10 kPa to 7 MPa or
more with accuracies up to 0.01%. See Chapters 3 and 4.
Oil
Piston gauges, or dead-weight testers, are the standards used for calibration over almost all the
range of pressure (always gauge) for which oil is used as a medium, up to more than 500 MPa.
They are dealt with in Chapter 5.

6.5.2 Ancillary equipment

Apart from the pressure standards discussed above some extra equipment will be needed.
Power supplied, obviously, must be of a quality so as not to degrade the performance.
Generally, a good voltage-regulated power supply with the capacity to vary the output voltage
(to test how this affects transducer output) will be suitable. Sometimes a constant-current
supply will be desirable for semiconductor-strain-gauge transducers as they have superior
temperature stability when operated in this mode.
The output must be measured with an uncertainty preferably ten times, but at least three times,
better than the repeatability of the transducer. This can be done easily with digital voltmeters
(DVMs) or multimeters (DMMs). These can be used to measure voltage and (by passing it
through a standard resistor, if necessary) current, but also can be useful in resistance ratio
measurement, as shown in Figure 6.9, when the DVM will give a direct reading of the ratio
/? //(Rj+Ri)-

If tests at temperature other than room temperature are required then a suitable temperature test
chamber and equipment for measuring the temperature of the transducer case will be needed.
This set-up could be quite elaborate for large temperature ranges or could be quite modestly
achieved with an insulated box and blower heater together with a thermocouple thermometer
fixed to the transducer case.

Dynamic temperature tests can be performed in a semi-quantitative way using cans of liquid
Freon or preferably the environmentally friendly alternative. A bath of Woods metal, a low
melting point alloy, maintained at a temperature slightly below the upper operating
temperature can often be useful, as described later.

If vibration is a significant parameter then a shaker table will be needed. A calibrated hammer
may be sufficient for simple tests of vibration sensitivity. For monitoring the output drift a
chart recorder could be used or the output of the DVM could be sent to a computer and a table

165
CHAPTER 6. PRESSURE TRANSDUCERS

of values printed out. Once the output values have been stored in a computer the drift rate can
be calculated, averages performed or statistical tests conducted.

EXT
STANDARD

POTENTIOMETER
OR HALF BRIDGE
Figure 6.9: A digital voltmeter may be used to measure the ratio of two resistors as used
in a strain gauge pressure transducer.

6.5.3 Static pressure calibration

There is no Australian Standard that deals with static pressure calibration but there are some
overseas publications listed in Further Reading. What follows is not a step by step procedure,
but rather attempts to outline the essentials needed for pressure transducer calibration.
Among the important characteristics of a transducer are: pressure range, full scale (FS) output,
sensitivity, zero output, linearity, hysteresis, resolution or the smallest increment of pressure
that produces an observable change in output, and repeatability. Sometimes the dynamic
characteristics are important and these will be considered in a separate section.

Essentially pressure transducers can be calibrated as for Bourdon tube test gauges as indicated
in AS 1349-1986 with commonsense differences. One difference is that the pressure transducer
and extra equipment must be allowed adequate time to warm up. The pressure connection or
mounting torque may be important, and may be specified by the manufacturer, but in any case
the setting should be repeated for consistent results.

A calibration cycle consists of a series of measurements with pressure ascending and then
pressure descending. The first point and the final point are normally at ambient pressure. The
zero may be adjusted only at the beginning of the tests. Three cycles would be the maximum
performed.

Recording the excitation voltage or current, time and ambient temperature at the beginning,
middle and end of the cycle will yield (say) 11 points (for 20% increments) or 21 points (for
10% increments).

166
NOEL BIGNELL

Care must be taken to keep the pressure changes at the transducer in one direction, except at
the maximum, to test for hysteresis. This can be achieved by using an isolating valve
(preferably a constant volume type) between the transducer and the pressure standard and
making the pressure correct for the next point before opening the valve. A volume adjuster
may be needed in the system to make up for the volume change of the valve and compression
of the fluid in the transducer. This would be the screw press of a dead-weight tester.

Resolution is measured by slowly changing the input pressure and observing jumps in output.
This is usually done for potentiometric types of transducers and if the number of turns of wire,
n, on the element is known the resolution can be calculated from FS/n. These types benefit
(like Bourdon gauges) from tapping, and a small buzzer attached to the case can be used to
provide a continuous vibration (dither).

Normally the tests would be performed at the temperature of operation and perhaps at
temperatures lower and higher. The transducer should of course be allowed time to stabilize at
any new temperature. This involves an enormous amount of work if a wide temperature range
is to be checked. The transducer may be checked for only zero output at several temperatures
in its range of operation to shorten this procedure, but of course this gives no information
about changes in linearity, sensitivity or hysteresis due to temperature.

Dynamic temperature tests can be done by observing the zero output while the transducer has
cold and/or hot air blown over particular parts of it or while it is partly plunged into a bath of
molten Wood’s metal. This latter is mainly useful for flush diaphragm transducers. The zero
output will probably shift quite a lot, first in one direction and then in the other, before slowly
coming back to a value close to the original value. The times involved very much depend on
the physical size of the transducer, the smaller the quicker, but several minutes would be
typical value for regaining zero.

Vibration tests are performed with the transducer mounted on an electromagnetic shaker table
and should be done in three orientations at right angles. The output is monitored and
resonances at specific frequencies identified by the rise in output. In potentiometric transducers
the resonance of the slider may cause a complete loss of output.

In using a calibrated hammer, the transducer is tapped with the hammer and the magnitude of
the tap can be measured as can the output change it causes, by recording hammer output and
transducer output on a storage oscilloscope. Several orientations should again be used. A
simple alternative to the hammer is to drop ball bearings of known weight from a known
height on to the transducer. The impact delivered will depend on the surface struck but, though
indefinite in value, the impact will be repeatable.

The warm-up time can be measured by measuring the cycle - zero output, FS output, zero
output - immediately after switch on and repeating at intervals of 5 minutes for the first half
hour, 10 minutes for the second and so on until stability is reached.

Ageing tests, to estimate stability with time, can be best done by building up a history of
calibration over a period of time. The ageing can be accelerated by holding the transducer at
higher temperatures and by subjecting it to pressure cycling using solenoid valves which

167
CHAPTER 6. PRESSURE TRANSDUCERS

automatically switch the input pressure from minimum to maximum for many thousands of
cycles. The transducers would be calibrated from time to time during this ageing process.

Sometimes, it is necessary to know how well an unused transducer, kept in the normal
conditions of storage, will maintain its calibration. Calibration every 6 months or so is about
the only way of determining this.

6.5.4 Dynamic pressure calibration

There are three main approaches to the calibration of the dynamic response of pressure
transducers. A static calibration is sometimes (but not always) possible, and often used as the
only calibration. Where possible a static calibration should be performed because it will be
undoubtedly more accurate, at zero frequency, than the dynamic calibration extrapolated to
zero. For this reason it can be used to check, and even to adjust, the results of the dynamic
method.

There are three main types of dynamic calibration techniques discussed below. In all cases the
output from the transducer is recorded using storage oscilloscopes or some form of fast data
logging facility.

Pseudo-static calibration

In this technique a static pressure is applied to the transducer as rapidly as possible using fast
acting valves. The static pressure may be measured using any of the standards discussed in
section 6.5.1.

One apparatus used for calibration with gas consists of a manually operated valve and a 300 L
tank that may be pumped up to 1 MPa. When operated, the valve simultaneously closes a leak
from the transducer chamber to the atmosphere and opens this chamber to the tank pressure.

Another device is shown in Figure 6.10 and consists of a similar set-up as the one above, the
novelty being in the way the valve is opened. The chamber housing the transducer is protected
from the pressure by a sealed plate. This is rapidly moved away by a heavy weight dropping
on to a handle attached to the plate. The data-recording device is triggered at the same time.
Rise times of 100 ps are claimed and it is designed to be used up to 1 MPa.

The same principle is also used in reverse; pressure applied to the transducer is suddenly
released. There are some advantages in this variation in that there is no correction for the drop
in pressure due to the expansion into a large volume and also that the rise times are said to be
faster. This pressure-dumping technique has been used in oil using a dead-weight tester to
establish a steady pressure before release. Since most dynamic pressure transducers are used
to measure positive pressure pulses there is a slight assumption that the characteristics
measured in reverse, for negative pressure jumps, are the same.

Shock-tube calibration

168
NOEL BIGNELL

TO ATMOSPHERE

OUTPUT

TRANSDUCER

HIGH PRESSURE
INPUT

HEAVY WEIGHT

Figure 6.10: A rapidly operating valve activated by the fall of a heavy weight used in the
dynamic calibration of pressure transducers.

A shock tube is a device, or a technique, for producing very fast pressure pulses. It consists of
a long tube of, say, 150 mm diameter and about 6 m long divided into two compartments
separated by a diaphragm. The smaller compartment contains air (or other gas) at high pressure
Ph (about 1 MPa). When the diaphragm is ruptured, usually by a mechanical device that
pierces it, a pressure discontinuity, or shock wave, travels down the tube containing air at
pressure Fp, and a rarefaction shock travels up the tube containing the high pressure. In air the
pressure pulse height produced is:

5 V 7
= )’ (6.4)
o 6c
where c is the velocity of sound in air at the pressure Po and V is the velocity of the shock
wave. This can be easily measured with transducers placed in the wall of the shock tube so
that the magnitude of the pulse is easily calculable.

From knowledge of the pulse response of a pressure transducer it is possible to calculate all the
dynamic characteristics.

Sinusoidal pressure calibration

A sinusoidal pressure variation can be generated by placing a liquid column on a shaker table.
The vibration subjects the column to a sinusoidal time-varying acceleration and, hence,
through the relation:

P = pg'h (6.5)

a varying pressure is generated. Here, p and h are the density and height of the column as
usual, but, now g' is not just the acceleration due to the gravity, but also that due to the shaker
table and gravity in combination. When the two accelerations are equal but of opposite sign the

169
CHAPTER 6. PRESSURE TRANSDUCERS

pressure generated will be zero. This limits the pressure variation at the bottom of the column
to from zero to 2pgh. By pressurizing the top of the column with gas at high static pressure,
larger accelerations can be used and hence larger pressure variations. Accelerations up to 29 g
have been used.

The frequency range depends on the table, and typically will be up to at least 2 kHz. This
enables resonances in this range to be identified and, in general, the frequency response to be
measured. There will be a resonant frequency due to the column itself that will complicate
measurement, as it will have to be avoided. For a 220 mm column of water this will be at about
1500 Hz but may be damped out by suitable techniques so that it does not interfere with the
instrument. Diaphragm transducers should be oriented on the table so that the vibration is
parallel to the diaphragm, to avoid diaphragm resonances and direct sensitivity to the
acceleration. From the frequency response the dynamic characteristics of the transducer can be
calculated.

6.6 Pressure transducers using elastically deformable elements


By far the largest class of pressure transducers are those that rely on the elastic properties of
some material to produce some counteracting force to that of the pressure. The stability of the
transducer is then directly dependent on those elastic properties and special materials have
been developed whose elastic properties are stable, and in some cases, temperature
independent. The most important elastic materials used are fused and crystalline quartz, single
crystal silicon and certain alloys such as Ni-SPAN-C902, beryllium copper, phosphor bronze
and stainless steels.
Quartz has excellent stability and low hysteresis over a wide temperature range, and is used in
a number of the better quality pressure transducers. Forming and machining quartz is difficult
and so these transducers tend to be expensive. Single crystal silicon is used mainly because it
is the material from which integrated circuits are made, though its elastic properties are also
good. There are a number of pressure transducers using integrated circuit techniques to form
strain gauges and signal processing circuits directly on to the silicon that acts as the diaphragm
for the pressure transducer [5].

Guillaume, in his work on iron-nickel alloys, for which he won the Nobel Prize for Physics in
1920, discovered an alloy of iron, chromium and nickel with zero thermo elastic coefficient,
i.e. the modulus of elasticity does not change with temperature. It was called Elinvar. Ni-
SPAN-C902 is one of a series of new versions of this alloy that include titanium. By varying
the chromium content and the heat treatment, positive, negative or zero thermo elastic
coefficients can be obtained. This and the other alloys mentioned are much used in Bourdon
tubes as well as in diaphragms, capsules, bellows and so on, for pressure transducers.
6.6.1 Diaphragms
In its simplest form a diaphragm is a flat circular plate clamped at the edges. Corrugated types
are also used to increase sensitivity. A capsule consists of two corrugated diaphragms placed
back to back so that there is a space between them. This space is either evacuated or
connected to a reference pressure source. When it is pressurised more on one side than the

170
NOEL BIGNELL

other it distorts elastically and for small movements the deflection is proportional to the
pressure difference. The radial and tangential strains are also proportional to the pressure
difference and so either the deflection or the strain may be used to measure pressure.

The deflection occurs according to the following equation

(6.6)
where S is the diaphragm deflection at the centre, v is Poisson’s ratio, r is the active radius, E
is Young’s Modulus and t is the thickness.

Strain Gauge
Diaphragms with strain gauges for detection of the strain caused by the applied pressure are by
far the most common design in use. The diaphragms range in size down to about 1 mm
diameter, small enough to be used in a catheter to monitor blood pressure in a vessel.

The classic strain gauge is a resistance element of nichrome, other resistance alloys, or
platinum, mounted so that it stretches a little with applied strain. A wire stretched between
two points can be a strain gauge and its resistance will increase if the points move apart. To
increase sensitivity several loops of wire may be used. These unbonded types are not much
used now. Bonded strain gauges are mounted permanently, using an adhesive compound, to
the element to be stressed. They are much less sensitive to vibration and are smaller than
unbonded types.

The most modern strain gauge is the semiconductor that is either a bonded type or is made
directly from the same material as the diaphragm itself by integrated circuit technology.
Semiconductor strain gauges have a much higher output than wire types.

Often the strain gauge is not mounted directly on the diaphragm but on another element, on
which the diaphragm acts, to produce the strain detected.

One single strain gauge can measure strain, but is also sensitive to temperature so that the
electrical arrangement is usually that shown in Figure 6.11 and is a Wheatstone bridge, or half­
bridge, arrangement with compensating and adjusting resistors which decrease the sensitivity
in favour of improvements in thermal stability. The sensitivity is quoted in mV/V, millivolts
per volt of excitation for full-scale output, and typical values are 1 or 3 mV/V for wire types
and 10-30 mV/V for semiconductor types. A typical maximum excitation voltage is 10 V
giving 20 mV for full-scale output for a sensitivity specification of 2 mV/V.

Strain-gauge pressure transducers have fast response, good (so-called infinite) resolution, are
not sensitive to vibration, usually well compensated for temperature effects and are simple in
construction. However, the zero needs frequent adjustment and the output is so small that
further signal processing is needed.

An advantage that pressure transducers using the microcircuit techniques of integrated circuits
have is that signal processing can take place on the chip itself. Thus these devices are more

171
CHAPTER 6. PRESSURE TRANSDUCERS

like measurement systems combining the functions of excitation, conditioning, strain-gauge


bridge, amplifier and compensation in one microcircuit. They are also capable of being quite
inexpensive. The accuracy is rarely better than 0.1 % FS with 0.3% - 1 % being more typical.

SENSITIVITY

OUTPUT

Figure 6.11: A resistance bridge configuration often used in strain-gauge pressure


transducers.

The pressure-induced strain, which changes semiconductor carrier mobility, alters the
properties of MOS field-effect transistors. A pressure transducer using these devices in a ring
oscillator has been developed which has a frequency output. By using a combination of two
ring oscillators with opposite temperature coefficients, a pressure transducer with good
temperature stability has been reported (Neumeister, Schuster and von Munch, 1985).
DIAPHRAGM

PI P2
Figure 6.12: A variable reluctance pressure transducer in which diaphragm movement is
detected by the change in the ratio of the inductances LI and L2.

Inductive and reluctive


The movement of the diaphragm may be measured by means of magnetic induction, either
acting directly on the diaphragm, as in Figure 6.12, or via a magnetic core and a linear variable
differential transformer (LVDT), as in Figure 6.13 [3]. Generally the sensitivity is much
greater than for strain gauge types, e.g., the instrument in Figure 6.13 can detect 1 Pa, but the
zero stability still relies on the mechanical stability of the diaphragm.

172
NOEL BIGNELL

The electronic support needed tends to be greater than for strain gauges and the transducers are
larger, but, because less strain need be generated in the diaphragm, the accuracy is better.
They are tending to be replaced by strain gauge types or by the type following.

Capacitive
This type uses the change in capacitance between the diaphragm and a fixed electrode or
electrodes to detect its movement. The technique is capable of the best sensitivity but the
electronics needed is more than for the others so that these gauges are expensive. They are
often used for gas pressure, sometimes in the vacuum range. They are available in differential
or gauge mode as well as in absolute mode. In this case there is a sealed vacuum space on one
side of the diaphragm, maintained vapour free by chemical absorption or gettering, to which
pressures applied to the other side of the diaphragm are referenced. To improve stability these
units are sometimes temperature controlled at about 40°C or more, which also prevents
vapours from condensing which would seriously disturb performance. Another way to prevent
contamination is to use a sealed cell filled with silicone oil and to transmit the pressure to the
oil via corrugated diaphragms.

Figure 6.13: The movement of the diaphragm in this transducer is detected with a linear
differential transformer. The oblique fingers to the left of the diaphragm
serve to produce a radial tension in it and so prevent an “oil-can” effect.

The best types offer an uncertainty of 0.05% of reading, repeatability to 0.0023% of reading,
sensitivity of 2 x 10*^ of FS and good temperature stability, with operation up to pressures of
500 kPa. Less expensive types are available, which use miniaturized electronics to produce a
small package with 0.15% of reading uncertainty and repeatability of 0.01% FS.

The PRESSFET is a MOS field-effect transistor in which the gate structure is lifted and made
into a pressure sensitive diaphragm [2].

Light sensing
The weightless amplification of a long lever arm of light has been used in a capsule transducer,
but this is a technique that has been superseded.

173
CHAPTER 6. PRESSURE TRANSDUCERS

A more modern light technique is to measure the deflection of a diaphragm using a bundle of
optical fibres whose ends stop just short of the diaphragm [6]. Half of these carry light from
light-emitting diode to the diaphragm and the other half carry light reflected from it away to a
photo detector. The diaphragm movement changes the amount of light received by this
detector whose output then becomes a measure of the pressure that moves the diaphragm.
These transducers ranging down to 2 mm in diameter are used in clinical work for
measurement of the pressure of body fluids. Here the absence of electrical leads that can carry
small leakage currents into the body is a distinct advantage. They are also very resistant to
corrosion.

Oscillating diaphragm
If a diaphragm is displaced and allowed to return to its undisturbed position, the speed with
which it does so depends not only on the elasticity of the diaphragm but also on the pressure
acting. A pressure transducer has been constructed using this principle [14], in which a piezo­
electric device displaces the diaphragm and another senses the displacement. This
arrangement can be made to oscillate using an amplifier and the frequency output is then
dependent on the pressure acting on the diaphragm. The frequency range is only 1324 to 1336
Hz but since frequency can easily be measured to great accuracy this is not a limitation to
accuracy. This device can be compared with the oscillating cylinder described later. For both
devices the relationship between frequency and pressure is non-linear.

A similar, but older device [11], used a vibrating contact, similar to an electric bell, attached to
the diaphragm. In this device the ratio of the time the contact was closed to the time it was
open was used to measure the pressure. This type of signal is easily sent long distances without
appreciable error.

Bellows
Bellows might well be considered a super corrugated diaphragm. Like diaphragms the
flexibility depends on the diameter, more precisely the stroke per unit load varies as the square
of the diameter. The flexibility also varies directly as the number of convolutions per unit
length and inversely as the cube of the wall thickness and the modulus of elasticity of the
material.

A bellows exerts a force when pressurised. The relationship between pressure, P and force, F,
can be thought of as if the bellows had an effective area A^ff so that

F = PA^^ . (6.7)
The value of Aeff- is approximately equal to the area of a circle whose diameter is halfway
between the maximum and minimum diameters of the bellows, D and d, so that

4, j. (6.8)

174
NOEL BIGNELL

Potentiometric. The force due to the pressure is balanced by the spring action of the bellows
as it expands so that

PA^ff =kx or P =----- (6.9)

where k is the spring constant and x the distance moved. To measure x, and hence P, a moving
arm can be made to move over a resistance element, suitably energized, as shown in
Figure 6.14. The output voltage should be connected to a high input impedance device such as
a DMM to preserve linearity.

The advantage of potentiometric pressure transducers is that they are inexpensive, both in
themselves and in the auxiliary equipment to handle the output signal. Their disadvantages are
electrical noise as the wiper moves over the element and static friction (or stiction) which
limits their usefulness, with slowly varying pressures, to 0.5% FS. Sometimes they are vibrated
to minimize this effect.

PRESSURE POTENTIOMETER MECHANICAL


IN AMPLIFICATION

Figure 6.14: Movement of the bellows caused by the pressure is amplified by the lever
arm and detected by the movement of the tapping point on the
potentiometer resistor to give an electrical output.

Oscillating Quartz Force Transducer. In one type of high quality pressure transducer, the
force produced by the pressure in the bellows is measured by a force transducer [12]. This
device uses the oscillation of a quartz crystal beam, the frequency of which changes with
applied force, reducing for compressive and increasing for tensile forces. A sketch of the
device is given in Figure 6.15. The balance weights are positioned to place the centre of mass
of the moving parts at the pivot point so that vibration has little effect. The quartz force
transducer is very stiff so that the bellows does not have to move very far. This is good
because hysteresis depends very much on the strain experienced by the elastic device - the less
strain the less to recover from. For this device repeatability and hysteresis are 0.005% FS.

175
CHAPTER 6. PRESSURE TRANSDUCERS

PIVOT

QUARTZ CRYSTAL
FORCE TRANSDUCER

BELLOWS

PRESSURE PORT

Figure 6.15: The force produced by the pressure in the bellows is transferred by levers to
a quartz force transducer for measurement.

Because the frequency of oscillation is a complex function of the force, the pressure is not
linearly related to the output frequency. Usually the relationship is expressed as a function of
the period r(the reciprocal of the frequency ) thus

P = A^Bt + Ct- . (6.10)

The task of calibration is to determine the constants A, B and C usually from a least squares
fitting of the equation to the calibration data. A package of electronics is available which both
measures the period and calculates P using the equation above, with stored values of A, B and
C. Alternatively, a frequency counter operating in period mode and a small computer can be
used. Some transducers of this type have corrections for temperature variation available for use
mainly outside the laboratory where the temperature is likely to quite different from the
calibration temperature. A full calibration of these parameters is a difficult and specialised task
and usually the manufacturers’ values are taken.

Cylinders
When a closed end cylinder is subjected to internal pressure P there are two strains, one
circumferential, £„ and the other longitudinal, s/. These are coupled because of Poisson’s ratio,
V, which links a strain in one direction with a strain at right angles.

For tubing of outside to inside diameter ratio d?and Young’s Modulus E


P
(6.11)
0) E
2P
(6.12)

Due to Poisson’s ratio there is a longitudinal contraction


(6.13)
~ ((6y^-1)£)

176
NOEL BIGNELL

so that the total longitudinal strain is

E 1^69^ (6.14)

This is zero when co = {l/(l-2v))‘^ which for v = 0.3 is for co = 1.581. Thus, the
circumferential strain rather than the longitudinal strain is the more reliable for measurement
of pressure.

Strain Gauges. The strain can be measured by applying strain gauges in a bridge
configuration to the outside of internally pressurised cylinders (or to the inside of externally
pressurised ones). To overcome problems with cancellation of strain, the cylinder used is
often equipped with moving seals at the ends so that there is no longitudinal stress. Such a
transducer is shown in Figure 6.16. These devices are capable of working at very high
pressures, at least to 1.5 GPa, and with accuracies of 0.2% FS.

Figure 6.16: A high pressure gauge using the measurement of external strain in an
internally pressurised cylinder.

Oscillating Cylinder. A thin-walled cylinder can be made to vibrate in a four-lobed hoop


mode, indicated in Figure 6.17, by means of electromagnetic pick-up and drive coils placed at
right angles. (A similar system is used in certain gas densitometers). The frequency of
vibration, f, depends, in part, on the circumferential stress, which, of course, depends on the
pressure P inside the cylinder. The relation between them is of the form

/’ = -4(/-/o) + S(/-/o)' + C(/-/o)’ (615)

with A, B and C constants and/, the frequency when P is zero. The stability and particularly
the hysteresis as claimed are good (<0.01 %) [13].

177
CHAPTER 6. PRESSURE TRANSDUCERS

Figure 6.17: The mode of oscillation of the cylinder in an oscillating-cylinder pressure


transducer.

Ultrasonic pulse technique


The velocity of ultrasonic pulses in liquids or solids depends on pressure and can be used as a
pressure gauge. For example, in one transducer [7], pulses generated at one end of a quartz rod
travel down the rod, reflect from other end, and return. This arrangement can be made to
oscillate continuously at a frequency dependent on the transit time, r, of the pulse, which is
given by

Ipr A 1 p aM
(6.16)
t[dPJ 2{m dP 12J
where M is the appropriate modulus of elasticity and K is the isothermal compressibility. The
transit time is also sensitive to temperature but, in some materials, e.g., aluminium, it is
positive and in others, e.g., quartz, it is negative, so that by using some of each in a compound­
gauge, temperature compensation should be possible.

6.7 Transducers not using elastic deformation


All the transducers in Section 6.6 use the elastic deformation of some material to measure
pressure. There are some transducers that do not use deformation. There are some advantages
in this because always, if looked at sufficiently closely, the deformation is not elastic and this
means non-linearity and hysteresis in the output. Generally speaking, these effects increase as
the deformation increases though the amount will depend on the material. The oscillating
quartz transducer discussed in previous sections and shown in Figure 6.15 owes much of its
good performance to the stiffness of the force transducer so that the bellows hardly has to
expand at all. Taking the logical next step gives the idea behind the force feedback technique.

6.7.1 Force feedback or force balance

Consider the simple pressure transducer shown in Figure 6.18 using a diaphragm and LVDT to
detect its movement. On the other side of the diaphragm is attached a small magnet which sits
in a coil so that current through the coil will push or pull it. The signal from the LVDT, when
amplified, can produce a current through the coil that causes the magnet to pull the diaphragm

178
NOEL BIGNELL

in the opposite direction. With suitable electronics the system can be made to look very stiff
and so the diaphragm does not need to move very much at all.

This technique of force feedback improves the performance of pressure transducers up to about
ten times. It has also been applied to diaphragms using a capacitive technique, both to
measure and to apply the force by putting high voltages on the capacitor plates so as to attract
the diaphragm back to its null position [4]. It has been applied to improve the characteristics
of quartz Bourdon tubes and to bellows.

COIL OUT OF
CURRENT BALANCE
TO RESTORE SIGNAL
BALANCE

AMPLIFIER

Figure 6.18: A force-feedback transducer in which the pressure force on the diaphragm
is cancelled by the magnetic force between the magnet and the coil.

The use of this technique involves quite a lot more electronics than in transducers based on
elastic deflection and they are correspondingly more expensive. Though the cost of electronics
is decreasing the use of materials with superior elastic properties, especially quartz, seems to
be dominant at the moment. The two techniques have been combined in at least one instrument
that uses a quartz spiral Bourdon tube.

6.7.2 Pressure-sensitive resistance transducers

Pressure can act directly on a resistor to alter its resistance, which can thus be used as a
pressure transducer. This happens to a useful extent only at high pressures where the manganin
resistance gauge is a much-used secondary standard. Manganin is an alloy used for making
resistors because its resistance is insensitive to temperature. Other alloys, such as zeranin, have
also been used as pressure transducers.

The resistance change for manganin is about 2.5 x 10 *’ Pa’’ which is small so that for accurate
work to precisions of 1 kPa, expensive resistance bridges are needed, as is good temperature
control even for low temperature coefficient alloys such as manganin.

179
CHAPTER 6. PRESSURE TRANSDUCERS

6.7.3 Semi-conductors

Semiconductor diodes have been used for some time. Tunnel diodes have high sensitivities to
pressure (l/V)dV/dP = 290 x lO'^Pa’’) up to 400 kPa and Zener diodes have been used to
measure pressures up to 2GPa with sensitivities about 1000 times less. More recently, a light
emitting diode (LED) has been used for a dynamic pressure transducer with sensitivities of up
to 440 X 10'’"Pa ' and very fast response [8].

6.7.4 Piezo-electric transducers

The piezo-electric effect, discovered in France by the brothers Curie in 1880, is the appearance
of charge on the surfaces of certain crystals when they are mechanically stressed. Quartz is the
most used piezo-electric material for pressure transducers because it is stable, has a high
mechanical strength, and can operate at high temperatures. Other materials used are lead
zirconate titanates and special ceramic materials. Such transducers consist only of a diaphragm
to transmit the force due to the pressure on to the crystal, so that they are essentially very
simple devices. For best effect, however, they need to be used with charge amplifiers that
convert the electric charge to a voltage. Because the charge leaks away they do not respond to
static pressures but quartz types can measure slowly varying (0.1 Hz) and rapidly varying
(IMHz) pressures from about 200 Pa to 700 MPa over temperatures from -196°C to 350°C.
They are used in measuring explosive pressures including those in internal combustion
engines.
6.7.5 Pressure-sensitive capacitance transducers
A gauge for use at high pressures using the change in dielectric constant of a crystal of calcium
fluoride has been described [1]. It was used up to 250 MPa with an accuracy of better than
0.01%. The technique involved the accurate measurement of the capacitance of a capacitor that
used the crystal as a dielectric.

6.8 Example specification sheet


This section contains a composite specification sheet. All the entries in each of the sections are
from real specification sheets, but not for the same transducer. The actual wording is given in
bold type and the explanation below it is in ordinary font.

6.8.1 ‘NML’ industrial pressure transmitters

We have used the house brand to avoid endorsing (or otherwise) specific manufacturers. The
term “transmitter” is another term for transducer. Note that not all manufacturers will give a
specification for all of the headings below but each of them is taken from a manufacturer’s
specification sheet. When something is not given, assume the worst until you can find definite
information, for example accuracy may be quoted as 1 % or it may be quoted as 1 % of reading
or 1 % FS (full scale). The latter is a worse specification and it is almost certain that when FS
is left out then it will refer to FS and not to % of reading. Similarly, unless otherwise stated.

180
NOEL BIGNELL

take the uncertainty as a standard uncertainty as this will give a smaller value than the
expanded uncertainty and so will be more likely to be quoted by the manufacturer.

Operating pressure ranges:


Full scale (FS) from 100 mbar to 70 bar gauge or absolute
The transducer is intended to measure from zero to the pressure chosen as full scale. “Gauge”
means that the pressure is referred to atmospheric pressure and “absolute” means that it is
referred to an inbuilt vacuum.

Pressure units:
psi, mbar, bar, hPa, kPa, MPa, mmHaO, torr, kgf/cm^, cmH2O, mHiO, inH2O, ft, mmHg,
kg/cm^
The variety of units offered does not mean very much for ordinary voltage or current output
transducers unless there is supposed to be a simple relationship between, for example, the
voltage and the pressure, such as 1 volt is equivalent to 1 kPa. For digital instruments this is
usually the case, that is, the instrument displays the pressure directly in the units specified. It is
easy for the manufacturer to offer a wide range of units, so they do. Sometimes the units
displayed are user selectable.

Burst pressure (or pressure containment):


10 X FS for ranges below 500 psi
5 X FS for 501 psi to 3000 psi
This means that the manufacturer does not guarantee that the transducer will not burst, or leak
more likely, if it is pressured above this figure. It does not mean that the transducer will not be
damaged or its calibration changed by operating below this maximum figure. The next
specification item gives this.

Overpressure (or overpressure limit):


The operating FS pressure can be exceeded by the following multiples with negligible
effect on calibration:-
5 X for ranges above 500 mbar up to 2 bar
This is clear as to meaning though sometimes it is not as clearly put as in this example.

Wetted materials (or pressure media):


Glass, silicon. Aluminum (sic) Valox, Silicone, Nickel plated. Brass
(Fluids compatible with 316L stainless steel)
These are two alternate ways of expressing the same idea, namely that you must not put
corrosive things into the pressure transducer. Some manufacturers have a wide range of
materials that the pressure transmitting fluid comes into contact with and this limits their use to
simple gases such as air or nitrogen, or to liquids such as good quality oil. Other manufacturers
have gone to some trouble to produce a transducer that has perhaps an all-welded construction
of inert materials so that corrosive fluids may be used with it.

Accuracy:
Combined effects of non-linearity, hysteresis and repeatability
±0.1% FS BSL typical (±0.2% FS BSL max)

181
CHAPTER 6. PRESSURE TRANSDUCERS

0.15% TSL typical (0.3% TSL max)


Here the ways of expressing non-linearity mentioned in section 6.3.2 have been used. The
specification has combined the three effects of non-linearity, hysteresis and repeatability. This
is not always done and it is wise to find out what the components in the accuracy specification
quoted are, as in for example:
1%
This is a vague specification. Does the 1% refer to full scale or to reading? What is included in
the uncertainty? Very likely the performance quoted will only be achieved at room
temperature and the degradation in performance if addressed at all will be given by another
item in the specification such as that below. For digital instruments there is an obligatory ±1
digit to be added to the specification.
There may be a statement such as:
All accuracies include 90 day stability
This is important when calculating the uncertainty of the pressure measurement made with the
device. An alternative statement is:
Long term stability
±0.1 % FS per year
If the transducer is used, say, 6 months after calibration, then the uncertainty component for
this item will be half of this value, 0.05% FS.

Ambient operating temperature (or operating temperature range)


0°C to 50 °C
This is just the range of temperatures over which the transducer should be able to operate
without destroying itself or refusing to operate. This does not mean that the transducer will
operate equally well over all of this range. There will almost certainly be some temperature
effects or temperature coefficients as indicated below.

Temperature effects:
1. Output will not deviate from room temperature by more than:
1.5% FS typical (2% FS max) over -20 to 80 °C
2. Temperature Coefficients
Zero: 0.02% of F.S./°C, 50 psia to 3000 psia;
Span: 0.04% of Rdg./°C, 50 psia to 3000 psia.
These two specification items from different manufacturers refer to the same phenomenon, that
is, the temperature will change the reading on the pressure transducer. The first example does
not quote a value for the change in the zero. This may not always be important, for example
the zero may be able to be adjusted at the operating temperature.

Which of these two specifications is the better? If we take a transducer with a FS of 50 psia
and a temperature of operation of 80 °C, or 60 K above its calibration temperature, then for the
first one we have a (maximum) change of 2% FS or 1 psi. For the second specification we
have 0.04% for each degree away from the calibration temperature, that is, 60 x 0.04% of 50
psia or 1.2 psi. If on the other hand we were operating at a temperature much closer to the
calibration temperature, at say 30 °C, we have only 0.2 psi change for the second one and
apparently the same maximum change as before for the first one, 1 psi. It is likely however that

182
NOEL BIGNELL

the first transducer will also behave better at a temperature of 30°C than at 80°C but this is not
specified. Further information could be sought from the manufacturer. (Here we have used
degrees kelvin, or K, to indicate a temperature difference and “psi” not “psia” to indicate a
pressure difference. This might be a bit pedantic but when the scales have real zeros then
differences are related to the size of the unit not to the scale. It is as if we spoke of the distance
of Brisbane to Sydney as 1000 (Bris. to Syd.) km so that when we have 900 (Bris. to Syd.) km
we know we are in Newcastle and 100 (Bris. to Syd.) km puts us in Tweed Heads. It would not
be correct then to speak of the distance of Newcastle to Sydney as 100 (Bris. to Syd.) km
instead of just 100 km.).

The zero change will probably add to (or perhaps subtract from) the change due to the span or
sensitivity change. If the transducer cannot be zeroed at the operating temperature then this
should be clarified with the manufacturer. It is likely that for temperatures above the
calibration temperature the zero will change in the opposite way to temperatures below the
operating temperature. It is also possible that the span change may behave in a similar way.

Supply voltage or input power required:


+13 VDC to +32 VDC @ 10 mA max.
This is fairly straightforward. The supply can have a wide range because there is an internal
regulator that converts the input to the real supply voltage used in the circuit inside. The
minimum value has obviously to be several volts above the maximum output voltage and the
maximum value is to prevent the destruction of the internal regulator.

Electrical outputs:
0 to 10 VDC
4 to 20 mA (2 wire) proportional to the zero to F.S. pressure range
There are many possible output quantities but these two are very common.

Supply sensitivity:
0.005% FSA^olt
There is no reason why this should be a problem for transducers today as the input supply will
pass through a regulator and it is usually easy to have the input supply constant to ±1 V or
better. Thus, for the 50 psia transducer used previously there would only be an effect of 0.0025
psi or about ten times less than the effect of a 1 K change in temperature.

Response or pressure response:


2 readings per second
1 kHz band width (63% response to step change in pressure)
The first of these will be for a device that has a digital readout capability and is related to the
digitisation time. This is the time that the electronics package takes to convert the analogue
output signal from the transducer into digital form. The actual transducer will probably have a
much faster response, perhaps as indicated by the second version of this specification item.
The second version relates to the frequency response curve shown in Figure 6.4. The 1 kHz
value is determined from the reduction in response as the response is plotted from static or
zero frequency applied pressure to higher frequencies. It can also be determined as indicated
here (in the brackets) from the response to a step change in applied pressure.

183
CHAPTER 6. PRESSURE TRANSDUCERS

Readout:
±999999 or 19999 or 3.99
For digital instruments such as pressure calibrators this gives the number of figures shown on
the display. Sometimes a “half digit” is shown as “1” as in the second example above and this
is referred to as 4^2 digits. The third specification is referred to as 2% digits. It should never be
assumed that all the digits are meaningful.

6.8.2 Other items covered in specifications

There will be a range of other things mentioned in the specification sheet. There will certainly
be information on the physical size, the type of pressure fitting, the type of electrical
connections for the input and the output, the weight, and several other items that should be
self-explanatory. Digital devices such as pressure calibrators will have some information on
the readout options and perhaps the connections that can be made to a computer such as an
RS232 bus.

Sometimes the internal volume of the pressure transducer is important and will be specified by
the manufacturer. If the physical characteristics of the pressure fluid such as its density or its
viscosity are important they may be specified but this is not usually the case. The effect of
position or orientation may also be specified and here the reading may change more for a
liquid acting on a diaphragm than for a gas.

It is not usual for the manufacturer to indicate whether the uncertainties quoted are standard
uncertainties or uncertainties at the 95% level of confidence or some other. When they give a
value that says “max” after it then this can be taken to be at the 95% level. Otherwise they will
probably be standard uncertainties as this gives a smaller number.

6.9 Pressure transducers in pressure measurement


The uses to which pressure transducers may be put are extremely diverse. They go under
oceans to measure both depth, through the relation P = pgh, and wave motion through pressure
changes caused by depth changes. They are used to probe our bodies to reveal the pressure of
blood and other vital fluids. They are more and more used in motor vehicles to continuously
monitor performance and to make adjustments to the operating conditions as a consequence of
those measurements. Differential pressure transducers are widely used in measuring flows of
liquids and gases through the use of orifice plates, venturis and laminar flow elements. In
industry their use is ubiquitous. No nuclear test explosion occurs without the presence of
dozens of pressure transducers to monitor its progress.

Pressure calibrators incorporate one or more pressure transducers in conjunction with other
equipment to allow the convenient calibration of lower grade pressure instruments. These
instruments range in accuracy from about 0.01% to a few percent with a price range inversely
matching. They are convenient to use and are often able to be interfaced to a computer. Some
are pneumatic equivalents, in size and general usefulness, to hand-held digital multimeters.

184
NOEL BIGNELL

Some are specifically designed to measure barometric pressure and to compete with aneroid
barometers and cheaper mercury barometers.

Another variation acts as a reverse pressure transducer. This is a device which, when given a
supply pressure input, produces a variable output pressure in response to an electrical input
signal. It may be thought of as an electrically controllable pressure regulator and, in
combination with an ordinary pressure transducer, allows an instrument to be produced which
gives a pressure output corresponding to a digital input, either entered with dial-up switches as
in a resistance box, or as a set of electrical pulses. It is the pneumatic equivalent of a variable
voltage reference.

All of these pressure transducers eventually need calibration or, rather, re-calibration since
they all must be calibrated before they are useful for pressure measurement, and though some
types exhibit remarkable stability none can be relied upon indefinitely. The glossy exterior
displaying many digits should not distract us from the black-box approach to such devices and
the possibly erroneous relationship between measurand and output, digital or otherwise.

6.10 Uncertainty of pressures measured with a pressure transducer


As indicated in the specification sheet in section 6.8, there are a number of sources of
uncertainty in the use of a pressure transducer. There are of course also some in the calibration
of a pressure transducer (see section 4.7.3). As an example let us assume that we have
measured a pressure of 50 kPa with a digital readout transducer of full scale capacity of
100 kPa that was calibrated 9 months ago at a temperature of 20 °C with a standard that has a
uncertainty of ±0.01%. The zero of the transducer was checked before the measurement that
was performed at 25 °C. The standard was used to set the maximum response of the device to
read correctly at that time.

The approach taken when considering and combining the components of uncertainty for this
measurement are as given in the ISO Guide to the Expression of Uncertainty in Measurement
[15]. The general procedure and terminology are discussed in [16].

Once the components of uncertainty are assessed (see below) they are combined to give Uc, the
‘combined standard uncertainty’ for the pressure being measured, and given by:
=^Ic.m(x.)|2 ,
/
where n(x,) is the standard uncertainty for the error source and c, is the sensitivity coefficient
for that parameter that it affects. In this case, all c, = 1. The value of w(x,) is obtained as UJki,
where U, is the initial, raw estimate of uncertainty and k, is an appropriate reducing factor [16].

The various components of uncertainty are:


(1) The resolution of the readout, ±1 digit, i.e., U, = 1 Pa. This has a rectangular distribution,
so the reducing factor A:, = 1.73, and the number of degrees of freedom is v, = 1000.
(2) An accuracy specification from the manufacturer of 0.15% FS TSL. The FS is 100 kPa so
this gives an uncertainty of Ui = 150 Pa, considered to be a standard uncertainty, so kj = 1.

185
CHAPTER 6. PRESSURE TRANSDUCERS

(3) The uncertainty in the pressure standard used to set the maximum pressure or the top of
the terminal straight line (TSL). Its calibration report gives an expanded uncertainty of
0.01% with a coverage factor of 2.04, equivalent to 30 degrees of freedom. So, at half the
top value, the uncertainty component is V2 x 0.01% x 100 kPa or 5 Pa. Thus, we have that
Ui = 5 Pa, ki= 2.04 and v, = 30.
(4) Effect of temperature coefficient of the transducer — the device is used at 25 °C or 5 °C
different from the calibration temperature. From the manufacturer’s specification, the
coefficient is within ±0.002% of rdg/°C, which amounts to ±0.002% x 50 kPa x 5, or
±5 Pa . This is considered to be a standard uncertainty, so k, = 1, and the estimate is
reasonable (Vi = 8).
(5) Drift in calibration over 0.75 of a year since last calibration. According to the
manufacturer’s specification, the stability would be 0.1%/year, which is considered a
standard uncertainty (kj =1). So, Uj = 0.1% x 0.75 x 50 kPa or 37.5 Pa and we take
Vi = 8.
(6) Rounding of the final result. The result was rounded to the nearest 10 Pa, so Ui = 5 Pa,
ki= 1.73, being rectangular, and Vj = 1000.

The above data for the various components of uncertainty have been listed in Table 6.1, where
the combined expanded uncertainty is calculated. It is clear that the dominant uncertainty
component is the one derived from the accuracy specification.

Table 6.1 Uncertainty analysis table for a transducer used to measure pressure.

Component U,(Pa) Vj Ci \Ci.u ( x,)l \Ci.u( X/)P \Ci.u( x,)lVv/


1. Resolution of the readout 1 1.73 1000 1 0.6 0.3 1.12E-04

2. Accuracy specification 150 1 8 1 150 2250 6.33E+07

3. Calibration of standard 5 2.04 30 1 2.4 6.0 1.20E+00

4. Temperature coefficient 5 1 8 1 5 25 7.81E+01

5. Instability/drift 37.5 1 8 1 37.5 1406 2.47E+05

6. Final rounding 5 1.73 1000 1 2.9 8.4 6.98E-02

Sums 23946 6.353E+07


Vgif = 9.03
Uc = 154.7 Pa
k =2.27 Uc^ = 5.73E+08
U = 351.2 Pa

This expanded uncertainty should be quoted with two significant figures, i.e., 350 Pa (see
section 4.4 on Rounding in [16]).

186
NOEL BIGNELL

If the accuracy specification was 0.1% FS BSL (or best straight line), then the uncertainty
component associated with this is found from 0.1% x 100 kPa or 100 Pa. This is smaller
because it implies that more values have been taken in the calibration of the transducer than
just setting the terminal or top value.

If a series of calibration pressure points has been used to calibrate the transducer then it may be
possible to generate a calibration curve. This is more work but will give the best possible
uncertainty from the device (see page 106). Sometimes the calibration laboratory will do this
and a figure will be given for the uncertainty that can be expected when pressures are
measured using the transducer (e.g., the specimen calibration report from the National
Measurement Laboratory given in section 6.11).

187
CHAPTER 6. PRESSURE TRANSDUCERS

6.11 Example report on a pressure transducer

Report
on

Absolute Pressure Transducer


Type xxxxx
Serial Number xxxxxxx

The tests, calibrations and measurements reported in this document have been
performed in accordance with NATA requirements, which include the requirements of
ISO/IEC Guide 25:1990. Accreditation Number 1.

For further information contact: Dennis Tyrrell telephone (02)9413 7360


Facsimile (02)9413 7238
email dennis.tyrrell@tip.csiro.au

Ref- RN 45368 File: Checked: Date: Day Month 2003

188
NOEL BIGNELL

For: CSIRO National Measurement Laboratory

Reference: Internal request

Description: A 2000 kPa capacity absolute pressure transducer, which is


powered by an 18 volt plug pack power supply. Output
from the transducer is a 0 to 10 volt signal and is measured
by a digital multimeter.

Manufacturer: Transducer: TXTXTXTX


Multimeter: DMMDMMDMM

Instrument Markings: Transducer: Type: XXXXXXX


Serial number: 78910
Multimeter: Model: XXXXXXX
Serial Number: 123456

Date of Test: October 2002

Test Method

The instrument was connected to a pneumatic pressure balance that was used to apply gauge
pressures of known uncertainty. A digital manometer was used to measure the ambient
pressure during the calibration. A series of twelve increasing calibration pressures and twelve
decreasing calibration pressures was applied to the instrument. This was repeated. Voltage
output from the transducer was measured by the digital multimeter and recorded for each
calibration pressure together with the ambient pressure indicated by the digital manometer.
The uncertainty associated with the applied calibration pressures is not greater than ± 30 ppm.
The pressure fluid was nitrogen.

Results:

Calibration equations relating instrument readings and applied pressure were computed and are
given below.
R = p2 X 7.11591X10’Q + P X 4.996977 x 10’® + 6.49866 x 10'^

P = R2 X (-0.056523435) + R x 200.118808 - 0.0112096

Where:
R = Transducer output, V
P = Absolute pressure, kPa

189
CHAPTER 6. PRESSURE TRANSDUCERS

Uncertainty

The uncertainties asscx:iated with the transducer output and the equations given in this report
are estimated to be ±0.14 kPa.

Notes

1. During the calibration the transducer was mounted vertically with the pressure inlet
connection at the bottom.

2. The absolute pressure transducer has been calibrated in the range from 100 kPa to
2000 kPa absolute pressure. The results given in this report are valid only within this
pressure range.

3. The uncertainties stated above are valid only if the transducer output is measured by
the digital multimeter described in this report.

4. A coverage factor k=2.0 was used to calculate the uncertainty associated with the
measurement result.

5. The calibration of this instrument was carried out according to the Test Method and
Work Instruction detailed in the NML Project operations manual EAFA - Mass and
Related Quantities. The relevant Test Method for this calibration is EAFA 8.2.16
“Calibration of Pressure Transducers and Bourdon Gauges” and the relevant work
Instruction is EAFA 9.2.21 “Pneumatic Pressure Procedures”.

h. The uncertainty has been calculated in accordance with the principles in the ISO Guide
to the Expression of Uncertainty in Measurement, and gives an interval estimated to
have a level of confidence of 95%.

7. The stated uncertainty applies at the time of test only and takes no account of any drift or
other effects that may apply afterwards.

John Man Noel Bignell


for Dr B. D. Inglis NAT A Approved Signatory
Director
National Measurement Laboratory

190
NOEL BIGNELL

6.12 References
[1] Audeen, C., Fontanella, J. and Schuele, D. (1971) Rev. Sci. Instrum. 432, 495.

[2] Bergveld, P. (1985) Sensors and Actuators 8, 109.

[3] Bignell, N. (1976) J. Sci. Instrum. 9, 730.

[4] Cope, J.O. (1962) Rev. Sci. Instrum. 33,980.

[5] Gieles, A.C.M. and Somers, G.H.J. (1973) Philips Tech. Rev. 33, 14.

[6] Hansen, T-E. (1983) Sensors and Actuators 4, 545.

[7] Heydemann, P.L.M. (Sept. 1967) J. Basic Eng. P. 551.

[8] Kruger, A. (1985) J. Phys. E: Sci. Instrum. 18, 944.

[9] Lederer, P.S. (1967) Instrum. Control Systems, 40, 93.

[10] Meyer, R.C. (1972) Instrum. Soc. Am. Trans. 12, 156.

[11] Neubert, H.K.P. and Price, E.F. (1969) Instrum. Control Systems, 42, 81.

[12] Paros, J.M. (1972) Instrum. Soc. Am. Trans. 12, 173.

[13] Quinn, E.J. and Posipanko, T. (1974) Rev. Sci. Instrum. 43, 475.

[14] Smits, J.G., Tilmans, H.A.C., Hoen, K., Mulder, H., van Vuuren, J. and Boom, G. (1983)
Sensors and Actuators 4, 565.

[15] ISO Guide to the Expression of Uncertainty in Measurement. International Organization


for Standardization, Geneva, 1995. ISBN 92-67-10188-9.

[16] R.E. Bentley. Uncertainty in Measurement'. The ISO guide. CSIRO, Sydney, Monogr. 1:
NML Technology Transfer Series, 2003.

Further reading
• Electrical Transducer Nomenclature and Terminology, Instrument Society of America,
ISA-S37.1 (1969).

• Glossary and Terms in Metrology, AS 1514, Part 1 (1980).

• Dynamic Calibration of Pressure Transducers, Guide for ANSI MC88.1 - 1972 (1978).

191
CHAPTER 6. PRESSURE TRANSDUCERS

Piezoelectric Pressure and Sound-Pressure Transducers, Specifications and Tests


for, ANSI/ISA 537.10 - 1969.

Potentiometric Pressure Transducers, Specifications and Test for, ANSI/ISA S37.6 -


1975.

Strain Gauge Pressure Transducers, Specifications and Tests for, ANSI/ISA S37.3 -
1970.

Sydenham, P.H. Transducers in Measurement and Control. - 3*^ ed. (Adam Hilger, UK).

The periodical “Measurements and Control” regularly features surveys of commercially


available pressure transducers and gives approximate prices and short specifications.

192
Chapter 7

Differential pressure measurement


Noel Bignell

7.1 Introduction
The measurement of small differential pressures, in the presence of much larger line pressures,
is an important measurement in many industrial and commercial activities especially those that
involve the measurement of the flow of liquids or gases using orifice plates, or other flow
impedances, or Pitot tubes. It is not a difficult measurement to make with any of the
differential pressure transducers on the market. Establishment of suitable standards of
differential pressure so that these transducers may be calibrated is, however, a difficult task.
The pressures considered here are those appropriate to flow measurement with line pressures
up to 30 MPa and differential pressures down to 50 Pa full scale (FS).

7.2 Differential pressure measurement using liquid columns


In many applications water or oil columns are convenient. The inch of water (“H2O or in WG)
is certainly not an SI unit but is commonly seen on specification sheets. The mmHg is also
frequently seen and mercury columns are also used for differential pressure measurement,
though with a loss of sensitivity.

Even water columns lack sufficient sensitivity for many applications and ways of improving
the sensitivity of ordinary tubes and scales have been developed. Chapter 3 discusses some of
these.

7.2.1 Inclined tube

It is the vertical height, h, of the liquid that enters into the expression for the pressure, P,

P = pgh (7.1)

where p is the density of the liquid in kg/m'^ and g is the acceleration due to gravity, in m/s".
By holding the tube at an angle 0 to the horizontal the actual distance the column travels along

193
CHAPTER 7. DIFFERENTIAL PRESSURE MEASUREMENT

the tube is magnified by 1/sin^ so that the distance the liquid surface or meniscus moves
becomes

Vsin^J
Though in principle I could be made as large as required the meniscus becomes harder to
define and the length / harder to measure as 3 becomes smaller. Chapter 3 contains a
description of a Bradshaw gauge that uses this principle.

7.2.2 Pressure correction

When using liquid columns at high line pressures there will be a change in the density of the
liquid due to its compression. This can be corrected for, but as the effect, for water, amounts
to only 0.05% at 1 MPa it can mostly be neglected. For mercury the effect is even smaller.
Data to enable the correction to be made are to be found in [4]. The use of glass tubing at
pressures as high as 1 MPa should be avoided as bursting, even after hours of satisfactory use
at higher pressures, can occur spontaneously.

7.3 Differential pressure transducers


Pressure transducers are generally more convenient than liquid columns and also offer greater
range and sensitivity. Their ability to be used at very high line pressures is usually superior
and they are safer than liquid columns. Generally, differential pressure transducers use
diaphragms, sensed capacitively, magnetically or with strain gauges. The various types of
pressure transducers are dealt with in Chapter 6.

Some transducers are built into a unit that includes the flow impedance and pressure tapping
points so they can be sold as flow meters. Some have the square root function, which is needed
to obtain the flow from the differential pressure, also built in. When the transducer output goes
to a computer there is no need for this square-root facility, or for the external linearisation of
those transducers giving pressure as a quadratic function of frequency as these transformations
can all be performed by a suitable program.

The tubing used to measure low differential pressure at high line pressures needs to be rigidly
mounted because small volume changes can produce large transient pressure changes, on the
scale of magnitude of the differential pressure that is to be measured, as the system readjusts
itself. Similarly, variations of temperature in the lines connecting the transducer to the source
can cause erroneous pressure readings. The use of well-secured tubing with the two lines
placed side by side is recommended.

Long lines and large volumes will attenuate changing pressures, but not static pressures, so, if
it is required to follow pressure changes, lines and volumes should be kept small. On the other
hand, there can sometimes be a useful smoothing effect introduced by the volume in the
system, but care must be taken that its temperature is stable. It is best in this context to have

194
NOEL BIGNELL

equal volumes on the two sides of the pressure transducer and placed close to each other so
that they experience the same temperature fluctuations and these effects will then cancel.

7.4 Standards of differential pressure at high line pressure


It is a difficult task to establish accurate differential pressure standards. The problem arises
because the sensitivity of differential pressure transducers is such that they can easily detect
pressures of 1 Pa in the presence of line pressures 10 million times greater. It is impossible to
maintain pressures of 10 MPa constant to a part in 10 million let alone to measure them to that
accuracy.

There is, however, one differential pressure easily and accurately obtainable at high line
pressure, and that is zero differential pressure. Using this “standard”, we can see that the zero
of many differential pressure transducers changes with line pressure. The changes in the
transducer, which cause the zero to change, also cause the sensitivity to alter, by as much as
1% for line pressures as low as 3.5 MPa, so the often made assumption that the atmospheric
pressure calibration is satisfactory, with a zero adjustment at high line pressure, is invalid. In
this section various ways of establishing standards are discussed.

7.4.1 Two separate pressure standards

A requirement for a differential pressure standard of AP at a line pressure of P is a requirement


for two pressures P and P+AP and a straight-forward approach is to use two separate
standards, one set at Pi=P and one set at P2=P+PAP. What uncertainties could be expected
from this approach? For P in the range up to 30 MPa, piston gauges would have to be used so
that we have
n
‘ A
(7.3)
p
’’ A,

The differential pressure generated is just the difference between these

The uncertainty in AP is given by the combination of the uncertainties in Pi and P?- The typical
percentage uncertainty offered by manufacturers of piston gauges is often 0.03%, which at
(say) 10 MPa means an uncertainty of 3 kPa for each standard and, hence, the combined
standard uncertainty Uc (terminology discussed in [5]) is
u’ = Vsooo^ +3000’ = 3OO0V2 = 4243ft/. (7.5)
The differential pressure is likely to be less that 10 kPa and so this approach is totally
unsatisfactory. A better quality piston gauge of 0.01% would reduce the uncertainty by a
factor of three to about 1.4 kPa or a 14% relative uncertainty. This approach also is not useful.

195
CHAPTER 7. DIFFERENTIAL PRESSURE MEASUREMENT

7.4.2 Two balanced piston gauges


Suppose we do an initial balance of the two piston gauges so that AP is zero for the weights m\
and m2 as in equation (7.4). By opening the valve shown in Figure 7.1 joining the two piston
gauges, and with the other valves closed, the weights m\ and m2 can be adjusted to give an
exact pressure balance so that m\g/A\ = m2^/A2. Then if we add a small additional mass Am to
the (say) second piston the differential pressure produced will be given by

= + (7.6)
^2 \ ^2 ■^l /

where Am is a part of the load on the second piston. The quantity in the large brackets in
equation (7.6) is zero because we have balanced the two pistons. It must remain zero or the
differential pressure will be affected. If it is zero, then the uncertainty in the differential
pressure will be a combination of the uncertainty in 242, g and Am. If we just take
= (7.7)
At
then the equation relating a small change ^AP) to small changes in Am and g and A2 is
^(AP) = f—K+f-^V(Am) + f^W (7.8)
V "^2 } < ^2 J k ^2 J

where in front of the term indicates a small change and the quantities in brackets are the
sensitivity coefficients as explained in NML Monograph 1, section 2.3 [5]. Let us take A2 as
80 X 10'^ m^ with an uncertainty of 8 x 10'^ m^. Am as 82 g with an uncertainty of 0.5 mg, and
g as 9.8 ms'2 with an uncertainty of .001 ms'^. The sensitivity coefficients are then, in the same
order as in equation (7.8), 1.03 x 10', 1.23 x 10'^ and 1.25 x 10^ so that the uncertainties are
1.03 Pa, 0.06 Pa and 1.0 Pa, respectively. The combined uncertainty is then 1.4 Pa. This is
considerably better than the calculated uncertainty in equation (7.4). This arises from the effect
of correlating the uncertainties in the two piston gauges so that they cancel out. It is crucial
that they do cancel out, which means, in practice, that the balance remains stable.

How well the balance can be maintained depends on the conditions of the laboratory as well as
the stability of the two piston gauges. To improve this stability the two piston gauges should be
identical. They should be maintained at the same constant temperature and to aid in this they
should be mounted in a common block. They should be operated in absolutely draught-free
conditions, by putting them in a box if necessary.

How can we take the stability of the balance into account since it does not appear in equation
(7.7)? It is necessary to go back to equation (7.6) and consider the uncertainty of the term in
brackets. The value of the term will be zero with an uncertainty attached to it. The value of this
uncertainty can be guessed at or experimentally it could be determined from studying the
balance over a period of time, perhaps with a differential pressure transducer connected
between the units. Let us take it as 10 Pa, which represents 1 ppm at 10 MPa. This now
dominates the uncertainty calculation so that the uncertainty becomes virtually 10 Pa. This
means that the lowest differential pressure that can be measured with 0.1% accuracy is 10 kPa

196
NOEL BIGNELL

(40 in WG). For lower line pressures and for larger diameter pistons the uncertainty can be
reduced.

Figure 7.1: The two identical piston gauges shown are accurately balanced when the
valve joining them is open. When it is closed a small known imbalance may
be produced by adding extra weights to one piston gauge. This differential
pressure can be used to calibrate a differential pressure (DP) cell or
transducer.

There are several commercial units on the market that operate on this principle. They have a
set of weights for the main balance and a set of smaller weights for the differential pressures.
The accuracy quoted is comparable with the example given here, about 7-10 Pa. Because two
piston gauges of superior quality and a larger than usual number of weights are required, they
are expensive.

7.4.3 NPL design

The National Physical Laboratory (NPL) in the UK has developed a differential pressure
standard using a differential piston arrangement [1]. Consider the small piston joined to the
large one shown in Figure 7.2. The effective area of the bottom of the large piston is Ai but
the effective area of the top. A?, is smaller because the action of the small piston is to subtract
its area from that of the large piston. With pressures P| and P2 acting as shown, and with m
being the mass of the piston and any weights on it we can equate the forces acting when the
unit is in equilibrium.
P,A, =P^A^+mg (7.9)
or
mg
(7.10)
AJ A
If there is a disturbance to the system so that (say) decreases, then the piston will move up
slightly hence increasing the volume in the lower compartment and. for a closed system,
reducing its pressure also until equilibrium is regained. The reverse occurs for increases in P^.
From equation (7.10) we can easily see that the change in AP, written ^ JP), is given by

197
CHAPTER 7. DIFFERENTIAL PRESSURE MEASUREMENT

8(AP) = 5P2( 1 - WA,)


assuming does not change. The value of (I-A2/A1) for the NPL device is 1/50. This means
that the changes, <5^2 in P2, are not reflected to the full extent in differential pressure AP as
they are for the previous designs. Instead they are reduced by the considerable factor of 50 so
that however stable we can keep P2 (with a piston gauge as before, for example) our
differential pressure will be 50 times more stable.

From equation (7.10):


(A,-AJ APA,
m = P,----------------------- (7.11)
g g

TO PISTON
GAUGE

PORI 1

Figure 7.2: A type of piston gauge designed at the National Physical Laboratory (UK)
for the accurate generation of differential pressures.

The value of m required for a particular P2 and AP could thus be calculated from a knowledge
of Al and A2 but if we put AP = 0 by simply connecting the top compartment in Figure 7.2 to
the bottom, then
P.(A, - A. )
Z72o = (7.12)
g

is the mass required for a balance. It is then only necessary to calculate the second term
APAt/g, and only necessary to know Ai, which is easily determined in a separate experiment.
Note that since AP can be positive or negative depending on which way round we think of the
two pressures, it is always only necessary to add weights.

The procedure then is to balance an external piston gauge, producing pressure P2, against the
differential unit with both parts connected together. Close the valve connecting the two parts
and add weight APAJg according to the AP required. Re-establish a balance by adjusting the
volume adjuster connected to port 1 and the differential pressure APA\lg at line pressure P2

198
NOEL BIGNELL

appears between the two ports. In the NPL device oil is used as the pressure medium so that if
gas pressures are needed oil/gas interfaces must be used and their levels carefully monitored.

7.4.4 Mercury columns

A mercury column can be used in a U-tube configuration with nearly equal high pressures
applied to the tops of the two arms. Normally a U-tube manometer is a simple device of glass
tubes and a scale. At high pressures, however, glass is far too unsafe and so some other means
of detecting the difference in height between the two columns must be found. This is done in
one commercially available instrument using two linear variable differential transformers. The
arms of the manometer are of large bore stainless steel tube, well polished internally, and each
has a differential transformer around the outside of it. The transformers are mounted on
movable platforms constrained to move vertically and driven by servo motors so that the
differential transformers are placed in a null position with respect to a magnetic body floating
on the mercury surface.

Having obtained, externally, the positions of the mercury surfaces inside the tubes, there are
several methods that could be used to determine the vertical distance between them. In this
commercial instrument moire fringes from steel scales are used to give an accuracy of 0.01%
with a maximum differential pressure of 400 kPa and a line pressure of 40 MPa. Below 40
kPa the uncertainty is 4 Pa [2]. Another instrument, described in the research literature, used a
cathetometer to sight the casing of the differential transformers and achieved an uncertainty 14
Pa at 20 MPa [3].

There are corrections that have to be applied to the result for the change in the density of
mercury with pressure and for the different column heights of the pressure fluid, be it gas or
liquid, above the mercury surfaces. The density of these fluids will change with pressure also.
In addition there are the usual corrections that apply to mercury manometers, namely those due
to the effect of temperature on both the density of the mercury and the length of the scale and
the difference of the local value of gravity from the standard value (9.80665 m/s^). See
Chapter 3 for details of these. They are made automatically in the commercial instrument.

7.5 Calibration of differential pressure transducers


It is quite easy to show the line pressure variation of the output for zero input, but, at least until
recently, it was a common practice to calibrate differential pressure transducers at atmospheric
pressure and to assume that these characteristics did not change at the high line pressure of
operation. This assumption has been shown to be false and errors of up to 1% at line pressures
of only 3.5 MPa have been observed at NPL (UK) in tests on commercial instruments.

Given a suitable standard of differential pressure the calibration of differential pressure


transducers at a particular line pressure follows along much the same lines as that of other
pressure transducers. This can then be repeated at other line pressures to cover the whole range
required. Frequently the transducer will be used at only one line pressure or over a narrow
range so that the enormous work involved in multiple calibrations from zero line pressure to
the maximum allowed can be avoided. The special precautions mentioned in section 7.3 should

199
CHAPTER 7. DIFFERENTIAL PRESSURE MEASUREMENT

be observed. The task of calibration of differential pressure transducers is one that needs
laboratory conditions if any of the standards discussed in section 7.4 are used.

It is possible, however, to use secondary standards that are more convenient for conditions
more likely to be found in the industrial or field situation. NPL have developed such a
secondary standard using a quartz bourdon tube transducer employing the force-feedback
principle (see Chapter 6, section 6.7.7).

7.6 References
[1] Dadson, R.S. Greig, R.G.P. (1965). J. Phys. E.: Sci Instrum. 42, 331.

[2] Delhomme, R.J. (1968). Product Specification KDG Instruments P/L, DSMIO Servo
Manometer.

[3] Lindsay, W.T., Bulischeck, T.S. (1970). Rev. Sci. Instrum. 41, 149.

[4] Patterson, J.B., Prowse D.P. (1983). CSIRO Appl Phys. Tech.

[5] R.E. Bentley. Uncertainty in Measurement'. The ISO guide. CSIRO, Sydney, Monogr. 1:
NML Technology Transfer Series, 2003.

200
Chapter 8

Pressure calibrators
Randall Anderson

8.1 Introduction
Pressure calibrators are a diverse group, varying in their designed use and technologies.
Discussion here is centred around transducer-based instruments designed for the calibration of
other pressure instruments - ‘pressure calibrators’. However, it also has relevance to digital
indicators used for direct measurements and to transducers used as field instruments.
The great variation in the design, performance and intended use of these instruments means
that many points discussed here may not have relevance in a particular case. It will be
necessary to adapt the principles to the particular instrument under consideration.

8.2 Functional design of pressure calibrators


When using a pressure calibrator, test pressures need to be generated and controlled as well as
measured. Often in field use, a hand pump is employed or perhaps a small gas cylinder with
manual control valves. It is important to ensure that a pressure point can be held with
sufficient stability to allow settling of each instrument and a comparison of readings.
Dead-weight testers have an inherent advantage in this respect, because the floating piston
performs exceptionally well as a controller of the system pressure, while providing a
measurement as well. Some top of the range calibrators have built-in pressure controllers that
aim to provide the same function. Some use an external pressure source and carefully control
the pressure using differential solenoid valves while others have an internal pump as well as
control functions. Stability of the pressure test point, with these controlling calibrators, can be
very impressive, but care is needed to ensure any instability does not impact adversely on the
measurement.
In considering the pros and cons in the use of calibrators versus dead-weight testers, it is
worthwhile noting that calibrators are not influenced by local gravity, thus avoiding the need to
correct for its influence in field use. Portability is often much better, with calibrators not

' From Australian Pressure Laboratory, Montmorency, Vic., invited to give an industrial
perspective.

201
CHAPTER 8. PRESSURE CALIBRATORS

requiring heavy stacks of weights, and a major advantage of the calibrator is its connectivity
with computer technology providing the possibility of data logging, automation of testing and
remote calibration. Many calibrators have electrical ranges for testing pressure transducers and
other electrical field instruments, a functionality that may be valuable for many users.
Calibrators on the other hand have less long-term stability and usually their temperature
sensitivity is larger and often unknown.

8.3 Desired outcomes of a calibration


• To ‘make it read right’. With the focus on quality of measurement and rigorous
uncertainty estimation we should not lose sight of the primary aim of the calibration,
which is to avoid problems due to inaccurate measurements.
• To provide the subsequent user of the instrument with sufficient information to enable the
best performance to be achieved. This requires complete and sufficient detail of how the
instrument is likely to perform in use. The best results may be achieved with information
from the user. This can influence how the instrument is tested.
• To provide traceability to national standards of pressure, the aim being that an instrument
should receive a comparable assessment regardless of which laboratory it is sent to.
• To alert the user to any aspects of the instrument’s performance that may impact on its
subsequent use.

8.4 Current standard test procedures


There is no current Australian Standard covering the testing of digital pressure calibrators and
indicators. Current practice is for each laboratory to develop a test procedure often based
loosely on the Australian pressure gauge standard AS 1349-1986 Bourdon tube pressure and
vacuum gauges. This can lead to considerable variation in the calibration treatment given to
these instruments.

8.5 Statements of compliance


A valuable adjunct to any calibration is an assessment and clear statement about whether the
instrument meets defined accuracy criteria. Without such a statement it can be easily assumed
by a user that an issued certificate of performance infers a satisfactory result when it may
actually prove the opposite.

Pressure calibrators have no classification systems detailed in a standard and, therefore, are
most often assessed against the performance criteria published by the manufacturer.
Assessment against an arbitrary performance can also lead to confusion in subsequent use. In
the case where a lesser standard of performance is to be accepted it is clearer to make
statements such as: ‘The instrument fails to meet the manufacturer’s accuracy specification but
meets the lesser requirement where corrections are less than 0.5% of range’. In order to verify
that the calibrator conforms to the accuracy specification, the laboratory requires reference

202
RANDALL ANDERSON

standards to be certified within these limits. End users will benefit from being informed of the
laboratory’s capabilities prior to testing.

8.6 A guide for testing a digital pressure calibrator


8.6.1 Test design

A test should be designed to operate the instrument with a repeatable set of test conditions in
such a manner that the measured performance is a good indication of the performance over the
next period. Practical and cost considerations will also play a part. For example, it may be
valuable to test a calibrator at several operating temperatures in order to predict its
performance in the field. However, since calibration time would be greatly increased, it is
common to test at one temperature and rely on the manufacturer’s specification regarding
temperature sensitivity.

Where possible, the test should emulate the conditions of use. Without user involvement it is
necessary to guess the likely conditions and methods of use. Where they affect the output,
they should be detailed in the test report. An example would be the warm-up time before
testing. It may not be appropriate to allow 4 hours warm up time at the time of calibration if it
is known that the unit is only allowed 10 minutes of warm-up time in use. On the other hand,
an extended period of warm up time may be appropriate if the test method aims to evaluate the
best performance of the instrument. A warm up time of 30 minutes is a reasonable compromise
to ensure stability is achieved with a realistic working protocol. If this is stated in the report the
user then knows this is required to achieve the performance indicated by the calibration report.

8.6.2 Test procedure

The following procedure is a general set of requirements that will need to be adapted to meet
the special requirements of a specific calibrator.
1. Test fluid
The fluid used to test the calibrator must be compatible with the instrument. Most calibrators
are marked with warnings if a particular fluid is required. Some, for example, require clean
dry gas due to the electronics in the transducer being unprotected or to avoid inaccuracies with
liquid use. Some calibrators have small internal tubing that can easily become blocked with
fluid contamination. If poor repeatability is observed, then clearing these tubes may be
necessary to obtain best performance.

Some calibrators are for use in an oil free system and will require testing on water or alcohol.
Many instruments can be calibrated on liquid or gas, but particularly for low pressure ranges a
test on gas may yield a much improved test uncertainty.
2. Stabilise the temperature of the calibrator
Sixty minutes is sufficient time to allow most instruments to come to thermal equilibrium near
room temperature. Extra allowance should be made if the instrument was initially at a very
different temperature (e.g., uncontrolled environments during transportation reaching

203
CHAPTER 8. PRESSURE CALIBRATORS

overnight temperatures below 5-10 °C, or daily temperatures well above 25-30 °C) or if it has
a very high thermal mass.
3. Check power supply and warm up
Instruments can be affected by low battery level or if a charger is in use. Before testing the
batteries should be recharged or the dry cells should be replaced. A test should also be
performed to determine if the output is affected by operating the unit with the charger
connected. The manufacturer's specified warm up time, if one is available, should be
determined. Without specific instruction, 30 minutes should provide adequate time. Some
instruments need to have the automatic “time off’ setting disabled to enable warm up and
testing. A shorter warm up may be called for if it is known the instrument can’t reasonably be
warmed up in use. Testing should not continue with display of a low battery signal.
4. Instrument setup and unit selection
It is possible to optimise some calibrators for testing. Select the highest possible resolution
and set any damping to a minimum. The resolution can sometimes be affected by the selection
of a particular pressure unit.

Select an SI unit for reporting the results of testing the device and try to avoid manometric
units, such as metres of water, to avoid confusion through conversion factors. Conversion
factors for any non-SI units should be stated in the report.

Instruments not based on a microprocessor technology may have selectable units in which to
report measurements and these scales may require separate calibration. In this case, a full test
should be performed on the scale graduated in the SI unit and at least a single point at full scale
should be included for each other unit.
5. Temperature during testing
It is important that the instrument is held at a constant temperature during testing as variation
can affect the output significantly.

The temperature sensitivity is a very important aspect of the calibrator’s performance. Often
for use in the field in a broad range of conditions, the performance at temperatures away from
20 °C can be critical. Some instruments are much better compensated for temperature than
others. The expected temperature sensitivity is usually specified by the manufacturer and this
should be used as a guide in the absence of it being measured. The assessment of temperature
sensitivity can be time consuming and expensive, however, current trends in user and technical
requirements suggest that it will in the future become a normal part of the assessment of a
pressure calibrator.
6. Head considerations
The head P/„ of fluid within the test rig is given by:

204
RANDALL ANDERSON

where pr is the test fluid density, pA is the air density, gravity g and h is the difference in
height between the reference level of the test instrument and that of the standard. The term is
often significant with low pressure instruments tested on liquid as well as high pressure tests
performed on gas.

For gases, the density of a compressible gas is dependent upon the pressure. A useful
approximation to the formula above for nitrogen at 20 °C, where h is in mm, becomes:
P/, =Pxl.118x10“’/z
It is preferable to arrange the two reference levels to be at the same height therefore alleviating
the need for the term to be calculated. It is good practise to calculate the term automatically
rather than having to determine whether the term is significant or not.
7. Initial spanning and zeroing of the instrument
Before testing, the instrument should be pressurized to full scale to exercise the transducer.
This phase of the testing may also form part of the assessment of repeatability. Spanning in
this way will often cause a change in the zero reading and the instrument should be re-zeroed
after providing sufficient settling time. Applying zero pressure may not be straight-forward
and it is recommended that a low range instrument being tested with a fluid be disconnected at
the instrument reference level to ensure there are no small head pressures on the inlet.

It may be necessary to zero the device frequently during testing and if this is found to be the
case, then it should be clearly indicated in the report. Zero drift can be observed by checking
full scale during testing against the value observed during initial spanning.
8. Evaluation of repeatability
Evaluating repeatability before further testing provides information on stability, which may
influence aspects of testing such as settling time and also gives early information on whether
the instrument is performing within specification.

The rigorous evaluation and analysis of repeatability is extensively discussed in the ISO
“Guide to the Expression of Uncertainty in Measurement". The method described here
provides an estimation process that is appropriate to instruments for which a satisfactory
evaluation has to performed in a time efficient manner.

The method of estimation requires the instrument to be pressurized to the required pressure
point 3 times and the readings recorded. The estimate for the standard uncertainty Ur may then
be taken as:
Ur = (Max reading - Min reading)/^/3.

The justification for taking a simplified approach here is that we are using a very small sample
where the confidence in the estimate is not great. If the instrument analysis warrants a better
estimate then a greater sample size and the usual formulation may be used.

Repeatability will usually vary over the range of the instrument. As a minimum, the estimate
should be determined for a low and a high value with additional points being necessary with

205
CHAPTER 8. PRESSURE CALIBRATORS

changes in resolution. For pressure points in between, repeatability can be interpolated from
the maximum and minimum points tested.

Considerable information about repeatability is also gained during the course of testing. The
degree to which the instrument repeats on falling pressure compared to rising and the
smoothness of the corrections through the range may give further information and these should
influence the repeatability estimate.
9. Points selection.
Setting a general rule for the selection of test points is made difficult due to the variety of
calibrator accuracy classes and design variations such as multiple sensors. A reasonable guide
would be at least 8 test points including zero. A lesser number of points can often be
satisfactory if the family of instruments is well known and the performance between points is
well established. Setting the number at 8 ensures a reasonable assessment of calibration curve,
and provides some redundancy, helping to identify incorrect readings.

Many calibrators have an accuracy specification expressed as a percentage of reading where


the accuracy tolerance is much tighter at lower pressures. In these instances, it is best to
include more test points at lower pressures and less near full scale to better confirm the
specification. Instruments with multiple sensors for a single range are also best treated in this
manner and care should be taken to ensure each sensor is tested at a minimum of 3 points.

Vacuum ranges: Many calibrators have a vacuum or compound capability. Points should be
selected in the vacuum range if the instrument is designed with this capability or a check made
with the end user for the vacuum range not to be tested. If the vacuum range is not tested this
should be made clear in the test report to ensure the user is informed.

Absolute ranges: Testing of a calibrator with an absolute pressure range requires a reference
standard with absolute pressure capability or a certified barometer. Some instruments have a
pressure port to connect to, while others may need to be put into a chamber that can be
pressurized while allowing readings to be taken.

Electrical ranges: Agreement from the user is required for testing of a calibrator’s electrical
ranges. The testing of electrical ranges is not, however, covered in these notes.

10. Testing ‘on nominaP and ‘off nominaP


Observation of a single test point can be performed in several ways. ‘Off nominal’ testing
means applying an arbitrary pressure near the test point and taking a reading. ‘On Nominal’
testing means the applied pressure is adjusted to be very close to the test point and a reading
taken A third method can be used in which the applied pressure is adjusted until the test point
value is read on the calibrator.

The uncertainty of measurement is greater with ‘off nominal’ testing, effectively doubling the
uncertainty due to resolution of the display. If we apply test pressures that are central to the
display increment then we might assume the uncertainty in the correction due to the resolution

206
RANDALL ANDERSON

to be the rectangular limits of +¥2 a count. If the applied pressure is not controlled to be central
to a display increment, then the limits of the rectangular distribution become ±1 count.

If the accuracy specification at a pressure point is more than 4 counts, or the repeatability is
more than 2 counts, then the choice of method becomes less important. However, if the
accuracy specification is perhaps only a couple of counts, or if the repeatability is within 1
count, the quality of assessment is greatly improved with ‘on nominal’ testing.
11. Readings with rising and falling pressures
Readings need to be taken with applied pressures rising from zero and falling from full scale
and both should be reported. Some calibrators exhibit considerable hysteresis, causing rising
and falling values to differ. Some test methods call for repeating the set of readings rising and
falling, presumably to help in the assessment of repeatability or to help prevent reading errors.
Repeating all points is not a rigorous assessment of repeatability as the sample of two readings
at each point is insufficient repetition. The observer may gain a better feel for the repeatability,
but the method by which the repeatability is formally calculated is unclear. From experience,
the calibration process is much better served if the repeatability is formally assessed and care is
taken with one set of observations.
12. Instruments with damped response
Many instruments have a designed damped response that assists with stabilising the display,
particularly in field use, where the system pressure may not be steady. Over-damped
instruments can present a problem for calibration and use. Some instruments require perhaps
30 seconds to be near enough to the final reading to be within specification. This makes
calibration very slow and raises the question about what settling time is reasonable. During use
particularly with a hand pump for pressure generation, the pressure may not be very stable and
it is questionable as to whether the instrument would ever reach its stable value. To this end it
is reasonable to set some limit to the settling time and increase the uncertainty allowance for
repeatability to cover any further settling of the output.

A reasonable limit for settling time is 30 seconds unless there is particular knowledge about
the instrument being given a longer or shorter period in use.
13. ‘Accuracy’ specification and adjustment prior to calibration
In the absence of a standard covering pressure instruments of this type, the most useful
specification of ‘accuracy’, or more-correctly, uncertainty, is the manufacturer's specified
tolerance. ISO/IEC 17025 does not insist on a statement of compliance to be made. However, a
report without one is more difficult to evaluate by the user.

Specification of accuracy can be very confusing. Some instruments are given a tolerance
indicating a constant uncertainty throughout their range, expressed usually as a percentage of
full scale. Some, however, have a % of reading statement with an absolute value to be added
to avoid the tolerance going to zero at zero pressure. It is recommended in these cases that the
tolerance specified for accuracy should be tabulated with the results to assist with
interpretation.

207
CHAPTER 8. PRESSURE CALIBRATORS

If adjustment prior to calibration would significantly improve the performance of the


instrument, then the adjustment should be undertaken whether or not the instrument is outside
the accuracy specification. This is in line with the client’s principal objective of calibration
being to ‘make it read right’. If adjustment has been undertaken, performance of the calibrator
before adjustment should also be included in the report in line with ISO/IEC 17025.

13. Estimation of uncertainty


This subject is covered in greater detail elsewhere in the Monograph, but the following notes
emphasise important components of the uncertainty relevant to digital calibrators.
The principal terms to include in an uncertainty analysis are:
Description: Notes:
Repeatability Covered above, Ua = (Max-Min)/A/3 repeatability test
Resolution Assume rectangular distribution half range = Vi Resolution
Ub = O.5Res/^/3. Doubled, if testing ‘off nominal’.
Head From measurement of reference height, gravity and density.
Height measurement usually contributes the principal element
Applied Pressure Depends on standard used, refer to certificate on standard
Drift since calibration of The calibration of the standard used is an on ‘on the day’ analysis
reference instrument. and consideration needs to be given to the expected drift since
calibration. The manufacturer may specify a drift expectation; a
history may provide some information. An estimate is better than
no allowance at all.

15. Information to be included in the report


All information required by the user to obtain the performance detailed in the report should be
included. Obviously, some common sense is required to prevent the document becoming a
user’s manual. However, if the instrument needs to be operated with a vertical orientation to
achieve the results expressed then it is imperative that this is mentioned in the test conditions.
Common test conditions to record are test medium used, warm up time, settling time,
instrument orientation, temperature of test, zeroing frequency and reference level chosen for
testing.

Some information is useful in avoiding confusion, such as inclusion of conversion factors for
non-SI units. Recording of internal calibration constants and the like can help the user ensure
the integrity of the instrument's use.

The table of results should include the applied pressures, the readings and corrections to apply
and the uncertainty calculated at each test point. Uncertainty will usually increase throughout
the pressure range.

208
RANDALL ANDERSON

Lists of standard detail to be included in test reports are detailed elsewhere and will not be
repeated here.
16. Removable modules
Many calibrators are multi functional with the one instrument able to measure a great range of
physical parameters. The use of removable modules is common. It is important to make a
distinction in the design of remote modules. Some require calibration data to be stored in the
base instrument and, therefore, the calibration is only valid when the module is used with a
particular base unit. This should be made clear on any certificate and the base unit clearly
identified. However, some modules hold all the calibration data and electronics in the module
itself and their output is entirely independent of the base unit. This means the modules can be
shared among a group of base instruments avoiding duplication of modules and calibrations.

209
CHAPTER 8. PRESSURE CALIBRATORS

210
Chapter 9

Measurement and uses of high pressures


Ed Morris and Dennis Tyrrell

9.1 Introduction
If one considers that high pressures begin at 1 atmosphere, then the field of activity covered in
this review is a very large one. Pressures in the range 0.1 to 10 MPa are quite common in
industry and even in everyday life, as are the Bourdon-tube gauges that are frequently used to
measure them. Pressures between 10 and 100 MPa are much less widespread and occur
mainly in the chemical industry and in hydraulic equipment such as rams, metal formers etc.
Above 100 MPa, however, the number of applications rapidly diminishes so that pressures of
1000 MPa and above are very much the province of the specialist.

The entire range above 0.1 MPa will be considered here, though topics covered elsewhere in
this course will be omitted or dealt with very briefly. Sections 9.2 to 9.5 concentrate on the
range up to 2000 MPa as within this area the techniques used at low pressures are still
basically applicable. Higher pressures are discussed in section 9.6 while sections 9.7 and 9.8
are devoted to brief reviews of safety issues and industrial applications.

9.2 High pressure standards


Although a very accurate and convenient standard, the mercury column is obviously limited to
relatively low pressures. Advances in the design, theory and fabrication techniques of piston
gauges have made them the obvious choice for pressure standards for pressures up to at least
2000 MPa. The difficulty presented by these instruments is the determination of the effective
area of the piston-cylinder assembly.

Below about 10 MPa the variation of the effective area with pressure, often referred to as the
distortion coefficient, is small and may be neglected, except for high precision laboratory
measurements. For instruments of a simple design it is possible to determine the distortion
coefficient theoretically. Provided certain conditions are met (e.g., the piston and cylinder are
constructed from the same material), the following simple formula results:

211
CHAPTER 9. MEASUREMENT AND USES OF HIGH PRESSURES

(9.1)

Where:
Ap = the effective area at pressure P
Ao = the effective area at 1 atmosphere
the Poisson’s Ratio for the material used
E the Young’s Modulus for the material used

Equation (9.1) gives a useful but not entirely satisfactory determination of the distortion factor
for standard instruments. At the National Physical Laboratory in England a direct measurement
of the distortion was made using what is known as the similarity method. This extremely
painstaking approach involves constructing two piston-cylinder assemblies that are
geometrically as nearly identical as possible. However, they are constructed from materials
with Young’s moduli that differ as much as possible. By comparing their behaviour a
measurement of the distortion factor can be obtained. Although the results of this work were
excellent, the controlled clearance piston gauge, which solves the distortion problem in another
way, has become common today.

Figure 9.1: A schematic illustration of a controlled clearance piston gauge

A controlled clearance piston gauge is illustrated schematically in Figure 9.1. The pressure to
be measured is introduced at A, while an independent adjustable pressure is present at B and in
the jacket surrounding the cylinder. The latter allows one to vary the piston-cylinder clearance.
For a given pressure at A, there will be a critical jacket pressure for which the clearance
between piston and cylinder is zero. Under these circumstances the effective area of the

212
ED MORRIS AND DENNIS TYRRELL

assembly becomes exactly equal to the area of the piston, which can be calculated with much
greater confidence than can the effective area of an ordinary piston-cylinder combination.

The gauge cannot, of course, be operated when the jacket pressure is equal to the critical
pressure. However the measurement that would be made at the critical pressure can be
accurately determined by extrapolation after making a series of measurements for which the
jacket pressure is less than the critical value.

The ability to control the clearance has the additional advantage that one can minimise the
leakage past the piston. This becomes very important above about 700 MPa and at 2000 MPa
such a facility is essential.

At pressures above 1000 MPa the difficulties involved in constructing and using piston gauges
increase enormously. At the same time, the uncertainty associated with the pressure
measurements increases, being less than 0.01% at low pressures and greater than 0.2% at
2000 MPa.

Figure 9.2: The controlled clearance piston gauge and associated equipment installed at
the National Measurement Laboratory

A controlled clearance piston gauge has been installed at the National Measurement
Laboratory and is shown in Figure 9.2. The console in the centre of the picture includes oil
reservoirs, air operated pumps (similar to the pump shown in Figure 9.4), pressure intensifiers
(similar to that shown in Figure 9.5), valves and interconnecting piping. It is effectively a dual
hydraulic system that can generate measured pressure and jacket pressure in independent

213
CHAPTER 9. MEASUREMENT AND USES OF HIGH PRESSURES

circuits. The structure on the left side of the console supports the piston and cylinder
assembly, the piston rotating mechanism and a 500 kg weight set. The weight set is raised and
lowered by an electrically driven screw jack to vary the weight, and hence, the force applied to
the piston. A data logging unit connected to a dedicated PC monitors jacket pressure, piston
fall rates and temperatures. With a suitable piston cylinder assembly this equipment has a
maximum pressure capability of 1.4 GPa.

Work is currently underway to characterise a 250 MPa piston cylinder assembly. In time, this
will lead to the establishment of the Australian pressure scale with this equipment recognised
as a primary pressure standard.

9.3 Fixed points


Most laboratories could afford to own and operate a piston gauge for measurements up to 200
to 300 MPa. For higher pressures however, the costs rapidly increase and the operation
requires highly skilled personnel. This problem is similar to one occurring in thermometry, and
has been overcome in the same way by establishing fixed points. The earliest such point to be
used was the pressure at which mercury melts at 0 °C. The pressure of this transition is now
known to be 756.95 ±0.15 MPa.

A fixed point for pressure measurement is a phase transition that, at a given temperature,
occurs at a certain fixed pressure. A suitable transition may, for instance, be the melting of a
solid, or a transition between two crystalline states in a solid. The pressures corresponding to
many fixed points have now been measured using the best available standards such as
controlled clearance piston gauges.

The melting of a solid is preferred as a fixed point because such transitions are rapid and
highly repeatable. Above 1200 MPa however, it is necessary to use transitions between two
solid phases. Various techniques are available for detecting the occurrence of a phase
transition under pressure. One can, for instance, measure the electrical resistance, volume or
refractive index of the sample, or measure the temperature changes resulting from the
absorption or liberation of latent heat.

The melting curve of mercury is now accurately known over the entire range 0 to 1200 MPa.
This curve can be thought of as not one but an infinite series of fixed points. Mercury held in a
pressure vessel at a temperature of T Kelvin will melt at a pressure given by

P = J9.32835d + 0.00J7068d^ + 6.087x 1d^ (9.2)

where:
d = T- 234.309, and P is in Megapascals.

If the temperature of the mercury can be controlled and measured with an uncertainty of
0.005 K, then the pressure at which the sample melts can be calculated from Equation (9.2)
with a total uncertainty of about 0.2 MPa. If a pressure uncertainty of ±1 MPa is acceptable.

214
ED MORRIS AND DENNIS TYRRELL

then a temperature uncertainty of ±0.05 K can be tolerated. The temperature range


corresponding to the 0 to 1200 MPa pressure range is -39 to +25 °C.

Fixed points used at higher pressures are discussed later.

9.4 Transducers
A wide variety of physical effects are dependent on pressure and many have been used with
varying success as the basis of pressure transducers. Only those frequently used at high
pressure will be considered here.

A distinction can be made between intrusive and non-intrusive transducers. Non-intrusive


transducers do not expose a sensitive measuring element to the high pressure fluid. Instead,
the strain induced by the pressure in a tube or diaphragm is generally measured by a device
that is entirely outside the high-pressure region. The resulting instrument is usually more
robust and more tolerant of dirty or corrosive pressure media. However the measurement of
strain limits the pressure range to about 1000 MPa, and limits the accuracy to about ±0.1%.

The Bourdon tube gauge is probably the most common and most accurate non-intrusive
instrument. Although normally used at low pressures, high-pressure gauges are available
commercially and an accuracy of 0.1 % of full scale deflection (FSD) is achievable. Although
normally available only to 700 MPa, gauges measuring to 1000 MPa have been developed.

A high-pressure variant of the Bourdon tube consists of a cylinder with an eccentric bore.
Gauges based on this device have an accuracy of only a few percent but are very robust.

In diaphragm and cylinder gauges, the elastic deformation of the diaphragm or cylinder is
measured with an electrical strain gauge bonded to the surface. The diaphragm types can be
used to 400 MPa, and the cylinders to 1400 MPa. An accuracy of 0.1% FSD is possible under
ideal conditions, but usually only 1 % FSD is obtained. Temperature effects are largely
overcome by comparing the resistance of the active strain gauge with that of a dummy which
is bonded to the same material and held at the same temperature, but not strained by the
pressure. In this way resistance changes due to temperature fluctuations can be cancelled out.
This principle of temperature compensation is often used in pressure gauges.

The bulk modulus cell measures volume strain by measuring the displacement of mercury
from a cylinder exposed to external pressure. The cell has an accuracy of about 1 % FSD.

Intrusive transducers are more accurate and have no intrinsic pressure limitations. On the other
hand, they are much more likely to be damaged by dirty or reactive pressure media and the
introduction of electrical leads through the wall of the vessel can sometimes be a problem.

Capacitance pressure gauges measure the change in capacitance that results from the change
with pressure of the dielectric constant of the material used in a capacitor. The change is
typically 3% for 1000 MPa. Since the temperature dependence of the dielectric constant is
usually quite large, care must be taken with the temperature compensation.

215
CHAPTER 9. MEASUREMENT AND USES OF HIGH PRESSURES

Resistance pressure gauges measure the change in the resistance of a wire that results from a
change in the pressure to which it is subjected. They are the most commonly used secondary
pressure gauge for accurate work up to about 3000 MPa, and have several notable advantages:
(1) They can be made fairly readily from quite simple materials
(2) For moderate accuracies, the sensing instruments required are cheap
(3) The transducer can be made small to reduce thermal effects.
(4) They are stable and reproduce well.
Silk-covered manganin wire is frequently used. Its resistance changes by 2.5% for 1000 MPa,
and at room temperature, the temperature dependence of resistance is small as it was originally
developed for use in precision resistors. Zeranin, gold-chromium, other alloys employed in
precision resistors, and even ordinary carbon resistors have been used successfully. Manganin
is probably the best choice if the gauge is to remain at room temperature, otherwise zeranin
may be better.

Figure 9.3: A manganin pressure gauge of the type used at the National Measurement
Laboratory. Both the main element (the longer one) and a reference element
consist of long loops of silk-covered manganin wire. The four wires emerging
at the top pass through a seal of araldite or a similar substance in the conical
section. They then connect in pairs to opposite ends of the manganin element.

The resistance elements are normally prepared with a resistance of about 100 Q. They must
then be seasoned to ensure stability. Manganin elements for instance, are seasoned by heating
them in oil to 140 °C for 24 hours. They are then subjected to a pressure higher than those at
which they are likely to be used.

216
ED MORRIS AND DENNIS TYRRELL

When mounting an element, it is important to ensure that the wire is not under any constraint,
otherwise strain in the wire will cause errors in the measurements. Thus, for instance, it would
be incorrect to wind the wire tightly on a bobbin. The manganin gauges made at the National
Measurement Laboratory are illustrated in Figure 9.3. It will be seen that the elements consist
of freely-hanging loops.

For the most accurate work, the element is connected as a four-terminal resistor and the
measurement is made with a high-precision bridge. Manganin gauges also need to be
temperature controlled to within about ±0.1 °C because of complications in the temperature
dependence of manganin resistivity as a function of pressure. An accuracy in the order of
0.03% of the measured pressure is then possible, and changes as small as 10 kPa can be
detected. For an accuracy of 0.2% the temperature control is unnecessary, and a digital multi­
meter of reasonable quality could be used.

A pressure change is inevitably accompanied by a change in the temperature of the


compressed fluid, roughly 1 °C for 100 MPa. This effect is important for electrical gauges in
particular, and even with the best temperature compensation there will be a delay before
accurate measurements can be made after a large pressure change. The delay will be
minimised by reducing the internal volume and hence the amount of fluid within the gauge.
For some bulky commercial gauges the delay can be 40 minutes or more, but for the gauge
shown in Figure 9.3 it is never more than 4 minutes.

9.5 High pressure generation and equipment


9.5.1 Pumps

At low pressure a distinction is made between gas pumps and liquid pumps, because gases are
far more compressible than liquids. Small volumes of liquid at pressures up to 200 MPa are
quickly and easily delivered from a small hand pump. Larger volumes may require a pump
driven by electricity or compressed air, the latter being the more usual. Figure 9.4 shows an
automatic reciprocating pump delivering liquid at pressures up to 200 MPa.

Two or even three stages of pumping are necessary to pressurise even small quantities of gas
to 200 MPa. The first stage could be provided by a conventional piston pump while the last
may consist of a pump similar to that shown in Figure 9.4.

Pressures above 200 MPa are best achieved with intensifiers, which consist of two piston­
cylinder assemblies of different cross-sectional area, see Figure 9.5. The larger piston drives
the smaller one, so that the pressure generated in the small cylinder is greater than that in the
large cylinder by a factor equal to about 90% of the area ratio. Friction is responsible for the
10% lost. Pumps such as those described in earlier paragraphs are required to drive the
intensifier and to provide the initial pressurisation (to at least 100 MPa) of the fluid in the
small cylinder. Intensifiers operating to 1400 MPa are available commercially.

217
CHAPTER 9. MEASUREMENT AND USES OF HIGH PRESSURES

Figure 9.4: An automatic reciprocating pump for liquids

Figure 9.5: An intensifier with a 16:1 area ratio

Most pumps require lubrication of moving parts and this can lead to contamination of the
pressure fluid. The contamination is usually fairly slight, but if it is to be entirely avoided then
compressors in which the piston is replaced by a pulsating diaphragm are used. These are
available at pressures up to 400 MPa.

218
ED MORRIS AND DENNIS TYRRELL

9.5.2 Tubing

Provided the flow rate is not very large, stainless steel (316) tubing is usually the most
convenient. Hard drawn stainless steel tubing with an outer diameter of 4.8 mm and an inner
diameter of 0.6 mm is used up to at least 1400 MPa.

Various types of tube connectors are available, particularly at lower pressures. At pressures
above 200 MPa, a simple cone in socket connection (socket angle 60°; cone angle 59°) works
quite well. See Figure 9.6, for example.

9.5.3 Pressure vessels

Users are often obliged to make their own pressure vessels or have them custom built,
particularly at the higher pressures. Some engineering experience and a comprehensive set of
stress/strain formulae are usually all that one needs to design small vessels. However, it is
advisable to be aware of the hazards peculiar to high-pressure work (see for instance the High
Pressure Safety Code given in the references) [7].

9.5.4 Seals

A wide variety of seals is described in the literature, and only a small selection will be
discussed here.

Figure 9.6 A simple cone-in-socket connection for


high pressure lines

The simple 0-ring seal is popular at low pressures, but is limited to about 100 MPa. At higher
pressures the O-ring tends to extrude from its groove, and should be used in conjunction with
back-up washers and anti-extrusion rings. Metal to metal seals, and seals using gaskets of
copper or bronze are used to high pressures. They are preferred only when the area enclosed
by the seal is small, and cannot of course be used in moving seals.

The most important high-pressure seal is undoubtedly the unsupported-area seal, which is
particularly useful as a moving seal such as on a piston. An example is shown in Figure 9.7.

219
CHAPTER 9. MEASUREMENT AND USES OF HIGH PRESSURES

The high pressure acting on the front face of the piston exerts an upward force on it. The only
downward force on the piston is that resulting from the pressure in the packing. Since these
forces must be equal,

P,A,=P,A, (9.3)
where:
Pf = the pressure in front of the piston
Pb = the pressure in the packing
Ap = the area of the front face of the piston
Ab = the area behind the piston on which the packing exerts a downward force

The area Ab is less than Ap by an amount equal to the area of the piston stem, which is not
supported either by the packing or by the piston collar. It follows then that Pb is greater than
Pp. It is this overpressure in the packing that very largely ensures that there is no leakage
through it from the high-pressure region.

9.5.5 Valves

Simple needle valves operating to 400 MPa are readily available. Valves operating at higher
pressures, and specialised types such as one-way, pressure-relief, or remote-control valves are
not so easily found, but are available commercially for pressures up to 1400 MPa.

For anyone wishing to design their own seals or valves there is extensive literature on the
subject. See for instance [1,4, 6].

Figure 9.7: An unsupported area seal on a piston.

220
ED MORRIS AND DENNIS TYRRELL

9.5.6 Pressure fluids

As the pressure increases, all fluids become more viscous and most will eventually freeze.
Care must thus be taken to select a fluid that will retain a relatively low viscosity throughout
the intended pressure range. Some hydraulic oils become sluggish at relatively low pressures,
but many may be used to at least 1400 MPa. For pressures above 1400 MPa petroleum spirit,
kerosene, and alcohols are preferred, as these media remain fluid to at least 2500 MPa.

9.6 Pressures greater than 2 GPa


Systems designed for use below 2 GPa usually consist of several units (valves, gauges, pumps,
etc.) interconnected by pipes. Above 2 GPa this convenient arrangement is no longer feasible
as it is not possible to construct reliable pipes, and in any case all fluids at room temperature
either freeze or become very viscous at about 3 GPa. The region exposed to the pressure must
therefore be highly localised; indeed it is restricted to the internal volume of the pressure
generator itself. This volume is seldom more than a few tens of millilitres, and at the highest
pressures may be less than one microlitre.

The need for the pressure sensor to lie within the working volume, which is itself within the
pressure generator, poses considerable problems. Another problem arises from the fact that,
with increasing pressure, true hydrostatic conditions become difficult and finally impossible to
achieve. Media that are fluid at low pressures become viscous and finally glassy hard. This
allows shear stresses to develop which can affect the pressure measurement and material
properties. Since there is no longer any advantage in using fluids they are usually dispensed
with, and most work is carried out in solid media such as soft metals, metal oxides or
pyrophyllite.

9.6.1 Standards

Piston gauges have been used to about 2.6 GPa but it is most unlikely that their range will be
extended. From 2.6 to 10 GPa the best available standard is the piston-cylinder press in which
a piston is forced into a cylinder by a hydraulic press. The cylinder in particular is of highly
specialised design and consists of several components. The pressure medium is a soft solid,
some of which extrudes into the gap between piston and cylinder thereby sealing the piston but
also creating large frictional forces. Since the pressure is obtained by dividing the force
generated by the press by the area of the piston-cylinder assembly, the friction is a major
source of error.

The friction can fortunately be reduced by partly rotating the piston, and what remains can be
largely removed from the calculations by making a measurement on falling pressure as well as
on rising pressure. The force measured on the upstroke is then averaged with that obtained on
the downstroke. In this way the pressures of several solid-to-solid phase transitions have been
established with an uncertainty in the order of 1%. These transitions now serve as fixed points.
Above 10 GPa there is no primary pressure standard. There are however two secondary
standards which appear to be quite accurate. They are the equation of state (i.e. density) of
sodium chloride, and the shift with pressure of the ruby fluorescence spectral line. Users of

221
CHAPTER 9. MEASUREMENT AND USES OF HIGH PRESSURES

these scales must introduce ruby chips or a small quantity of sodium chloride into their
pressure media. They must also be able to probe their pressurised media with either X-rays
(for NaCl) or laser light (for ruby) in order to make a pressure measurement. The sodium
chloride scale can be used to 30 GPa while ruby fluorescence has been used to pressures above
100 GPa. Below 10 GPa, both scales have been calibrated indirectly against primary standards.
Above 10 GPa they agree with each other to within 3%.

9.6.2 Fixed points

There are about ten well-documented fixed points covering the range from 1.2 to 30 GPa. The
lowest is a bismuth transition at 2.550 GPa and the highest a sodium chloride transition at
29.6 GPa. The bismuth transition pressure is known with an uncertainty of ±0.2% as it was
measured with a piston gauge, but some of the others have uncertainties of about ±3%. For a
comprehensive list of fixed points see [2].

9.6.3 Pressure generation

At pressure above 5 GPa the anvil is the dominant apparatus for pressure generation. It exists
in many forms, some of which are very complex. The simplest, known as the Bridgman anvil
in honour of its inventor, consists of two pistons that can be forced together in a press (see
Figure 9.8). The piston-cylinder press described in an earlier paragraph is not used very much
beyond 5 GPa, and beyond 10 GPa the anvil alone can provide stable and measurable
pressures.

The small flats on the anvils shown in Figure 9.8 are sealed by a cylindrical gasket that
encloses the pressure medium proper. After an excursion to high pressure the gasket will be
greatly extruded. With tungsten carbide anvils pressures in the order of 25 GPa are readily
achieved, and the more recently developed diamond anvils have been used to reach pressures
greater than 150 GPa.

The working volume of a Bridgman anvil is usually only a few microlitres, but belted or
girdled anvils can enclose working volumes of several millilitres. These devices combine some
of the features of both the Bridgman anvil and the piston-cylinder press. For the most part they
are restricted to pressures below 10 GPa.

9.6.4 Pressure measurement

Resistance gauges calibrated against fixed points are used extensively at the lower pressures,
and have even been used at pressures in the order of 20 GPa. Using fixed points it is also
possible to calibrate the relationship between the forces applied to an anvil and the resulting
pressure. Over most of the range however, the most accurate measurements are obtained by
using the sodium chloride or ruby scales directly.

222
ED MORRIS AND DENNIS TYRRELL

Gasket

-Workspace

Figure 9.8: A simplified illustration of the working area of a Bridgman anvil.

9.6.5 Dynamic pressures

Extremely high pressures (up to a few hundred GPa, at least) can be generated dynamically.
Such pressures occur for instance when a projectile strikes a target. The duration of the high
pressure is very short, being sometimes much less than one microsecond. The brevity has the
advantage that it is not necessary to build elaborate pressure containment devices. It has the
disadvantage that all measurements have to be made quickly. The temperature may also rise
very high, particularly if the target is a compressible material. Temperatures of 30,000 K are
quite possible.

Most work is carried out at the lower pressures (less than 1 GPa) where the duration of the
shock is much longer than one microsecond. An electro-elastic transducer such as a diaphragm
coupled to a strain gauge or to a capacitor can then be used for measuring the pressure. These
devices are re-usable and are available commercially. At pressures up to 30 GPa resistance
gauges are used and these are sometimes not re-usable.

9.7 Safety
This is a subject that is often neglected by the authors of books on high pressures. The best
starting point in the literature is undoubtedly the “High Pressure Safety Code” of the High
Pressure Technology Association of the U.K [7]. This is a short booklet which can be read in
a few hours and which covers a very wide range of hazards including many that are only
incidental to high pressure work. It also lists a dozen or so further references.

Since the above booklet provides an excellent introduction to the subject, what follows here is
just a brief review of the most important topics.

Even pressures less than 1 MPa can be dangerous under certain circumstances. Three
examples are

223
CHAPTER 9. MEASUREMENT AND USES OF HIGH PRESSURES

(1) When large volumes of compressed gas are involved.


(2) When the pressure fluid is flammable or toxic.
(3) If the apparatus and particularly tubing are not securely anchored.

Pressures greater than 5 GPa are usually not particularly dangerous as the volume of
compressed materials is small. The energy released by an explosion depends not only on the
pressure, but also on the volume by which the pressurised matter expands. For this reason
gases are inherently much more dangerous than liquids, as liquids do not compress to less than
50% of their original volume even at the highest pressures. Gases on the other hand can be
compressed to less than 0.2% of their volumes at one atmosphere.

There are three important hazards associated with high pressures:


(1) Projectiles
(2) Shock waves
(3) Fire or hot fluids

Shock waves are a hazard only in gas explosions, but projectiles are important for both gas and
liquid explosions. The third hazard is usually present only in special circumstances such as
when a flammable fluid is used. Formulae exist for estimating the strength of likely shock
waves, and for designing barricades to withstand possible projectiles. It is sometimes
necessary, particularly if dealing with gases above 1 GPa, to house the apparatus in a separate
room or bunker that must not be entered while the pressure exceeds a certain limit.

At pressures up to about 1 GPa, a high level of safety can be achieved simply by building large
safety factors into designs. At higher pressures this becomes increasingly difficult, and one
must be prepared for occasional failures and/or adopt very elaborate procedures to avoid them.

There are some special hazards peculiar to high pressures. Probably the most notable of these
is the fact that mercury and hydrogen under pressure can greatly weaken the steels used for
pressure vessels.

9.8 Industrial applications of high pressure


The vast majority of industrial applications involve pressures below 100 MPa. The large
capital costs of high-pressure equipment are strong incentives to use relatively low pressures
whenever possible.

In the chemical industry, pressure is used to promote reactions which would be far too slow or
which would not occur at all at atmospheric pressure. The pressure required is usually less
than 10 MPa, though quite a few processes operate in the range 10 to 100 MPa [1]. In heavy
industry, pressure is used to generate the large forces required. These are very large topics that
are covered by an extensive literature, so little more will be said about them here. Instead,
some applications that mostly involve pressures greater than 100 MPa are discussed briefly.

Spain and Paauwe [5] give an excellent review of industrial applications of high pressure.

224
ED MORRIS AND DENNIS TYRRELL

9.8.1 Chemical synthesis

There are a number of chemical processes that operate at or above 100 MPa. The synthesis of
ammonia by the Haber process dates from the turn of century and is still carried out at
100 MPa. Today however most ammonia is synthesised at pressures in the order of 20 MPa.
Low-density polythene, which is manufactured at pressures in the order of 350 MPa, is the
most significant high-pressure polymerisation process. However, other polymers such as
Teflon are sometimes manufactured at pressures at or above 100 MPa. The hydrogenation of
coal to form fuel oil is typically carried out at 50 MPa, but pressures approaching 100 MPa
have been used. Some specialised organic compounds are synthesised at pressures up to
1500 MPa.

9.8.2 Crystal growth

Crystals of substances such as ZnS, GaP and ferrites are grown from the molten material at
pressures in the range 10 to 20 MPa. Some of these crystals cannot be grown at atmospheric
pressure, and even when they can, better crystals usually result from the application of
pressure.

Pressures up to 400 MPa together with temperatures in the range 300 to 600 °C are required to
grow a wide variety of crystals from aqueous solution. The process is called hydrothermal
crystal growth, a term originally used by geologists to describe its natural occurrence. The
pressure is necessary as the substances involved (eg. SiOi and CaCO.O are not soluble at
atmospheric pressure. Pressure increases some aqueous solubilities by up to a factor of 2000.
The most commercially significant crystals grown in this manner are quartz and calcium
carbonate.

9.8.3 Production of superhard materials

Pressures of at least 5000 MPa along with temperatures up to 2000 K are required for the
production of superhard materials, which are used as abrasives in polishing powders, cutting
tools etc. The most well known are diamond and cubic boron nitride. Both are high-pressure
phases of substances that exist in other forms normally, and while diamond occurs naturally,
cubic boron nitride does not.

Von Platen and Bundy, each working independently, produced the first synthetic diamonds in
1953. By 1957 a commercial process had been developed by General Electric. For industrial
purposes the synthetic diamonds are superior to the natural ones, and presently account for half
the world consumption. Gem-quality diamonds have also been produced, but at a cost that
renders them non-commercial.

9.8.4 Hydrostatic extrusion

Extruded products such as wire, tubes and rods are normally made by forcing a billet through a
die by pushing it with a rod. An alternative that has so far been used only on a fairly small

225
CHAPTER 9. MEASUREMENT AND USES OF HIGH PRESSURES

scale is to force the billet through the die with hydrostatic pressure. The pressures required
range up to 1000 MPa.

9.8.5 Differential hydrostatic extrusion

Some materials such as tungsten, magnesium, molybdenum and beryllium are very brittle at
low pressures but become relatively ductile at high pressure. Such materials cannot normally
be extruded as they would crack. They can however be extruded if the entire process is carried
out at an elevated pressure [3]. This is achieved by having high hydrostatic pressure on both
sides of the die. The pressure on the output side needs to be at least 100 MPa.

9.8.6 Isostatic pressing

Many specialised metals or ceramics are made by melting or sintering powders. Sintering is
often preferred as melting can involve complex processes that are hard to control. Sometimes
the powders are pressed forcefully into moulds so that the article can be machined before being
sintered. An article that has been pressed but not yet sintered is said to be “green”.

The procedure, known as isostatic pressing, involves compressing powders within flexible
moulds using hydrostatic pressures up to 1000 MPa. This results in some important advantages
such as greater green strength and uniformity [3]. In a process called hot isostatic processing
the compaction and sintering are combined into a single operation.

Products of isostatic pressing include refractory parts, ceramic tool tips, turbine blades, filter
elements, spark plug insulators and ferrite parts. Hot isostatic pressing is specially suited to
forming parts and tools from super-alloys such as cemented carbides, high-speed steel and
high-temperature alloys.

9.8.7 Cutting materials with high-pressure jets

Both continuous and pulsed jets are used to cut a wide variety of materials. Pressures up to
1000 MPa are required. The procedure can speed up normal cutting operations, and is suited
to complex shapes and pliable materials such as plastics.

9.8.8 Explosive forming

The shock wave from an explosion forces a metal sheet into a die. The method is suited to the
forming of large or complex parts. The pressure in the shock wave can reach hundreds of
megapascals.

9.8.9 Explosive welding

Two plates are mounted at a small angle to each other and an explosive charge is detonated to
drive them together. The force of the impact welds the plates together. The pressures
generated can be extremely high, but the velocity of impact is generally a more important

226
ED MORRIS AND DENNIS TYRRELL

parameter than the pressure. The procedure is used for instance to bond a protective sheet to a
steel plate.

9.8.10 Autofrettage

Autofrettage is a process used on pressure containing components (high pressure cylinders,


gun barrels, high pressure tube, etc.) to generate favourably oriented stresses within the
component. These stresses increase the elastic strength of the component, its resistance to
fatigue and inhibits its crack propagation rate.

The pressure required to “autofrettage” a component will vary according to the size of the
component and the strength of the material.

In the case of a pressure cylinder, to achieve these favourable stresses, sufficient internal
pressure is applied to elastically and plastically deform the component in the direction of
normal loading. When the pressure is released, the cylinder is permanently deformed. The
outer portion of the cylinder has residual tension stresses that are in equilibrium with the
residual compression stresses at the bore of the cylinder. Machining to finished size after
autofrettaging is optional.

The Autofrettage process has been used in various forms dating back to the 19“’ century.
Equipment with a capacity to 1.4 GPa has been commercially available for many years.

227
HAPTER 9. MEASUREMENT AND USES OF HIGH PRESSURES

9.9 References
[1] Commings, E.W. (1956). High Pressure Technology, McGraw-Hill, New York,

[2] Peggs, G.N. (Ed.) (1983). High Pressure Measurement Techniques. Applied Science
Publishers, London.

[3] Pugh, H.L.I.DL (Ed.) (1970). The Mechanical Behaviour of Materials under Pressure.
Elsevier, London.

[4] Spain, I.L. and Paauwe, J. (1977). High Pressure Technology Vol. I, Marcel Dekker, New
York.

[5] Spain, I.L. and Paauwe, J. (1977). High Pressure Technology Vol. II, Marcel Dekker,
New York.

[6] Tsiklis, D.S. (1968). Handbook of Techniques in High-Pressure Research and


Engineering, Plenum Press, New York.

[7] High Pressure Safety Code, High Pressure Technology Association (U.K.) (1977).

228
Cover: aerial view of NML, Sydney

You might also like