You are on page 1of 65

CONTENTS

1. TRACEABILITY, WEIGHT INSPECTION AND WEIGHT HANDLING 1


1.1 TRACEABILITY 1
1.2 WEIGHTS 1
1.3 WEIGHT INSPECTION 1
1.3.1 Construction 1
1.3.2 Surface Condition 2
1.3.3 Magnetism 2
1.3.3a Compass 3
1.3.3b Gauss Meter 3
1.3.3c Davis Susceptometer 3
1.3.3d Permeability Meter 4
1.3.4 Surface Roughness 5
1.4 WEIGHT CLEANING 5
1.5 WEIGHT HANDLING 6
1.5.1 Gloves 6
1.5.2 Tweezers 6
1.5.3 Clean surfaces 7
1.6 WEIGHT STORAGE 7
1.7 CALIBRATION INTERVAL 7
2. BALANCE ASSESSMENT 8
2.1 INTRODUCTION 8
2.2 PREPARATION 8
2.2.1 Balance Location 8
2.2.2 Measurement Equipment 9
2.3 BALANCE TYPES 9
2.3.1 Two-pan balances 9
2.3.2 Single-pan mechanical balances 9
2.3.3 Single-pan electronic balances 9
2.4 ASSESSMENT 10
2.4.1 Two-pan balances 10
2.4.1a Visual Inspection and mechanical checks 10
2.4.1b Rider bar 10
2.4.1c Rider weight 10
2.4.1d Internal balance weights 10
2.4.1e Equality of arm length 10
2.4.1f Linearity of scale 10
2.4.1g Sensitivity 11
2.4.1h Repeatability of reading 11
2.4.1i Repeatability of measurement 11
2.4.2 Single-pan mechanical balances 11
2.4.2a Visual inspection and mechanical check. 12
2.4.2b Drift 12
2.4.2c Calibration of internal weights 12
2.4.2d Effect of off centre loading 12
2.4.2e Hysteresis 12
2.4.2f Scale sensitivity (scale value) 12
2.4.2g Scale linearity 12
2.4.2h Repeatability of reading 12
2.4.2i Repeatability of measurement 13
2.4.3 Single-pan electronic balances 13
2.4.3a Hysteresis 13
2.4.3b Effect of off centre loading 13
2.4.3c Scale error and linearity 13
2.4.3d Repeatability 17
2.5 FREQUENCY OF ASSESSMENT 17
2.6 END USER REQUIREMENTS FOR ASSESSMENT 19
2.6.1 General 19
2.6.2 What tests do we need to perform and when 19
2.6.3 The annual check 20
2.6.4 Daily or weekly checks 21
2.6.5 Automated checks 22
2.6.6 Reducing external influences 22
2.6.7 Keeping records 22
3. WEIGHING TECHNIQUES 23
3.1 INTRODUCTION 23
3.2 SINGLE PAN BALANCES 23
3.3 DIRECT READING MEASUREMENTS 23
3.3.1 Example 23
3.4 WEIGHING BY DIFFERENCES 24
3.5 SUBSTITUTION WEIGHING 25
3.5.1 ABA Calibration 25
3.5.2 ABBA Calibration 28
3.6 CYCLIC WEIGHING 29
3.7 WEIGHING BY SUB-DIVISION 33
3.8 MAKE-WEIGHTS 34
3.9 TWO PAN BALANCES 34
3.9.1 Double-Double Weighing 35
3.9.1a Use of Make-weights 35
3.10 DOUBLE-SUBSTITITION WEIGHING 35
3.11 BUOYANCY CORRECTIONS 36
3.12 GRAVITATIONAL CORRECTIONS 36
4. ESTABLISHING A MASS LABORATORY 38
4.1 CONSTRUCTION 38
4.1.1 Location 38
4.1.2 Construction 39
4.1.3 Services 40
4.1.4 Balance plinths and other furniture 41
4.2 EQUIPMENT 42
4.2.1 Balances and mass comparators 42
4.2.2 Weights 43
4.2.3 Environmental monitoring 44
4.2.4 Other equipment 45
4.2.5 Legal metrology 46
5. AIR DENSITY MEASUREMENT AND BUOYANCY CORRECTION 47
5.1 INTRODUCTION 47
5.2 DEFENITIONS 47
5.2.1 True Mass 47
5.2.2 Conventional Mass 47
5.2.3 Buoyancy correction 48
5.3 THE APPLICATION OF BUOYANCY CORRECTIONS 48
5.4 MEASUREMENT OF AIR DENSITY 49
5.4.1 Determination of air density from parametric measurements 49
5.4.2 Other methods for the measurement of air density 50
6. UNCERTAINTY IN MASS CALIBRATION 51
6.1 INTRODUCTION 51
6.2 THE MEASUREMENT 51
6.3 THE UNCERTAINTY BUDGET 51
6.3.1 Symbol 52
6.3.2 Sources of uncertainty 52
6.3.3 Values 53
6.3.4 Probability distributions 54
6.3.5 Divisor 55
6.3.6 Sensitivity coefficient (ci) 56
6.3.7 Standard uncertainty in units of measured, ui(W I) 56
6.3.8 Degrees of freedom vi 56
6.3.9 Adding it all up 57
6.4 AIR BUOYANCY UNCERTAINTY BUDGET 59
6.5 REPORTING THE RESULTS 61
6.6 FURTHER READING 61
TRACEABILITY, WEIGHT INSPECTION AND WEIGHT HANDLING
Ian Severn
National Physical Laboratory

1. TRACEABILITY, WEIGHT INSPECTION AND WEIGHT HANDLING

1.1 Traceability
Mass is unique amongst the base units of the international system of units in being
defined in terms of an artefact. A cylinder of platinum iridium alloy, held at the Bureau
International des Poids et Mesures near Paris, is defined as being exactly one kilogram
in mass. All mass measurements undertaken in the World should be traceable to this
artefact via National Measurement Institutes and accredited laboratories. The United
Kingdom’s national standard of mass, Kilogram 18, is held by the National Physical
Laboratory and is the basis of the entire mass scale in the UK. The NPL participates in
a wide range of international comparisons to ensure that measurements made in the UK
are equivalent to those made elsewhere. In the past there have been some problems
with organisations based in one country not accepting traceability to any NMI other than
their own (a prominent example of this is the American Federal Aviation Authority not
accepting measurements traceable to any source other than the American National
Institute of Standards and Technology - NIST), but this situation has now changed with
the advent of a structured approach to international comparisons that means that
measurements made throughout the world are accepted to be equivalent.

1.2 Weights
The use of appropriate mass standards and their correct treatment at all times is
essential to mass metrology. Weights are divided into classes from high quality
reference standards (Class E1) to those used in industrial settings (Class M3). These
classes were originally specified for legal metrology purposes, but they are now in
common usage throughout mass measurement. The International Organisation for
Legal Metrology (OIML) document OIML R111 specifies the properties of weights of
each class and specifies the tests that are necessary prior to carrying out a mass
calibration. A summary of the more important points is given here.

1.3 Weight Inspection


It is vital that weights are inspected prior to calibration to ensure the following:
a) they are of suitable construction for the type of calibration requested
b) they are in good condition
c) they are non-magnetic
d) they have a suitable surface finish

1.3.1 Construction
Weights should be made of a material that is chemically unreactive, non-magnetic, hard
enough to resist scratching and of a density that meets the OIML R111
recommendations for its class. Austenitic stainless steel is generally used in the

1
construction of Class E1 and E2 weights. Lower accuracy weights may be manufactured
from brass, iron or other suitable materials.

Class E1 and E2 weights must be integral in construction, ie be made of a single piece of


material. Other weights classes are allowed to be made up of multiple pieces with a
sealed cavity to allow for adjustment.

As with the other properties of weights the shape of weights for particular classes is
defined in OIML Recommendation R111.

1.3.2 Surface Condition


Weights should be inspected for surface damage such as scratching or contamination.
It is almost inevitable that weight surfaces will become slightly scratched in use, but
weights with gross scratching can potentially become unstable due to contamination
filling the scratch mark. It is recommended that weights are inspected on receipt at
calibration laboratories and any major scratches noted and reported to the customer.
Similarly it is good practice for an organisation that has sent a weight set away for
calibration to check it for any damage that may have occurred in transit from the
calibration laboratory.

It is common for weights to be submitted in a dirty condition - finger prints are a common
source of contamination. Dirty weights should be cleaned prior to calibration, but only
after the customer has given permission for this to take place. It is recommended that
weights should be calibrated both before and after cleaning.

1.3.3 Magnetism
There are two magnetic properties that must be measured to characterise weights,
permanent magnetisation and magnetic susceptibility. Magnetic susceptibility is a
measure of whether the weight can become magnetised by being placed in a magnetic
field (magnetism of this sort is transient) while permanent magnetism is a feature of a
weight that cannot be altered. It may be possible to de-Gauss magnetically susceptible
weights using a commercial de-Gausser, but such treatment has no effect on
permanent magnetisation. OIML Recommendation R111 sets out permissible limits for
these two properties for various classes of weights. Tables 1.1a and 1.1b shows these
limits.

Weight Class E1 E2 F1 F2 M1 M2 M3
Maximum Magnetization, µ0M (µT) 3 10 30 100 300 1000 3000
Table 1.1a: Maximum Permanent Magnetisation, µ0M (µT), for OIML Weights

Weight Class E1 E2 F1 F2
Nominal mass ³ 100 g 0.010 0.020 0.07 0.21
Nominal mass < 100 g 0.025 0.075 0.25 0.75
Nominal mass £ 1 g 0.12 0.37 1.2
Table 1.1b: Maximum Susceptibility, c

2
Tables 1.1a and 1.1b are taken from second draft of OIML R111, Feb 2000.

The susceptibility and magnetisation limits are specified such that weights with these
properties could cause an error in a balance reading of more than 10 % of the tolerance
of a particular weight class.

There are several methods available for testing the magnetic properties of weights
which range from the simple to the complex.

1.3.3a Compass
The most simple test is to bring a the weight under test into close proximity with a
compass. If the compass needle is deflected it indicates that the weight is magnetic.
However, this is a purely qualitative measurement and can only be used to indicate that
a weight is magnetic.

1.3.3b Gauss Meter


Gauss meters are commercially available from a number of manufacturers. They
measure the local magnetic field. The probe should be positioned such that the Earth’s
field and other ambient field are as near to zero as possible. The weight is then brought
as close as possible to the weight without contact being made. The permanent
magnetism should be taken to be twice the largest reading that is obtained at any point
around the weight’s surface. If carried out correctly the uncertainty in this measurement
is about 10 %.

1.3.3c Davis Susceptometer


This is maybe the most complex method of determining the magnetisation in a weight
and is generally only used in National Measurement Institutes. This method is able to
measure both permanent magnetisation and magnetic susceptibility. By comparison
with suitable magnetic susceptibility standards (at present these are only available from
the Magnetics Section at NPL) it is able to give quantitative measurements of the
magnetic field due to a weight. A schematic of the system is shown in Figure 1.1. This
technique based on a mass comparator of 2 µg resolution or better, although it would
probably be possible to adapt it for other balances of lower resolution if larger magnetic
fields were to be measured. This technique has the advantage of being able to map the
magnetic field in the weight under test by varying the height of the weight above the
magnet. The user is referred to Davis (1995) for further details of this method.

3
Figure 1.1: The Davis Susceptometer

1.3.3d Permeability Meter


This is a commercial instrument that has been on the market for some time. It is
illustrated in Figure 1.2. The magnet on the end of the pivoted lever is attracted to the
weight if it is more magnetically susceptible than the susceptibility standard on the
opposite side. The instrument is supplied with a range of susceptibility standards. When
testing a weight it is usual to start with the least powerful. If this does not attract the
lever the weight may be said to be less magnetically susceptible than that standard. If
the weight attracts the lever the strength of the magnetic standard is increased until it
attracts the pivot. The uncertainty in this measurement is ranges from 4 % to 30 %
depending on the magnitude of the weights susceptibility and suffers from problems in
obtaining traceable permeability standards. GREAT CARE MUST BE TAKEN WHEN
USING THIS INSTRUMENT. IT IS POSSIBLE TO INDUCE MAGNETISM IN
MAGNETICALLY SUSCEPTIBLE WEIGHTS THAT WERE NOT PREVIOUSLY
MAGNETIC.

4
Standard

Counter- Pivot
weight

Test-piece
Figure 1.2: Permeability Meter

1.3.4 Surface Roughness


The most common method for measuring surface roughness in mass calibration
laboratories is by visual inspection. In the case of dispute it may be necessary to make
a comparison with surface roughness standards. The hardware for more exotic
solutions involving detailed surface analysis are generally too expensive and time
consuming in their use to make them suitable for use in mass calibration laboratories.
The permissible surface roughness for various categories of weights is shown in
Table 1.2.

Class E1 E2 F1 F2
Rz (µm) 0.5 1 2 5
Ra (µm) 0.1 0.2 0.4 1
Table 1.2: Maximum permissible surface roughness

1.4 Weight Cleaning


Weights should routinely be dusted before use using a clean soft haired brush. With the
exception of this the cleaning of weights should be avoided unless it is absolutely
unavoidable. Permission must be obtained from the customer before cleaning weights
as it will affect their calibration history. It is recommended that weights should be
calibrated both before and after cleaning. It is good practice to write to a customer that
has submitted dirty weights suggesting how they should handle weights correctly.

The cleaning of weights should only be undertaken at the time of calibration as it


invalidates the previous calibration data. If a weight is found to be dirty the following
steps could be taken in order to clean it:

a) lightly dust with a clean soft haired brush


b) if (a) fails lightly rub the weight with a soft clean cloth (washed chamois leather is
ideal)
c) if steps (a) and (b) have failed, or are likely to further contaminate the weight (for
example if there is an oil patch on the weight that would be spread around by

5
rubbing), then solvent cleaning should be used as a last resort. In general this will
take the form of wiping the weight with a clean cloth that has been soaked in
solvent. Other more elaborate forms of cleaning include the use of ultrasonic baths
and soxhlet apparatus. After cleaning it is necessary to allow weights to stabilise
before calibrating them. The stabilisation time will vary according to the class of
weight and the extent of cleaning that has been undertaken.

1.5 Weight handling


Weights must always be handled with the greatest care. It is important that they are
never:

a) touched with bare hands


b) handled with sharp or abrasive tools and materials
c) in contact with tools or surfaces that are not scrupulously clean
d) slid across surfaces
e) knocked together
f) breathed on or spoken over by the operator
g) cleaned by inappropriate means

The following handling methods are recommended in order to avoid the problems
mentioned above.

1.5.1 Gloves
Gloves should be worn whenever practicable (sometimes they restrict the operator’s
ability to manipulate tweezers et cetera). Chamois leather is an ideal material for such
gloves as it has good thermal insulation properties (so reducing thermal influences on
the balance during the weighing process) and affords a good grip on large weights when
manipulating them. Wicket keeper’s inner gloves, available from large sports retailers
make ideal gloves for this purpose. It is important that the gloves are cleaned prior to
use using a mild detergent to remove the natural oils from the material.

1.5.2 Tweezers
Tweezers, forceps or other specialist lifting devices should be used whenever
practicable to pick up weights and manipulate them inside the balance. It is important
that these tools have no sharp edges and should, when possible, not have metal to
metal contact with the weights. The greatest problems arise at each end of the mass
scale. It is often necessary to handle fractional weights with metal tweezers due to their
small size. In this case the tweezers should preferably be made of copper or brass
which are relatively soft. It is necessary to manipulate large weights using mechanical
lifting equipment. This type of equipment usually takes the form of a manually operated
fork lift. In this case it is good practice to cover the forks with corrugated cardboard and
acid free tissue paper in order to reduce the potential to mark or contaminate the
weights.

6
When manipulating weights with tweezers, or similar tools, it is good practice to keep a
gloved hand beneath, but not in contact with, the weight being moved. This hopefully
minimises the potential for the weight to fall to the ground.

1.5.3 Clean surfaces


Surfaces in a mass laboratory should be kept clean and dust free at all times. However,
it is advisable to place acid free tissue paper, or something similar, on any surface prior
to putting weights on it. It is recommended that a minimum of three layers of acid free
paper are kept between the weights and any other surface (obviously this does not
apply to balance pans or storage boxes).

The use of paper to stand weights on also has the advantage of allowing the operator to
write the identity of each weight close to where it is standing during the calibration. Care
must be taken not to place weights on top of writing and the top sheet of tissue should
be disposed of at the end of the calibration.

1.6 Weight Storage


Weights should be stored in a manner that protects them from damage or
contamination and allows them to be easily identified. Manufacturers generally supply
weights in felt lined wooden boxes. Such boxes normally have a recess for each weight
and a lid which touches the weights when closed to hold then in place. Smaller weights
tend to be loosely confined within a small aperture in their box.

The major disadvantage with using boxes for storage is that they contact virtually the
entire surface of the weight so potentially causing contamination problems. It is
important that the interiors of weight boxes are kept clean to prevent a build up of dust
and other particles. A vacuum cleaner with a purpose built fine nozzle is ideal for this.

However, it is preferable for weights to be stored on acid free tissue paper under glass
domes. This is an ideal situation as only the base of the weight is in contact with other
surfaces. Regardless of how the weight is stored it is important to ensure that it is clearly
identified at all times so that it cannot get confused with similar weights.

1.7 Calibration Interval


There is no hard and fast rule relating to how frequently weights should be re-calibrated.
This depends on the uncertainty of measurement required, their frequency of use and
the environment in which they are used and stored. However, it would seem sensible to
not have a longer calibration interval than three years for typical weights. It must be
emphasised that in many applications much shorter calibration intervals are appropriate.

7
BALANCE ASSESSMENT
Stuart Davidson
National Physical Laboratory

2. BALANCE ASSESSMENT

2.1 Introduction

The need for regular and appropriate assessment of the equipment used in a calibration
laboratory is vital both in providing traceability of results and in providing accurate and
reliable results. This chapter covers the routine assessment of standard balances and
mass comparators used for the calibration of standard weights and for other laboratory
based mass measurement. It is not intended to cover the assessment of specialist
balances or to provide details of the test performed as part of the Type Approval of a
balance.

Balances can be used either as comparators or as direct reading devices. The


assessment of the balance will depend on the mode in which it is used. When used as a
comparator traceability for the calibration will be provided by the mass standard against
which the unknown is compared and the performance of the balance is effectively given
by its repeatability. When used as a direct reading device traceability for the
measurement is in effect provided by the balance itself and a number of other factors
such as the linearity, stability and temperature sensitivity of the balance become
important. In all cases it is vital that;

· the balance assessment reflects the way in which the balance is used in practice
· the assessment takes place in the same conditions as the balance would normally be
used under.

It is essential that both these conditions are met if the results of the balance assessment
are to give an accurate reflection of the performance of the balance and therefore
provide a valid contribution to an uncertainty budget for measurements performed on
the balance.

2.2 Preparation

2.2.1 Balance Location


To perform at its best the balance should be located in a position as free from vibration,
thermal instability, direct sunlight and draughts as possible. Care should also be taken
to avoid external influences such as magnetic fields and radio frequency interference.
The performance of the balance depends to a great extent on the environment in which
it is located so some thought should be given to optimising its position with respect to
the above parameters while allowing practical use of the balance. As previously stated
the balance should be assessed in the location in which it will be used. If a balance is

8
used in a shop floor environment it should not be assessed in a temperature controlled
laboratory environment.

2.2.2 Measurement Equipment


A basic assessment of the performance of a balance for repeatability, hysteresis and
eccentric loading can be performed with two un-calibrated weights of similar nominal
values. A more comprehensive assessment, however, requires a set of calibrated
weights and a temperature probe to monitor the temperature of the balance. Since the
sensing elements of most balances are temperature sensitive it is vital to keep a record
of the balance temperature while it is being assessed as unusual variations in
temperature may result in poor performance. A calibrated set of weights allow the scale
error of the balance to be assessed.

2.3 Balance Types

2.3.1 Two-pan balances


These balances consist of a symmetrical beam and three knife edges. The two terminal
knives support the pans and a central knife edge acts as a pivot about which the beam
swings. Two-pan balances are generally un-damped with a “rest point” being calculated
from a series of “turning points”. Some balances incorporate a damping mechanism
(usually mechanical or magnetic) to allow the direct reading of a “rest point”. Readings
from this type of balance tend to be made using a simple pointer and scale although
some use more complicated optical displays. In all cases the reading in terms of scale
units needs to be converted into a measured mass difference. Due to the construction
of the balance they are always used as comparators (comparing the load on one pan
with the other) but the way in which the comparison is achieved can vary - double-
double, double, double-substitution and substitution modes can all be used. It is vital
that the balance is assessed using the same mode of comparison as is used for the
normal operation of the balance.

2.3.2 Single-pan mechanical balances


These generally consist of a beam with two knife edges, one to support the weighing
pan and one acting as a pivot. A fixed counterweight balances the load on the pan. This
type of balance is usually damped and a series of built in weights allow the calibration of
a range of mass values by maintaining a constant load on the balance beam. The
displays on these balances tend to be of the optical variety and the sensitivity of the
balance can usually be adjusted by the user.

2.3.3 Single-pan electronic balances


These are usually top loading balances with the applied load being measured by an
electro-magnetic force compensation unit or a strain gauged load cell. Other sensing
elements exist and the way in which the balance load is applied to the sensing element
varies between manufacturers and balance capacities. Single pan electronic balances
give a direct reading of the weight applied whereas the other two mechanical balance

9
types rely on the comparison of two forces (an unknown weight with either an external
or internal weight). Despite the possibility of using these balances as direct reading
devices (applying an unknown weight and taking the balance reading as a measure of
its mass) single pan electronic balances will always perform better when used as
comparators, comparing a standard and an unknown in an ABA or ABBA sequence. As
with all the balance types discussed the method of assessment should reflect the way in
which the balance is used in practice.

2.4 Assessment

2.4.1 Two-pan balances


This type of balance is being used less and less frequently mainly due to the amount of
time it takes to make a weighing compared with electronic balances. For this reason
only a brief outline of the tests which can be performed on these balances will be given.

2.4.1a Visual Inspection and mechanical checks


With particular emphasis on the integrity of the knives and planes and checking that the
balance swings without fouling. The functioning of the arrestment the knife/plane
clearance and the poising of the beam should all be checked and adjusted if necessary.
The parallelism of the knife edges may also be checked (optically or mechanically) and
adjusted.

2.4.1b Rider bar


If the balance has a rider bar the length and notch spacing of the bar should be checked
either optically (for example with a travelling microscope) or using a special “heavy”
rider.

2.4.1c Rider weight


The effective value (mass value corrected for air buoyancy based on the actual density
of the weight) of the rider weight should be measured against suitable mass standards.

2.4.1d Internal balance weights


Instead of (occasionally as well as) a rider bar some two-pan balance have built in
poising weights. These should be removed and calibrated in the same way as the rider
weight

2.4.1e Equality of arm length


This can be checked by taking an unloaded zero point and comparing it with a zero
point with the balance loaded (calculated from a double weighing).

2.4.1f Linearity of scale


This can be checked using either fractional weights or with the rider and rider bar.

10
2.4.1g Sensitivity
Measuring the sensitivity of the balance allows a measured difference in terms of scale
divisions to be converted into a mass difference and is therefore vital to the use of the
balance. A suitable sensitivity weight, which can be easily transferred between the
balance pans without arresting the balance, is required. The value of this weight should
be enough to cause the pointer to move along the scale by between one quarter and
one third of the total scale length. The sensitivity weight should be calibrated against
suitable mass standards. The weight is then swapped between the balance pans (left to
right and then right to left) without arresting the balance. Arresting the balance will
change the effective rest point slightly and compromise the accuracy of the sensitivity
measurement. The average effect of swapping the sensitivity weight ( d ) is calculated
(the two figures should agree to better than 5%) and this value used to calculate the
sensitivity as follows:

2 ´ Ws
S=
d

Where: S is the measured sensitivity


Ws is the effective value of the sensitivity weight
d is the average scale difference

The sensitivity should be checked at at least four points across the weighing range of
the balance.

2.4.1h Repeatability of reading


The repeatability of reading can be checked by loading the balance and performing a
series of releases calculating a rest point for each series. Apart from arrestment the
balance remains undisturbed between releases making this measurement relatively
insensitive to the user. This makes this test useful in providing a temporal measurement
of the balance performance which is insensitive to external influences.

2.4.1i Repeatability of measurement


This provides a more accurate assessment of the balance’s performance “in use”. A
series of repeated weighing are performed using the balance in its normal comparison
mode (double weighing or substitution). The series of measured mass differences can
be statistically analysed to give a measure of the balance performance. The average
measured mass difference can also be compared with the certified mass difference
between the weights taking into account the uncertainties on these certified values.

2.4.2 Single-pan mechanical balances


In a similar way to two-pan balances, these single-pan mechanical balances are being
replaced by electronic balances which often offer better resolution and are easier to
use. Again, the range of test which can be performed on these balance will be only
briefly described.

11
2.4.2a Visual inspection and mechanical check.
Check knife edges and planes. Makes sure the released balance does not foul. Adjust
the zero of the balance.

2.4.2b Drift
This will give an indication of the balance’s stability and its sensitivity to changes in
environmental conditions, it will also give an indication of how long the balance should
be left to give a stable reading (the effectiveness of the damping). If the drift is linear it
can be eliminated by using suitable symmetrical weighing method (eg ABA weighing).

2.4.2c Calibration of internal weights


Ideally the weights built into the balance should be removed and calibrated externally. If
this is not possible they can be left in the balance and calibrated by dialling them up in
combinations.

2.4.2d Effect of off centre loading


The effect of off centre loading on the balance reading is assessed using a weight of
nominal value equal to about half the balance capacity (to avoid mechanical damage to
the balance). The weight is placed at the extremes of the pan and the results compared
with the reading when the weight is placed centrally. A significant difference between
the reading at the centre and extremes of the pan may indicate that the balance is
fouling.

2.4.2e Hysteresis
This is checked by taking readings of increasing and decreasing load. Significant
difference may indicate mechanical problems within the balance.

2.4.2f Scale sensitivity (scale value)


This is checked by placing a weight on the pan equal to the full range value of the
(optical) scale. For most single-pan mechanical balances adjustment of the scale
sensitivity is possible.

2.4.2g Scale linearity


The (optical) scale should be checked for linearity at regular intervals along its range
using a calibrated set of (fractional) weights.

2.4.2h Repeatability of reading


This is assessed by loading the balance (usually at maximum load) and taking a number
of readings by releasing and arresting the balance without otherwise disturbing the
balance. This provides a measure of the balance performance independent of external
variables and can thus be used as a repeatable check of the balance performance over
time.

12
2.4.2i Repeatability of measurement
This represents a series of actual comparisons using the balance in its normal weighing
mode. Statistical analysis of the repeatability of measurement provides a practical
assessment of the balance performance under normal weighing conditions.

2.4.3 Single-pan electronic balances


This type of balance is by far the most widely used and the simple principle of operation
means that a full assessment takes the form of relatively few tests. As with the other
types of balance the way in which the balance is assessed will depend on the way it is
to be used in practice. Before starting an assessment of an electronic balance it should
have been left switched on for a least an hour and preferably overnight (check the
manufacturers literature for details). Where appropriate the balance calibration facility
should be used before starting the assessment.

2.4.3a Hysteresis
Hysteresis is checked by taking readings for increasing and decreasing loads.
Significant deviation between the two readings indicates a poorly adjusted or dirty
balance.

2.4.3b Effect of off centre loading


This type of balance is generally more sensitive to off centre loading since there is
generally little de-coupling between the balance pan and the sensing element. The
effect of off centre loading on the balance reading is assessed by placing a weight, of
nominal value equal to about half the capacity of the balance, at the extremes of the
pan and comparing the results with the reading when the weight is placed centrally.
Typical results for such a test are shown below.

Reading Centre Right Front Left Back


sequence
C > R > F > L > B 0.86 g 0.81 g 0.83 g 0.87 g 0.82 g
B > L > F > R > C 0.85 g 0.87 g 0.80 g 0.86 g 0.83 g
Mean reading 0.855 g 0.84 g 0.815 g 0.865 g 0.825 g
Reading relative to centre -0.015 g -0.04 g +0.01 g -0.03 g

The weight is placed at each position on the pan twice and the weighing scheme is
symmetrical with time so any balance drift is calculated out.

2.4.3c Scale error and linearity


The scale error and linearity of the balance can be checked with a suitable set of
calibrated mass standards. Measurements should be made at about ten equal steps
across the range of the balance. Further checks of the scale accuracy can be made
across sections of the balance range which may be of particular interest. For example, if
the balance is usually used at particular load the linearity of the scale around this load
should be measured. Similarly if a part of the scale is shown to be particularly non-linear
from the initial test this part of the scale should be checked across a narrower range.
Two typical results for scale linearity tests are shown below. The results show zero

13
readings taken before and after the test. In practice zero readings should be taken
between each scale point and the balance readings corrected accordingly (this is
particularly important if the balance shows significant drift in use).

Nominal Applied Balance Scale


load load reading error
(g) (g) (g) (g)

Zero 0.00000

10 10.000001 10.00015 0.000149


20 20.000003 20.00037 0.000267
30 30.000004 30.00041 0.000406
40 40.000001 40.00060 0.000599
50 50.000011 50.00063 0.000619
60 60.000012 60.00052 0.000508
70 70.000014 70.00041 0.000396
80 80.000015 80.00022 0.000205
90 90.000013 90.00013 0.000117
100 100.000009 100.00005 0.000041

Zero 0.00000

Table 1: Scale linearity for an electronic balance with single point calibration

0.0007
0.0006
Scale error (g)

0.0005
0.0004
0.0003
0.0002
0.0001
0
0 20 40 60 80 100

Balance reading (g)

Graph 1: Scale error of electronic balance across its weighing range

14
These results show the typical performance of a balance with a single internal
calibration weight at full capacity (100 grams in this case). The maximum scale deviation
occurs at 50 grams half way between the two fixed points (zero and calibration at 100
grams). The small scale error at 100 grams may be due to an error in the internal
calibration weight.

If the balance is to be used as a direct reading device allowance should be made for the
scale errors measured either by making corrections to the balance readings or by
allowing for the (maximum) scale error in the uncertainty budget.

When using the balance as a comparator the absolute scale error is not significant since
it applies equally to both weights. The difference in the slope between the scale error
line and the horizontal is however significant. Graph 1 shows that between 40 and 50
grams the error is smallest since the slope of the line is near horizontal (interestingly,
the maximum scale error occurs at 50 grams). The maximum deviation from horizontal
occurs between 70 and 80 grams which is where the most significant errors will occur
when using the balance as a comparator. In practice the scale error between these two
points is equal to the difference in the two measured scale errors ie;

0.396 - 0.205 mg = 0.191 mg over the range 70 to 80 grams

Thus comparing a 70 gram weight with an 80 gram weight would introduce an error of
0.191 grams. On a more realistic level, calibrating a 0.7 Newton weight (nominal mass »
71.36 grams) against a 70 gram mass standard would introduce an error of;

. ´ (7136
0191 . - 70)
mg
10

which is equivalent to 0.026 mg. It can be seen that such an error may be significant and
should be taken into account when producing an uncertainty budget for work done on the
balance in question. Larger errors may be introduced when comparing components which
are further from nominal (for example comparing a piston weighing 75 grams with a mass
standard of either 70 or 80 grams will introduce an error of 0.091 mg).

The second set of data shows results for a balance which is calibrated at two points in
its range (half and full load).

15
Nominal Applied Balance Scale
load load reading error
(g) (g) (g) (g)

Zero 0.00000

10 10.000001 10.00015 0.000149


20 20.000003 20.00037 0.000367
30 30.000004 30.00041 0.000406
40 40.000001 40.00013 0.000129
50 50.000011 50.00001 -0.000001
60 60.000012 59.99991 -0.000102
70 70.000014 69.99971 -0.000304
80 80.000015 79.99965 -0.000365
90 90.000013 89.99985 -0.000163
100 100.000009 100.00000 -0.000009

Zero 0.00000

Table 2: Scale linearity for an electronic balance with two point calibration

0.0005
0.0004
0.0003
Scale error (g)

0.0002
0.0001
0
-0.0001 0 20 40 60 80 100

-0.0002
-0.0003
-0.0004
Balance reading (g)

Graph 2: Scale error of electronic balance across its weighing range

It can be seen a two point calibration has the benefit of reducing the maximum scale
error (in general the maximum error should be about one quarter of that achieved with a
single point calibration). This is useful when the balance is used as a direct reading

16
device. When used as a comparator the deviation in the slope of the scale error line
compared with the linearity line will introduce errors into the comparison. The graph
shows a maximum deviation between 30 and 40 grams. Taking the example of a 0.3 N
weight compared with 30 gram mass standard the error introduced is;

(0.599 - 0.406) ´ (30.58 - 30)


mg
10

Which gives an error of 0.011 mg. It can be seen that this error is of equivalent magnitude
to that obtained for the 0.3 N weight calibrated on the previous balance (0.37 ppm in both
cases) and thus significant in terms of an error budget. The results show that while a two
point calibration will generally improve a balance’s performance as a direct reading device
it will not make it any more accurate as a comparator. Generally, for this type of balance,
the maximum deviation of slope will occur between 40 and 60 grams where the scale
error graph crosses the X-axis.

Given the above scale error measurements illustrated, more detailed assessment of the
ranges 70 to 80 grams for the first balance and 30 to 40 grams for the second balance
may be undertaken. The results of these extra tests will help to quantify the maximum
potential error due slope error for inclusion in an overall uncertainty budget.

2.4.3d Repeatability
The repeatability of an electronic top pan balance can be assessed by repeated
application of a mass standard. In general ten measurements will give enough data to
analyse. The repeatability should be measured at at least two loads (usually half and full
load) and at any other load where the balance will be normally used. Generally the
following data should be calculated for a repeatability assessment:

· Maximum difference between consecutive weighings


· Maximum difference between any weighings
· Standard deviation of the series (sn-1)
· Standard deviation of the series accounting for drift

2.5 Frequency of assessment


A new balance should be fully assessed before it is used for any calibration or other
measurement. This allows a comparison of the measured performance with the
manufacturer’s figures. It also provides data against which subsequent assessments can
be compared and input data acquired for uncertainty budgets. It should be noted that
manufacturer’s figures for balance performance are acquired under near ideal conditions
(often loading the balance automatically) and figures obtained under non laboratory
conditions may exceed the specifications. A maximum period for the re-assessment of a
new balance should be 1 year. Eccentricity, scale error and linearity and repeatability
should be re-assessed and compared with the original figures. The subsequent re-
assessment period will depend on balance usage but should generally be 1 year or less.
If, after a number of assessments, the performance can be shown not to change

17
significantly the re-assessment period may be extended. In any case a balance should be
re-assessed if it is moved, modified or changed in any other way.

18
END USER REQUIREMENTS FOR ASSESSMENT

M J Buckley
Standards Officer and General Manager
South Yorkshire Trading Standards Unit

2.6 End user requirements for assessment

This section deals with the practical information that end users should obtain when
assessing balances and follows on from the assessment of balances.

2.6.1 General
When assessing balances we need to accumulate information that enables us:

· To establish that the calibration system meets the defined requirements


outlined in our quality management system and the uncertainty budgets
that we have established using the weighing equipment in the normal
location of use

· To monitor our balances to see if their performance is being maintained, or


is indicative that attention may be required now or in the future

· To monitor performance, taking into account not only the performance of


the balance at that moment in time, but also the environment surrounding
the balance at the time these measurements are made, and the
performance of a particular balance user.

Accordingly, an annual calibration is just not sufficient. We might perform an annual


calibration to ensure we cover all requirements at least once a year, but we need to
perform other checks daily or weekly, depending on the performance of the balance with
time and its contribution to the overall uncertainty of measurement when calibrating a
weight.

2.6.2 What tests do we need to perform and when


We cannot avoid the annual check. If you operate an accredited facility you will need to
not only carry out these tests but also keep records of the tests. An annual check also
enables us to carry out some tests that we might not do during daily or weekly checks.

It is necessary to carry out the routine tests referred to in Para 6.2: linearity, eccentricity
and repeatability are essential. In addition we need to check on other things such as the
effect of electrical disturbances or local magnetism. We may also have to assess local
effects such as long term vibration or temperature problems.

The results of these tests may not however be truly indicative of long term performance.
Short term repeatability problems may ‘colour’ the results of our measurements. It is

19
therefore necessary that a log is kept of all balance calibrations. We can also calculate
a long term process standard deviation for each balance.

So we must also carry out short term testing. We may need to carry out some form of
daily or weekly testing. We also need to monitor the short term repeatability of each
weight calibration. Different users may produce different repeatability values, perhaps
because of slight variations in operating techniques, or due to heavy breathing (more
about this later!).

2.6.3 The annual check


The annual check provides us with a lot of information. But before we start our formal
balance test, we need to carry out some start-up checks:

· Check that the balance is complete, clean and undamaged. If the balance
is electronic, has it been connected to the mains supply for sufficient time
(the ideal is to leave it permanently on, but in standby mode, if there is
such a facility, when the balance is not in use)

· Remember that dust or other material may get trapped below the balance
pan. Give it a clean

· Check that ancillary equipment works correctly. If the doors are motorized,
do they function correctly, opening and closing without banging. What
about auto-zero and stability settings? Have they been set for optimum
performance where the balance is located, and have the settings been
altered

· Check for any localized vibration or magnetism that was not previously
present

· Have there been any changes in the local environment (temperature,


pressure, the effects of air conditioning etc).

Now we can carry out an assessment of the balance. Many balances require some
preliminary weighings to condition the balance. Some larger capacity balances work
best if they are preloaded when not in use.

We should first zero the balance then, if this is the normal calibration practice, check
(and set) the calibration interval using the internal or external calibration cycle. On large
capacity balances the internal weight may be proportional to the displayed value – with
the result that any minor error in the value of the internal weight produces a large error
in the displayed value. In this case use an external weight. Remember that the
calibration cycle only works if after loading the internal or external weight, zero is
regained on removing the weight. If the weighing machine has an in-built calibration
cycle, check that it has worked by putting a similar value weight on the load receptor
(with the machine out of the calibration cycle).

20
Now we need to carry out a linearity check. Sometimes this need only be for part or all
of the electronic range. Other balances, particularly used for direct reading, may require
the whole range calibrating, together with the value of any range or dialled-on weights.
At the conclusion of the test recheck zero – to assess hysteresis (if this is a problem a
full linearity test with descending loads may be more appropriate). If you rely on part of
the electronic range to define the difference between the standard and test weight you
will need to assess the uncertainty arising from this procedure and include it in your
uncertainty budget. If you use mechanical balances you will need to perform a linearity
test in a similar way, calibrating the optical or mechanical scale as well as any in-built
weights.

Next we need to perform an eccentricity test – that is to establish if there is any variation
in the indicated value if we place a weight in the middle of the load receptor or in a
position which is off-centre. Where possible, however, we should try and design the
effects of eccentric loading out of the weighing machine, with auxiliary hanging
suspension, or by the use of specially designed load receptors. It is necessary that we
check that the eccentric pan is actually functioning as designed.

Now we need to assess the repeatability of weighing the same weight a number of
times. This test should be conducted using the same technique as when the balance is
used for normal calibration. Even if you do not normally carry out as many as ten
difference weighings when calibrating weights, it is convenient to assess balance
performance by weighing the standard and test weight ten times, using the aba or abba
technique normally used in calibration. When making these repetitive measurements we
may get different standard deviations if we use different operators to perform this
operation, or if the environment is different (using solid rather than adjustable weights
when the air pressure is changing illustrates one example of possible variation in
repeatability). Furthermore, placing the weights used in this test in a different position
from normal may result in different repeatability values – some parts of a balance are
hotter than other parts. Finally, local effects such as the effects of the operator breathing
may influence the local environment and so cause changes in repeatability.

2.6.4 Daily or weekly checks


We have already stated that an annual check is not sufficient. We need to carry out
daily or weekly checks as well to ensure that the balance remains in a satisfactory
operating state. As a minimum we may need to carry out the internal or external
calibration cycle every day or several times a day and then check its effectiveness using
a single or a small number of calibrated weights. Note that if we activate the calibration
cycle as part of the annual calibration we should then carry out the same procedure
each time we use the balance – adopting this technique every time does not invalidate
the annual calibration.

If more than a single measurement is made when calibrating weights (and a single
measurement is not in itself a very safe procedure) then we can evaluate the
measurement series and monitor the standard deviation of this measurement against a

21
process standard deviation. Regular monitoring of performance will provide information
as to the condition of the balance (and ensure by maintenance that a balance does not
suddenly become unusable).

2.6.5 Automated checks


Many modern balances have in-built weights. It is possible to set up automated checks
each day (perhaps at night) using the in-built weights, which are loaded and unloaded
automatically perhaps ten times, after some initial pre-weighings.
Although this will not give the same repeatability value as when a human being uses the
balance to perform calibrations, it enables us to again assess the long term
performance of a balance. We can also use the information obtained to decide whether
a balance requires maintenance.

2.6.6 Reducing external influences


Whether we are calibrating a balance or using it for routine calibration it is advisable to
reduce the effects of external influences. Proper mounting of the balance should avoid
the problems of vibration, but local thermal or pressure effects may occur due to
changes in the local environment (heating going off at night or solar heating on
windows), heat from observers or lighting, or air conditioning or opening and closing of
doors (which can cause pressure effects).

2.6.7 Keeping records


It is important to keep records of all the checks we make. These include the annual
calibrations, as well as the daily and in-process checks. If possible these records should
be made using a computer recording system to automatically record the balance
readings. This avoids the possibility of errors arising from the manual recording of data.
The software used however must be thoroughly checked for errors arising from unstable
readings, poor mathematics and even interface errors. What is displayed on the balance
should be what is recorded in the software – in many cases it is not!

22
WEIGHING TECHNIQUES
Ian Severn
National Physical Laboratory

3. WEIGHING TECHNIQUES

3.1 Introduction
There are many weighing techniques currently employed in mass metrology. This is
indicative of the wide range of processes that rely on weighing and the uncertainty
demands that are required by different industrial sectors and end-users.

3.2 Single Pan Balances


The operation of a single pan balance is outlined in Section 2. This type of balance
gives an output, normally in the form of a numerical display, that indicates the apparent
mass of the object under test. For low accuracy applications this output may be used as
a direct measurement of the artefact being measured, but for more demanding
measurements it is usual to make comparative weighings.

3.3 Direct Reading Measurements


The most simple method of weighing is to simply place a test piece on a mass balance
and take the displayed reading as its weight. This type of measurement is only suitable
for low accuracy applications but even in this most straight-forward application it is
essential to follow good practice.

As with any form of mass calibration it is essential to have the balance calibrated on a
regular basis. It is recommended that the balance undergoes a full assessment and
calibration, by a suitably accredited body, on a periodic basis which will be influenced by
the application and frequency of use. It is important that the balance scale is tared
before use. If there is an internal balance calibration feature this should be used prior to
making measurements on the balance. A reading with zero load should be taken (z1)
followed by the reading with the load on the pan (r1) and a final zero reading (z2). This
allows the user to compensate for any drift in the instrument. The drift corrected reading
(rd) is given by:

z1 + z 2
rd = r1 -
2

3.3.1 Example
An example of how a table of balance reading corrections may look for a 500 g capacity
instrument is shown in Table 3.1. It would be usual for only one of the fourth or fifth
column to appear in such a table of results, depending on the preference of the
calibration laboratory so the user must be aware of both possibilities. The column
marked correction indicates the correction a user must apply to the balance reading to

23
take account of the errors in the scale linearity while the error indicates the offset from
the correct value that was observed by the calibration laboratory. The calibration
correction has the same magnitude as the correction but has the opposite sign.

Nominal load Value of Balance Error in Correction


standard reading reading
(g) (g) (g) (g) (g)
100 100.000 0 100.003 +0.003 -0.003
200 199.999 7 200.006 +0.006 -0.006
300 300.000 3 300.005 +0.005 -0.005
400 400.000 4 400.003 +0.003 -0.003
500 499.999 7 500.002 +0.002 -0.002
Table 3.1: Balance Reading Calibration Table

If when an object is weighed on this balance the initial and final zero readings are 0.000
and 0.006 g respectively, while loaded the reading is found to be 303.235 g, the actual
corrected value of this object would be 303.230 g - (0.000 + 0.006)/2 g = 303.227 g. If a
reading is between two calibration points it is usual to apply the correction of the point
closest to the indicated reading. In the case of a reading of 242.342 g with zero readings
of 0.000 and 0.002 g the calibration correction to be applied would be -0.006 g giving a
drift corrected result of 242.336 - (0.000 + 0.002)/2 g = 242.335 g.

3.4 Weighing by differences


This technique is particularly common in analytical chemistry. In general It involves
placing a container on the balance pan, noting the reading, and then adding a
substance it. The final balance reading is noted and the difference between the two
readings is taken to be the amount of material in the container. This approach is fine for
relatively low accuracy requirements but is not ideal for more demanding
measurements. This type of measurement is susceptible to problems with temporal
balance drift (which is generally more of a problem with modern electronic analytical
balances than was the case with mechanical balances).

A more robust method of carrying out measurements of this type is to have two similar
containers, one to fill with the test substance and one to act as a reference. The
weighing should then be carried out as follows:

1. The reference container is placed on the balance pan and the reading noted (ref1).
2. The empty container (to be loaded with the test substance) is placed on the balance
(test1)
3. The container is filled with the test substance and the balance reading taken (test2)
4. The reference container is put back on the pan and the reading taken (ref2).

The weight of substance added to the container (wt) may then be calculated as follows:

24
w t = ( test 2 - test1 ) - (ref2 - ref1 )

This weighing scheme eliminates drift and should allow better weighing uncertainties to
be achieved.

3.5 Substitution Weighing


High accuracy weighing is generally carried out by comparing the test-weight with mass
standards of similar nominal value. This is known as substitution weighing.

When weight A (the first weight to be place on the balance pan) is compared with weight
B (the second weight used) there will generally be a difference (Dm) between the
readings with each of the weights on the pan. This may be expressed in terms of the
following equation:

A = B + Dm

It is important to follow a consistent nomenclature when carrying out weighings and


expressing the comparison in terms of a simple equation is perhaps the best way to do
this. Throughout this section the concept of an equation is used to describe a
comparison weighing.

If the test-weight is of an unusual nominal value it will be necessary to use a standard


made up of several calibrated weights. The number of comparisons carried out in order
to calibrate a weight, and the number of independent mass standards used will depend
on the uncertainty requirements of a particular calibration.

There are two popular techniques for comparative calibration ABA and ABBA
calibrations.

3.5.1 ABA Calibration


This calibration involves the comparison of two weights, normally an unknown and a
standard by placing each one in turn on the balance pan and noting the reading. The
process is symmetrical in that the weight that is placed on the pan first is also on the
pan for the final reading. It is possible to vary the number of applications of each weight
according to the application, but three applications of the test weight and two of the
standard would be a typical regime. This leads to the following measurement scheme:

A1 B1 A2 B2 A3

where Ai and Bi represent the balance reading for the ith application of weights A
and B respectively

The difference between weights A and B is then calculated from the equation:

25
A 1 + A 2 + A 3 B1 + B 2
Dm = -
3 2

This may be generalised for an ABA comparison involving n applications of weight B to


become:
n +1 n

å Ai åB i
i =1 i =1
Dm = -
n+1 n

Two quantities, the first differences and the second differences, followed by the average
difference may be calculated as the calibration progresses. This process is shown for an
ABABA weighing.

Reading A Reading B 1st Diff Average


A1 2nd Diff
B1 A1- B1 (A1 + A2)/2 - B1
A2 A2- B1
B2 A2- B2 (A2 + A3)/2 - B2
A3 A3- B2

In some applications the sample standard deviation of the final column is taken to
indicate the standard deviation of the weighing process. This is not statistically correct
as some of the readings contribute to more than one of the average values (so meaning
that they are not truly independent measurements). Hence this type of pseudo-statistical
analysis should be avoided.

It is much better to make several totally independent ABA based comparisons and look
at the statistical agreement between them rather than attempt to make an invalid
statistical analysis on a single set of comparisons. In the case of an E2 calibration made
at NPL four or five ABABA comparisons would be made, the results for each calculated
as shown on the previous page, and the sample standard deviation of all four or five
results evaluated.

26
Example
The calibration of a 10 g F1 weight against two standards:

Standard 1 has a conventional mass value of 9.999 969 3 g

Test weight (A) versus standard 1


A1 A2 A3 B1 B2
9.999 974 9.999 976 9.999 978 9.999 892 9.999 896
Mean 9.999 976 9.999 894
Difference 0.000 082

Conventional mass value = 9.999 969 3 + 0.000 082 g = 10.000 051 g

Repeat
A1 A2 A3 B1 B2
9.999 978 9.999 978 9.999 980 9.999 894 9.999 896
Mean 9.999 979 9.999 895
Difference 0.000 084

Conventional mass value = 9.999 969 3 + 0.000 084 g = 10.000 053 g

Standard 2 has a value of 9.999 969 0 g.

Test weight (A) versus standard 2


A1 A2 A3 B1 B2
9.999 976 9.999 980 9.999 978 9.999 910 9.999 908
Mean 9.999 978 9.999 909
Difference 0.000 069

Conventional mass value = 9.999 969 0 + 0.000 069 g = 10.000 054 g

Repeat
A1 A2 A3 B1 B2
9.999 980 9.999 980 9.999 982 9.999 910 9.999 912
Mean 9.999 981 9.999 911
Difference 0.000 070

Conventional mass value = 9.999 969 0 + 0.000 070 g = 10.000 055 g

Mean of four measurements = 10.000 053 g

Standard deviation = 0.000 001 7 g

Standard Uncertainty = 0.000 000 9 g

27
3.5.2 ABBA Calibration
The ABBA calibration method is potentially more efficient than the ABA method in terms
of the number of weight applications required to produce a calibration result of a
particular uncertainty. In this case the weights are applied to the balance pan in the
following order:

A1 B1 B2 A2

where An and Bn represent the balance reading for the nth application of weights
A and B respectively.

It is important in this weighing method that weighings B1 and B2 must be independent of


each other. To achieve this on a typical electronic comparator the weight must be
removed from the pan between readings. If a balance with an arrestment mechanism is
being used it is usually sufficient to arrest the balance between the measurements.

The mass difference (Dm) is calculated from the following:

A 1 + A 2 B1 + B 2
Dm = -
2 2

28
Example
This example illustrates the calibration of a 1 g E2 weight by carrying out two ABBA
comparisons against each of two standards.

Using standard 1 (calibrated value 1.000 005 7 g)

A1 A2 B1 B2
-0.024 1 -0.024 0 -0.017 5 -0.017 3
Mean -0.024 0 -0.017 4
Difference -0.006 6

Conventional mass value = 1.000 005 7 - 0.000 006 6 = 0.999 999 1 g

A1 A2 B1 B2
-0.024 0 -0.024 3 -0.017 5 -0.017 7
Mean -0.024 2 -0.017 6
Difference -0.006 6

Conventional mass value = 1.000 005 7 - 0.000 006 6 g = 0.999 999 1 g

Using standard 2 (Calibrated value 0.999 997 8 g)

A1 A2 B1 B2
-0.024 2 -0.023 8 -0.026 2 -0.026 1
Mean -0.024 0 -0.0262
Difference +0.002 2

Conventional mass value = 0.999 997 8 + 0.000 002 2 g = 1.000 000 0 g

A1 A2 B1 B2
-0.023 7 -0.024 1 -0.025 7 -0.026 1
Mean -0.023 9 -0.025 9
Difference +0.002 0

Conventional mass value = 0.999 997 8 + 0.000 002 0 g = 0.999 999 8 g

Mean value = 0.999 999 5 g


Standard deviation = 0.000 000 5 g
Standard uncertainty = 0.000 000 3 g

3.6 Cyclic Weighing


Often it is necessary to calibrate several weights of the same nominal value. In this case
it would be inefficient to calibrate each weight against two standards of the same value.
A technique known as cyclic weighing is used to rationalise the number of
measurements.

29
Cyclic weighing gets its name because the first weight weighed is also weighed last so
that it is possible to account for drift in the measurements. This first weight is normally a
mass standard (S1) that is used to assign values to the test-weights (T1...Tn). A second
standard (S2) is also included in the weighing scheme so that the values assigned using
the first standard may be checked against an independent standard. At its most simple
a cyclic weighing scheme for four test-weights would consist of the following
measurements:

Weight on pan Balance Resulting equation Corrected Result


Reading
S1 R1
T1 R2 T1= S1+ R2 - R1 = S1 + Dm1 T1=S1+Dm1-corr
T2 R3 T2= T1+ R3 - R2 = T1 + Dm2 T2=T1+Dm2-corr
S2 R4 S2= T2+ R4 - R3 = T2 + Dm3 S2=T2+Dm3-corr
T3 R5 T3= S2+ R5 - R4 = S2 + Dm4 T3=S2+Dm4-corr
T4 R6 T4= T3+ R6 - R5 = T3 + Dm5 T4=T3+Dm5-corr
S1 R7 S1= T4+ R7 - R6 = T4 + Dm6 S1=T4+Dm6-corr
Corr = (R7 - R1)/6

In this type of cyclic weighing it is imperative that all of the measurements are made in
quick succession with an approximately equal time between each. From the above it
may be seen how the value of each weight is calculated from that of the previous one.
The value of S1 is used to assign the value of the first test-piece. Theoretically the
measured differences should sum to zero, but in practice there will be some drift in the
measurements. This drift is eliminated by applying a correction (equal to the sum of the
measured differences divided by the number of measurements) to each of the
measured differences. The second stage of the process involves summing all of the
calculated mass differences (SDmi) and then calculating the correction to be applied to
each difference in order to compensate for drift. In this simple example the sum will be
equal to R7 - R1 (ie the difference between the two readings made on S1). This figure is
then divided by the number of calculated differences (in this case six) to obtain a
correction (Corr) to be applied to each difference. As a test of the correction applied and
the arithmetic used the corrected differences should be summed. If this quantity does
not sum to zero, within the uncertainty of the arithmetic used, there is an arithmetic
problem and the calculations must be repeated. As a final check on the calculations the
value of S1 should be calculated using the last difference. It should agree, to within the
accuracy of the calculations, with its quoted value.

The calculated value of S2 should be compared with its standard value. If the calculated
value of S2 is higher than its certificated value by a quantity ES, then the values
calculated should all be reduced by ES/2. The uncertainty in the weighing scheme is
given by ES/Ö3.

30
Example
A set of four 50 g test-objects to be calibrated against S1 = 50.005 g and S2 = 50.001 g.

Weight on pan Balance Resulting equation Corrected Result


Reading
(g)
S1 50.002
T1 49.992 T1= S1+49.992 - 50.002 T1=S1 - 0.011
T2 49.995 T2= T1+ 49.995 - 49.992 T2=T1+ 0.002
S2 50.003 S2= T2+ 50.003 - 49.995 S2=T2+ 0.007
T3 49.895 T3= S2+ 49.895 - 50.003 T3=S2 - 0.109
T4 49.992 T4= T3+ 49.992 - 49.895 T4=T3+ 0.096
S1 50.008 S1= T4+ 50.008 - 49.992 S1=T4+ 0.015
Corr = (50.008 - 50.002)/6 = 0.001 Sum = 0

From the equations in the right hand column the weights would be assigned the
following values from that of S1.

Weight Value from Final Value


S1 (g) (g)
T1 49.994 49.993
T2 49.996 49.995
S2 50.003
T3 49.894 49.893
T4 49.990 49.989
S1 50.005

In this case the assigned value of S2 is 0.002 g higher than its quoted value. Hence the
assigned values of all of the test-objects should be reduced by 0.002/2 g to take
account of this discrepancy. The uncertainty from this weighing scheme is 0.002/Ö3 g =
0.0012 g.

The simple cyclic weighing scheme described above is only suitable for low accuracy
applications as it is extremely sensitive to balance drift and only involves one application
of each test-piece onto the balance pan (this means that there is no check that the
correct weight has been applied).

If a cyclic weighing scheme is to be used for a more demanding calibration it is more


usual to make a series of comparisons in the manner shown below. Here each weight in
the cycle is compared with the previous one using either an ABBA or ABA technique.

31
Comparison Measured Corrected Using S1 value Using S2 value
Difference Difference
S1 v T1 d1 d1 - Corr T1= S1-d1+Corr T1= S1*-d1+Corr
T1 v T2 d2 d2 - Corr T2=T1-d2+Corr T2=T1-d2+Corr
T2 v T3 d3 d3 - Corr T3=T2-d3+Corr T3=T2-d3+Corr
T3 v S2 d4 d4 - Corr S2*=T3-d4+Corr S2=T3-d4+Corr
S2 v T4 d5 d5 - Corr T4=S2*-d5+Corr T4=S2-d5+Corr
T4 v T5 d6 d6 - Corr T5=T4-d6+Corr T5=T4-d6+Corr
T5 v T6 d7 d6 - Corr T6=T5-d7+Corr T6=T5-d7+Corr
T6 v S1 d8 d8 - Corr S1=T6-d8+Corr S1*=T6-d8+Corr
Sum d1+ d2+...+ d8 0
Corr sum/8

A standard is compared with the first weight under test. The first test weight is then
compared with the second test weight et cetera until the final weight is compared with
the standard. A second standard is used at the centre of the weighing scheme in the
same manner as one of the test weights.

This is shown in the third column above. If the correction has been calculated and
applied correctly the sum of the corrected differences will be zero (within the precision of
the arithmetic used). If this is not the case the correction must be re-calculated and
applied.

The values for each of the test-pieces may be calculated from that of the first standard
in the manner shown in the fourth column. In this manner the value of test-piece one is
calculated from that of standard 1 and the first corrected mass difference and the value
of the second is calculated from that assigned to test-piece 1 and the second corrected
difference et cetera. It may be seen from this that a value will be calculated for the
second standard that is based on the weighing scheme and the value of the first
standard.

The final value assigned to each test-piece is the mean of the values obtained from S1
and S2. It is only necessary to calculate the value based on S1 (or S2 if the calibrator
prefers) and then to consider the offset between the assigned value of the other
standard (S2*) and its quoted value (S2). If a weight has been assigned a value of Tx
using S1 then its value taking into account both standards is given by:

S2 - S2 *
T = Tx +
2

Alternatively the values of the test-pieces may also be calculated based on the value of
S2 before taking a mean of the two results. In this case the process starts by assigning
T4 a value based on S2 and the corrected difference d5 - Corr. The procedure then
assigns values to all of the weights in turn with S1 being given a value based on that
assigned to T6. This assigned value of S1 is then used to give a value to T1 with the

32
process continuing until S2 is assigned a value from T3. However, this is a much more
laborious method of carrying out the calculation.

The uncertainty in the weighing process may be calculated by looking at the difference
between the assigned value of one of the standards (eg S2*) and its quoted value (S2):

S2 - S2 *
Uw =
3

The denominator reflects the fact that this uncertainty has a rectangular distribution (ie it
is based on two points which represent the extremes of a range).

3.7 Weighing by Sub-Division


Sub-division is used for the most demanding applications. It involves the use of
standards of one or more value to assign values to weights across a wide range of
values. A typical example of this would be to use two or three 1 kg standards to
calibrate a 20 kg to 1 mg weight set. Equally it would be possible to use a 1 kg and a
100 g standard for such a calibration. Typically this is used for E1, and sometime E2,
calibrations.

This is most easily illustrated by considering how values would be assigned to a weight
set using a single standard. In reality the weighing scheme would be extended to involve
one or two other standards. The standard is compared with any weights from the set of
the same nominal value and also with various combinations of weights from the set that
sum to the same nominal value. A check-weight, which is a standard treated in the
same manner as any of the test-weights, is added in each decade of the calibration so
that it is possible to verify the values assigned to the weight set. In the case of a 1 kg to
100 g weight set the following minimal weighing scheme may be used:

1000 =1000S
1000 = 500 + 200 + 200D +100C
1000S = 500 + 200 + 200D +100
500 = 200+ 200D+100
500 = 200+ 200D+100C
200 = 200D
200 = 100 +100C
200D = 100 +100C
100 = 100C

In a simple case such as this it is possible to calculate the values of the test weights
manually, but in a realistic situation when several standards are used and information is
required about the weighing scheme uncertainty it is necessary to undertake a least
squares analysis using a computer.

The subdivision weighing scheme has the following advantages:

33
a) it minimises use on (and hence wear on) standards
b) it produces a set of data which provides important statistical information about the
measurements and the day to day performance of the individual balances
c) there is a redundancy of data (ie more measurements than unknowns).

It has the following disadvantages


a) it requires a relatively complex algorithm to analyse the data
b) it necessitates placing groups of weights on balance pans (this can cause problems
for instruments with poor eccentricity characteristics or automatic comparators
designed to compare single weights.

3.8 Make-weights
Make-weights are weights that are added to a load in order to make it approximately
equal in weight to the object it is being compared with. Make-weights are added so that
only a small part of the comparators scale is used during a comparison. For example if a
100 g scale had a 1 % error it would equate to a 0.2 g error if there were a 20 g
difference between the loads under test, but this could be reduced to a 0.01 g error if a
make-weight is used to balance the loads to within 1 g.

If a make-weight of value Mw is added to load A during a comparison the weighing


equation becomes:

A + M w = B + Dm

so in order to calculate the value of A from of B it is necessary to subtract the value of


any make-weights used in association with A from the sum of the value of B and the
calculated mass difference. Similarly, if make-weights had been used with weight B the
value of the make-weights would be subtracted from the value of A before taking the
difference in the balance readings into account.

Example
A 1.000 023 g make-weight is added to weight A when it is compared with standard B.
The calculated difference between the readings indicates that A together with the make-
weight is 0.000 546 g heavier than B. Therefore B is heavier than A by 1.000 023 -
0.000 546 g = 0.999 477 g.

3.9 Two Pan Balances


The use of two pan balances is gradually decreasing with the increased popularity of
electronic balances. However, there are still applications were the two pan balance has
advantages over a single pan balance. The use of a two-pan balance is generally more
complex than that of a single pan balance with there being several more issues to
consider in its use.

34
In using a two pan balance it is usual for the rest-point (the orientation of the loaded
balance if its swing was allowed to fully decay) to be calculated from a set of three or
more turning points (the extreme points of the beam’s oscillation). If five turning-points
were taken the rest-point would be:

æ æ tp + tp 3 + tp 5 ö æ tp 2 + tp 4 ö ö
rp = ç ç 1 ÷ +ç ÷÷ ¸ 2
èè 3 ø è 2 øø

The most common methods to use in comparing a pair of weights on a two pan balance
are double-double or double substitution weighing.

3.9.1 Double-Double Weighing


If the weighing is expressing in the same form as for a single pan balance (ie A=B+Dm)
then weight A should be placed on the left hand balance pan and weight B on the right
hand pan. The beam rest point would be determined for this configuration and then the
weights swapped over before taking a second reading. The second configuration is then
repeated before returning the weights to their original configuration for the final
measurement. The readings (r1, r2, r3, r4) are combined as follows to give a mass
difference Dm:

ær + r r +r ö s
Dm = ç 1 4 - 2 3 ÷ ×
è 2 2 ø 2
where s is the balance sensitivity

A more simplistic variation of this, the double weighing, involves just a simple exchange
of the weights and the calculation of the difference between them. However, such a
measurement regime does not take any account of drift in the balance.

3.9.1a Use of Make-weights


Great care should be taken in the use of make-weights on a two pan balance. If a
make-weight is applied to a pan during the first measurement it must also be applied
during the fourth one. Similarly the make-weights must be applied consistently during
the second and third measurements. It is important to apply the correct sign to any
make-weights applied during the weighing process. The following table indicates the
sign of the correction to be applied to the final calculated mass difference when using
double-double weighing:

Position/Measurement Measurement 1 and 4 Measurement 2 and 3


Left pan -ve +ve
Right pan +ve -ve

3.10 Double-Substitition weighing


When a two pan balance in used in substitution mode its operation is analogous to that
of a single pan balance. The loads under comparison are placed in turn on one of the

35
pans (for example the right hand pan) while the left hand pan is loaded with a counter-
poise weight that remains in the same position throughout the entire weighing. Care
must be taken to ensure that the counter-poise is heavier than either of the weights
under comparison. An ABBA or ABABA weighing scheme is recommended for this type
of comparison.

3.11 Buoyancy Corrections


If two weights under comparison have different volumes they will be subjected to
differing upthrust from the air. The concept of air buoyancy and the measurement of air
density is described in the section on air density measurement. If weights A and B have
volumes VA and VB respectively the weighing equation for a conventional mass
comparison, including buoyancy correction would be:

A = B + (VA - VB )(r a - 12
. ) + Dm
where ra is the air density in g cm-3

3.12 Gravitational Corrections


The Earth’s gravitational field varies with altitude and latitude. In practice this means
that the gravitational field at sea level is larger than at altitude. If comparative weighings
are carried out using a single pan balance, or a two pan balance which has the loads at
the same height it is not normally necessary to make gravitational corrections. The
exception to this is when extremely high accuracy weighings of the type made in a
national measurement institute are being carried out. In that case it is sometimes
necessary to correct for any difference in height between the centres of gravity of a pair
of weights.

Typically for every 1 mm difference between the heights of the centres of gravity of two
1 kg weights there is approximately a 0.3µg correction to be applied. Therefore if a
10 kg standard with its centre of gravity at height 10 cm is compared with a 10 kg slab
weight with its centre of gravity at 3 cm there will be a 0.21 mg correction to be applied
to the mass difference.

It should also not be necessary to make corrections when a balance is used in a direct
reading mode if it has been calibrated in the location where it is being used. If a balance
has been calibrated at a different location (not best practice) or weights are being used
in a dead-weight tester (for example a pressure balance or a force machine) it may
prove necessary to make a correction for gravity.

In many practical applications it should be possible to calculate the local gravitational


field as follows:

36
g = ge (1 + 0.0053024 sin2 f - 0.0000058 sin2 2f ) - 3.088H m s-2
where ge = value of gravity at equator
g = local gravity
f = latitude of location
H = Height of location above mean sea level

This typically has a 20 parts per million error associated with it due to geological
differences.

The British Geological Survey is responsible for the maintenance of a grid of local
gravity reference stations around the UK and is able to provide data relating to a
particular location. Alternatively a useful internet site containing data relating to locations
around the United Kingdom is run by the Bureau Gravimetrique International
(http://bgi.cnes.fr:8110).

37
ESTABLISHING A MASS LABORATORY

M J Buckley
Standards Officer and General Manager
South Yorkshire Trading Standards Unit

4. ESTABLISHING A MASS LABORATORY

4.1 Construction
This section will deal with the basic requirements for establishing a mass laboratory, in
terms of the construction of the facility and the equipment needed for calibrating
weights.

When establishing a calibration laboratory it is important to consider the type of weights


to be calibrated. There are many laboratories who can calibrate class M1 weights, with
progressively fewer laboratories able to calibrate higher class weights. In most countries
there are only one or two laboratories who have the facilities, equipment, experience
and the need to calibrate class E1 weights.

Consideration also needs to be given to the needs of the end user – what are the typical
sizes and classes of weights that need to be calibrated. For lower class weights there
may be as much demand for sandblasting and painting cast iron weights as there is for
calibrating them. You may also need to consider whether an adjustment service is to be
provided.

Whilst there are only a small number of weights in Imperial units still in use in industry,
there are weights which are used for force measurements, or as part of some other
measuring equipment, such as a pressure balance. Many of the newer mass
comparators cannot accommodate these weights because of their physical size or their
non-nominal mass value. Furthermore, for force weights you may need to know about
the difference between mass and conventional mass, and you may need to have a good
working knowledge of the technical units of force, and their relationship with the SI unit
of mass.

Although many of the basic requirements apply to all types of mass laboratory, it is true
that the conditions for higher class weights are more onerous – the costs are also
greater. If the basic facilities are right, however, you should be able to make higher
precision weighings in the same facilities at a later date if you need to.

4.1.1 Location
The choice of location of a laboratory is important. A mass laboratory should not be
located adjacent to railway lines, major roads, heavy industry or canals. All of these
locations are unsuitable due to possible localised vibration. If a site is chosen where
vibration is present, no amount of pneumatic or hydraulic isolators will enable good

38
quality mass measurements to be made. Likewise overhead electric cables, substations
and radio transmitters should be avoided.

If you are likely to need to calibrate a significant number of large weights (over 50 kg)
you will have to consider access – but large goods vehicles near the calibration
laboratory, or overhead cranes installed in an adjacent area can also cause problems.

For laboratories undertaking high precision weighings, it is often recommended that part
of the laboratory should be below ground level – whilst this could give better
temperature stability, it is nowadays possible to build above ground facilities which have
similar levels of isolation from environmental effects. What is not recommended is that
the laboratory should be on the upper floors of a building (although some mass
calibration laboratories have been built like this). Mention has been made of cranes for
unloading – it is also recommended that the laboratory is not near a person or goods lift,
used as a throughway to some other room, used for other purposes (eg equipment
storage) or be adjacent to heavy traffic areas such as canteens.

4.1.2 Construction
The construction of a mass calibration laboratory should ensure that stable conditions
can be maintained without compromising access for weights trolleys or equipment.

The laboratory needs to be sufficiently large to accommodate a wide range of mass


comparators or balances. For laboratories carrying out a full range of work, more than
one calibration area may be necessary, to ensure that when calibrating large weights
the localised environment is not so disturbed that it affects the calibration of smaller
weights. Test areas dedicated to the calibration of weights up to 100 g or 200 g are
generally separated from areas where larger weights are calibrated. In addition it may
be necessary to provide accommodation outside the calibration area for cleaning,
painting and adjusting weights. If a large number of class M1 weights are calibrated,
they should ideally be calibrated in a special test area, to avoid dust contamination of
higher accuracy calibration areas.

Small test rooms can lead to temperature gains if too many people and weighing
machines are located in a small area. If possible allow a minimum of 10 m² for each
person.

Consideration needs to be given to the materials from which a laboratory is constructed.


Doors should be sufficiently wide to enable weight trolleys to be moved in and out of the
laboratory. If possible use sliding doors, to avoid pressure shocks when closing a more
conventional door. Particularly in high precision areas, the materials used for walls
should be low maintenance – regular painting of walls should be avoided. If possible
avoid windows – if this cannot be done use solar control film and consider building an
internal wall or glazed frame some 300 mm or so in from the external wall. Remember
also to insulate all walls – an adjacent room kept at a higher temperature will cause the
parting wall to be hotter.

39
It is worth installing a secondary ceiling, using a grid system and non- fibrous wipeable
tiles. The gap between the primary and secondary ceiling can be used for service
cables. If air conditioning is installed, conditioned air can enter the gap between ceilings,
and then be allowed to diffuse through perforated tiles fitted with removable filter
materials (the filter material should be dust free). This will allow better thermal stability in
the calibration area and ensure that pressure effects caused by the air conditioner do
not affect the balances.

There are problems in choosing a good floor material. It must be easy to clean and yet
not subject to dust formation. Sealed vinyl sheet such as used in operating theatres and
clean rooms is best. Although it will require cleaning this should be by damp mopping
rather than using large quantities of water or cleaning fluids which will change the
humidity levels in the laboratory. Chemical cleaning fluids may also affect the stability of
your standard weights.

4.1.3 Services
When constructing a laboratory you will need to consider the provision and location of
services. At all costs avoid water or heating pipes from being located in the laboratory or
immediate vicinity of a calibration laboratory, as they may affect the temperature of the
test area.

There should be sufficient power points for the balances, environmental measuring
equipment and computers. Ideally, a stabilised supply with battery back-up should be
provided to ensure a continuous stabilised power supply to the mass comparators. By
stabilising the power supply, voltage fluctuations and lightning strikes will not affect the
balances used for calibration. Furthermore, any power outages will be contained, with
the balances being kept thermally stable by being fed from the battery back-up system.

Lighting is important in the calibration area. It should be sufficiently bright to enable


even the handling of milligram weights to be accomplished with ease, but it should not
be so bright that the light fittings generate heat which affects the environment of the
calibration laboratory. Modern fluorescent fittings with electronic starters are ideal
provided they have a metal cover to contain the heat. Since turning lights on and off
affects the laboratory temperature, it is recommended that they are left on at all times.

When calibrating weights it is important to record the balance readings accurately. This
is best undertaken by interfacing the balance to a computer so that data can be
transmitted automatically, and manual transcription errors can be avoided. However,
this may not be possible at first. Nonetheless, it may be expedient to install the trunking
and sockets (even if blanking plates are used at first) when establishing a laboratory, so
that building alterations do not have to be made if a data recording system is installed at
a later date.

40
4.1.4 Balance plinths and other furniture
Balances need installing on suitable tables with (ideally) only one balance on each
table. Do not use ordinary tables, or anti-vibration tables designed for other types of
measurements. Plinths should be strong, not transmit vibration from the floor or
surroundings and not exert any influence on the balance through movement, flexing or
magnetism (through metal reinforcements).

The best and most economic solution is probably the construction of solid brick piers (in-
filled with concrete) supporting a cross table of granite. To avoid dust generation the
bricks should be sealed with a non-flaking paint. Polished granite is suitable as a
balance top (100 mm to 200 mm thick depending on the size of balance). Do not use
reconstituted marble or stone as the dry atmosphere of the calibration area will cause
the glued stone to disintegrate.

When installing the brick support piers please ensure that they are not mounted onto
the laboratory floor, otherwise vibrations will be transmitted to the balance when
operators walk across the floor.

Apart from balance tables you will need to provide cupboards which are dust proof for
housing the standard weights (they should be stored under glass rather than contained
in their transportation boxes - this will avoid instability due to contamination from the
materials from which the box is made).

Finally, it will be necessary to provide trolleys or tables on which computers or


notebooks are kept. When making weighings, notebooks should not be placed on the
balance table. Likewise, a clean area for keeping chamois gloves, clean weight handling
devices and tweezers and brushes clean is required.

Depending on the overall temperature control required it may be necessary to install air
conditioning. Although measurements are referenced to a temperature of 20 °C it is not
necessary to make measurements precisely at this temperature, particularly if making
lower grade measurements or calibrating stainless steel test weights using stainless
steel standards. However it is very important that the temperature remains stable for
long periods, with little or no change in temperature.

Annex C to the draft revised OIML Recommendation R 111 specifies that temperatures
in mass calibration laboratories calibrating higher class weights should generally be
within the limits 18 °C to 27 ° C, and 40 % rh to 60 % rh. (Lower limits might apply when
certain mass comparators are installed which have reduced operating specifications).
The rate of change of temperature (per 12 hours) and humidity (per 4 hours) should not
change by more than:

41
Class E1 Class E2 Class F1 Class F2 Class M1

0.3 °C / 12 hours 0.7 °C / 12 hours 1.5 °C / 12 hours 2.0 °C / 12 hours 3.0 °C / 12 hours

5 % rh / 4 hours 10 % rh / 4 hours 15 % rh / 4 hours Not specified Not specified

If air conditioning is installed it must be arranged in such a way that the conditioned air
does not cause a pressure effect on the load receptor. Furthermore there needs to be a
uniform temperature throughout the test area. This can often be achieved through
blowing the conditioned air into the partition between the primary and secondary ceiling,
allowing it to diffuse through perforated ceiling panels fitted with a dustless filter
material.

4.2 Equipment

4.2.1 Balances and mass comparators


It is not possible to buy just one balance or mass comparator to cover the whole range
of weights which are normally calibrated.

In a calibration laboratory undertaking a normal range of mass calibration work, most


weights will fall in the range 20 kg to 1 mg, although the OIML International
Recommendation currently covers the range 50 kg to 1 mg. (Incidentally the draft
revision of this publication will cover weights as large as 5 000 kg). There are only a
small number of weights in use larger than 20 kg and it should not be necessary to
provide equipment for the larger weights. If you do decide to calibrate larger weights it
should be remembered that the balances and standard weights are extremely
expensive and you will also require manual handling equipment.

In the range 20 kg to 1 mg you need to decide whether you will only be calibrating
weights that comply with R 111. If you intend to calibrate other weights you should
remember that they may be of varying physical forms and dimensions, and may be of
non-nominal values.

When providing balances or mass comparators, you should bear in mind the uncertainty
that you want to achieve in calibrating weights, and the practicality of calibrating the
number of weights that are likely to be submitted to you for testing.

Laboratories calibrating M1 weights in small numbers may find it convenient to have


perhaps four balances or comparators, typically of capacity 2 g, 200 g, 2 kg and 20 kg.
The balances or comparators should be chosen for their ease of use, their performance
under varying temperature and other environmental conditions, and their likely
repeatability when used by you in your calibration laboratory.

If you intend to calibrate higher accuracy weights you may need to have a larger number
of balances or comparators, with extended resolution. If you want to calibrate a large

42
number of weights of all classes then you may need to provide many balances, some
with the same resolution and some of the same capacity but with a different resolution,
size of load receptor or other operating characteristics. For work of the highest
accuracy, such as the calibration of Class E1 weights, some comparators will be
provided with weight handlers or robots to enable weights to be interchanged
automatically, data being collected by computer, with the whole calibration being
undertaken without a human observer. This leads to greater efficiency, enables more
measurements to be made, and allows for better environmental conditions during
measurement by carrying out measurements in a ‘closed’ environment.

Most modern machines used for calibration are now electronic, and do not use a
mechanical beam or lever system with knife-edges and bearings. They have the benefit
of a digital display and can be interfaced to computers for data capture. For low
accuracy weights, electronic balances of general design may be quite satisfactory, and
are relatively inexpensive. For higher accuracy you may need to use a mass
comparator, a high accuracy electronic balance. Unlike the normal balance they may
have higher resolution, but may have several discontinuous weighing ranges or
windows. They may also require to be installed in more environmentally stable
conditions.

For the highest accuracy it is convenient to use a system of balances, interfaced to a


computer operating system, which can record both the weighing results as well as the
environmental conditions at the time weighings are made. This facilitates the calibration
system, enables weighings to be undertaken during the night when conditions are more
stable, and eliminates errors resulting from the incorrect manual recording of
measurement data. A computer operating system also facilitates compliance with
documented measurement procedures, by regulating techniques and prohibiting
unauthorised actions. If a wired computer system is inappropriate, the balances or mass
comparators can be hard wired, with laptop computers linked to a server by radio link.

4.2.2 Weights
Most standard weights are cylindrical in shape with a knob and a recessed base.
Weights of nominal value less than 1 g are of polygonal wire or sheet metal. For the
highest precision weights are solid and of one-piece. However lower grade weights may
be made with a screw knob which protects an adjusting cavity (this facilitates adjustment
of the weights to compensate for wear, but means that the weights are less stable than
solid weights and cannot be used for the highest precision calibrations). Small weights
should preferably be made of wire rather than sheet, as they have better stability. If
sheet metal weights are used they must be kept very clean as more of the weight is in
contact with the load receptor than is the case with wire weights.

Generally, sets of weights come in the series 5, 2, 2, 1, although older sets may have a
different sequence or contain values which are no longer in normal use, such as 5, 3, 2,
1. Sets of weights are supplied in wooden transit boxes, but they should be stored in the
laboratory out of their boxes under glass.

43
Virtually all weights are made in compliance with the OIML International
Recommendation R 111, which prescribes constructional requirements and specifies
materials, tolerances, surface conditions, density and markings. The highest precision
weights, Classes E1 and E2 are solid stainless steel, with a prescribed density. They
have no weight markings, as the marking would itself attract dirt. Lower grade weights
are usually adjustable, and may be of other materials including chrome-plated brass,
brass or painted cast iron. Except for the lowest grade of weights, stainless steel is the
preferred material, because of its stability, ability to be polished during manufacture and
for its density. Smaller weights may be of stainless steel, german or nickel silver, or
aluminium. Because of its density, aluminium should only be used for weights up to 10
mg.

When calibrating weights over 50 kg it is normal to calibrate the larger standards using a
series of weights of the same value, which can be calibrated on an automatic mass
comparator and then stacked to form a weight equivalent to the larger standard. Whilst
this facilitates quite low uncertainties, it is a very expensive process requiring
considerable investment in stacking weights and also special mass comparators.

In an ideal world a calibration laboratory should have a minimum of two sets of


standards, ensuring that reliance is not then placed on a single standard which may
change value between calibrations without the laboratory who are using this standard
having any knowledge of this change. For higher accuracy work, a simple one-to-one
calibration is insufficient, with calibration in groups, using a restraint or check standard
(usually at the 1, 10 or 100 level). When a large number of low accuracy weights are
calibrated it may be necessary to have other low grade weights to avoid wear on the
laboratory’s main standards.

For most calibration work, the standard weights will be at least one class better than the
weights to be calibrated. However, high accuracy standards require more care in
storage, handling and use if they are not to change significantly between calibrations.
As a general rule, the value of a standard weight should not change by more than 50%
of the calibration uncertainty between calibrations. Ideally, this rate of change should be
much less than this. Except in special circumstances laboratories should not buy Class
E1 weights for normal use as it is very difficult to maintain them.

4.2.3 Environmental monitoring


Apart from balances and standard weights we need equipment to monitor the
environment. This means equipment to measure air temperature, pressure and
humidity. For high accuracy measurements of carbon dioxide may also be necessary.
We use this equipment to ultimately measure air density, and to make buoyancy
corrections if required. OIML R 33 specifies a limit on air density deviations when
measuring conventional mass.

To determine air density you can use either the BIPM formula or, for lower grade
measurements, the simplified NIST formula. For the highest accuracy it is also possible

44
to measure air density using special weights which have different volumes. Other
experimental ways of measuring air density are also being researched.

In general, for low grade F2 and M1 measurements we only need to ensure that the
temperature remains within the limits specified. For higher accuracy measurements we
may also need to measure air pressure and humidity. We also need to monitor the rate
of change. This is best measured using a recording thermograph (not a recording
thermometer which may only measure the temperature once every few minutes). In an
automated laboratory an environmental measuring and recording system can be
installed with all environmental data being obtained automatically, twenty four hours a
day.

For simple temperature measurements a total immersion mercury-in-glass thermometer


such as a BS 1365 SB 25C thermometer is sufficient. Note that these are quite small
and can be installed in a balance case. They should be used vertically – if used
horizontally a measurement error may occur. For greater accuracy the best sensor is
probably a thermister, a small bead sensor is best. These can be calibrated to high
precision over a limited range. Such sensors can be installed throughout the test area or
in every balance if required (be sure that the sensor is measuring the temperature of the
environment and not a local hot spot such as near a door motor).

Although relative humidity measurements are relatively unimportant by comparison, the


operating conditions of modern mass comparators means that the humidity level in the
test area must be known even if corrections are not made for it. The simplest device is
now probably the electronic hygrometer with sensor. Used in humidities of around 50%
rh they have reasonable accuracy. These instruments are less suitable for measuring
extremes of humidity, but these conditions should not exist in a mass calibration
laboratory. Hair hygrometers and similar devices are generally unsuitable, being difficult
to calibrate and relatively unstable.

When measuring air pressure there are two convenient devices, mercury barometers
being generally unsuitable as they are difficult to use accurately and difficult to calibrate.
In previous times the precision aneroid barometer was used – these can often be
bought second hand from specialist repairers. The most practical instrument is now the
electronic pressure transducer (battery or mains operated). These transducers can be
bought in a barometric pressure option. They will need fairly frequent calibration at first
to ensure that they are stable. They have the advantage of being very easy to use.

4.2.4 Other equipment


If you are to offer a comprehensive calibration service, particularly at the higher levels of
accuracy, it will be necessary for you to consider other requirements given in OIML
Recommendation R 111 for such things as the density of weights, magnetic
susceptibility and surface roughness. The Annexes to the new draft Recommendation
given some details of what is required.

45
For density, there are probably three practical ways of measurement. For laboratories
working at the highest levels an automated density comparator is probably the best.
Here a mass comparator with a handler device for placing the standard volumes and the
test weights in a controlled temperature bath of FC40 liquid allows the unknown weight
to be calibrated against a standard with a known density and volume. The flourinert
liquid FC40 is used in preference to water as it is less susceptible to the formation of air
bubbles - the downside of this is that the density of the FC40 must be measured
regularly.

Another system for making density measurements involves hanging a pan below a
balance (a hydrostatic balance) and immersing the test weight in distilled water. This is
a fairly common technique.

Finally, you could use a pyknometer to immerse the weight in water and weigh it on a
conventional balance or mass comparator. These pyknometers are now commercially
available - they are generally best suited for large weights of nominal mass 5 kg or
more. With some experience they can be used to determine whether the density of the
test weight falls within the limits detailed in R 111.

The magnetic susceptibility of weights is also specified in R 111. There are a number of
ways of measuring this, including the use of a mass comparator and a magnetic
standard, a test method which was developed at BIPM. The most practical methods are
either with the attracting beam balance (which is commercially available in the United
Kingdom) or by an electronic measuring device which can use standard reference
samples supplied and calibrated by NPL.

The measurement of surface roughness requires a roughness measuring machine. For


most purposes a visual examination will often suffice.

4.2.5 Legal metrology


Where weighing or measurement is undertaken for trade purposes (including certain
pharmaceutical measurements) the equipment used may need to meet the legal
metrology requirements detailed in the Weights and Measures Act. Weights may fall
under this legislation, although few if any high class weights are certified for legal
metrology purposes in the United Kingdom. If you are unsure whether this legislation
applies to the equipment you are using or which you have been asked to calibrate, you
should contact your local weights and measures or trading standards authority. If they
are unable to help, contact the National Weights and Measures Laboratory.

46
AIR DENSITY MEASUREMENT AND BUOYANCY CORRECTION
Stuart Davidson
National Physical Laboratory

5. AIR DENSITY MEASUREMENT AND BUOYANCY CORRECTION

5.1 INTRODUCTION
The measurement of air density is necessary in the field of mass measurement to allow
buoyancy corrections to be made when comparing weights of different volume in air. It is
particularly important when comparing weights of different materials or when making mass
measurements to the highest accuracy.

5.2 DEFINITIONS

5.2.1 True Mass


The mass of a body relates to the amount of material of which it consists. In terms of the
calibration of weight it is referred to as true mass in order to differentiate it from
conventional mass which is generally used to specify the value of weights (see below).
The international prototype of the kilogram for which the mass scale throughout the
world is realised is defined as a true mass of exactly 1 kilogram. All high accuracy
comparisons should be performed on a true mass basis (including class E1 and E2
calibrations) although values may be converted to conventional mass when quoted on a
certificate.

5.2.2 Conventional Mass


This is the value normally quoted on a certificate and is the conventional value of the
result of a weighing in air, in accordance with International Recommendation OIML R
33. For a weight at 20 °C, the conventional mass is the mass of a reference weight of a
density of 8000 kg/m3 which it balances in air of a density 1.2 kg/m3.

A conventional mass value for an artefact can be calculated from a true mass value
using the following equation:

æ æ 1 1ö ö
. ´ 10 -3 ÷
M c = M t ç1 + ç - ÷12
è è 8 rø ø

Where: Mc is the conventional mass of the artefact in grams


Mt is the true mass of the artefact in grams
r is the density of the artefact in g/cm3

47
5.2.3 Buoyancy correction
This is a correction made when comparing the mass of artefacts of different volumes. It is
equal to the difference in the volumes of the artefacts multiplied by the density of the
medium in which they are compared (usually air). When comparing true mass values the
buoyancy correction to be applied between two artefacts can be given by the following
equation:

BC = (V1 - V2 ) ´ rair

Where: BC is the buoyancy correction to be applied


Vn is the volume of the artefact n (ie Mn / rn)
rair is the density of the air at the time of comparison

The calculated buoyancy correction should be applied in the form:

M t 1 = M t 2 + BC

Where: Mt n is the true mass of artefact n

When calibrating standard weights comparisons are normally performed on a


conventional mass basis. For such comparisons the buoyancy correction depends on the
difference in air density from the conventional value of 1.2 kg/m3 Because of the way the
conventional mass is specified, comparisons made in air of density exactly 1.2 kg/m3
require no buoyancy correction no matter what the volume difference of the weights being
compared. For comparisons performed on a conventional mass basis the buoyancy
correction is given by the following equation:

. ´ 10 -3 )
BC = (V1 - V2 ) ´ ( rair - 12

The buoyancy correction is applied with the same convention as for the true mass
correction, ie:

M c1 = M c 2 + BC

The equation for conventional mass buoyancy correction is more complicated than for
true mass and care must be taken with the sign of the correction. If weight 1 has a greater
volume than weight 2 and is compared in air of density greater than 1.2 kg/m3 the
correction (BC) will be positive.

5.3 The application of buoyancy corrections


The table below shows the magnitude of the buoyancy correction when comparing
weights of stainless steel with those of another material in air of standard density
(1.2 kg/m3) on a true mass basis.

48
Material compared with Buoyancy correction
Stainless steel (ppm)
Platinum Iridium 94
Tungsten 88
Brass 8
Stainless Steel 7.5*
Cast Iron 24
Aluminium 294
Silicon 365
Water 875

*This is the result of comparing two types of stainless steel, with densities 7.8 and 8.2
g/cm3

Table 1: Buoyancy Corrections when Comparing Dissimilar Materials in Air

The table shows that even when comparing weights of nominally the same material (such
as stainless steel) attention must be paid to buoyancy effects when the best uncertainty is
required. When comparing weights of dissimilar materials the effect of air buoyancy
becomes more significant and must be applied even for routine calibrations when true
mass values are being measured.

When working on a conventional mass basis the buoyancy corrections become smaller.
The OIML recommendations R 33 use a range for air density of 1.1 to 1.3 kg/m3 (ie.
approximately ± 10% of standard air density) meaning the corrections are about one tenth
of the true mass corrections. This together with the limits specified by OIML R 33 for the
density of weights of Classes E1 to M3 mean that the maximum correction for any weight
is one quarter of its tolerance. This is generally not significant for weights of Class F1 and
below (although allowance should be made for the uncertainty contribution of the un-
applied correction) but for Class E1 and E2 weights buoyancy corrections need to be
applied to achieve the uncertainty values required for weights of these Classes.

5.4 MEASUREMENT OF AIR DENSITY

5.4.1 Determination of air density from parametric measurements


The standard method for determining air density involves the measurement of
temperature, pressure and humidity. From these measurements, and taking into account
carbon dioxide concentration for the best accuracy, the density of the air can be
calculated. The empirical formula for the calculation of air density recommended by the
Comité International des Poids et Mesures (CIPM) was derived by Giacomo [1] and
modified by Davis [2]. Table 2 shows typical and best achievable uncertainties for the
calculation of air density from the above parameters using the CIPM formula.

49
Routine Measurement Best Capability
Uncertainty ppm Uncertainty ppm
Temperature (oC) 0.1 360 0.01 36
Pressure (mbar) 0.5 500 0.05 50
Humidity (% RH/oC dew pt.) 5% 350 0.25oC 58
CO2 content (ppm) - - 50 21
CIPM Equation 100 100
Total (x 10-3 kg/m3) 0.86 720 0.16 133
Table 2: Routine and Best Achievable Realisation of Air Density using the
CIPM Formula

5.4.2 Other methods for the measurement of air density


Other methods for the measurement of air density have been investigated, principally by
National Measurement Institutes (such as NPL) and balance and weight manufacturers.
The most common alternative method involves the use of artefacts of different volumes
whose mass difference gives a direct measurement of the density of the air in which
they are compared. In whichever way air density is measured, the conventions for the
application of buoyancy corrections remain the same.

50
UNCERTAINTY IN MASS CALIBRATION
UNCERTAINTY IN MASS CALIBRATION
Pauline Leggat
National Physical Laboratory

6. UNCERTAINTY IN MASS CALIBRATION

6.1 Introduction

When making measurements there is always an element of uncertainty in the result. We


cannot know ‘true’ values – there are limitations in our knowledge and in the
performance of the instruments we are using. Therefore a measurement is not complete
without an estimate of the doubt that surrounds it (the uncertainty) and the confidence
we have in that estimate. This document gives an introduction to the terminology and
the main sources of uncertainty in the calibration of a weight of nominally 100 g, when
using a comparator, plus a brief summary of the process for calculating the uncertainty
and associated confidence level.

6.2 The measurement

Although perhaps an obvious point, before starting it is worth confirming precisely what
the measurements are aimed at determining. In this example it is the conventional mass
of a weight and the following need to be considered:

· Which measurements and calculations will be required to enable you to establish the
mass value and the uncertainty in its determination? For example, will you need to
determine air density?
· How many measurements do you need to take? (The more measurements you take
the more representative the mean (average) value becomes, although there is a
reduction in benefit as the number of measurements increase beyond a certain
point. Ten measurements is a common choice and statistically valid but not always
practical – eg for economic reasons. One way of dealing with this problem is to use
some data from previous measurements to determine the performance of the
balance and then take a smaller number of measurements for the particular
calibration in hand. This is dealt with in paragraph 6.3.3 under repeatability.)
· How to take your measurements and calculate a mass value.

6.3 The uncertainty budget

Once you have determined the mass value you are ready to start calculating the
uncertainty in the measurement. The following table is a typical layout for an uncertainty
budget. It can be in the form of a computer spreadsheet – to make repeated

51
calculations easier - or it can be a paper table completed by hand using a calculator.
Each column in the table is dealt with separately below.

Symbol Source of Value Probability Divisor Sensitivity Standard vi or


uncertainty distribution coefficient uncertainty veff
± mg

6.3.1 Symbol

The symbol used in to denote the input quantity or the influence factor.

6.3.2 Sources of uncertainty

The sources of the uncertainty are dependent on the measurement process and
equation used - as defined in your procedures and by your laboratory environment.
Here we consider the most common sources.

Each of the input quantities in the measurement equation used to calculate the mass
value has an uncertainty. For example if the equation you are using is:

Wx = Ws + DW + Ab

where Wx is the unknown mass


Ws is the mass of the standard
DW is the difference in the balance
readings
Ab is the correction for air buoyancy

then there is an uncertainty associated with each of the input quantities Ws, DW and Ab. The
uncertainty in DW depends on uncertainty due to other influence factors:

· Rounding errors in the comparator readings (dId)


· The uncertainty due to less-than-perfect repeatability of the readings (W R)
· The value associated with comparator linearity (dC)

52
Another influence factor to be considered is

· The drift of the standard with time (Ds)

Symbol Source of Value Probability Divisor Sensitivity Standard vi


uncertainty distribution coefficient uncertainty or
± mg veff
Ws Calibration of
mass standard
dC Comparator
linearity
Ab Air buoyancy
Ds Uncorrected
drift of the
standard
dId Digital rounding
error
WR Repeatability

6.3.3 Values

The values associated with the sources of uncertainty are either measured, calculated
or come from a priori (previous) knowledge.

In our example:

Symbol Source of The value


uncertainty
Ws Calibration of The uncertainty in the mass standard is taken from its
mass standard calibration certificate.
dC Comparator Estimated from previous measurements according to your
linearity procedures.
Ab Air buoyancy A calculated uncertainty based on the air buoyancy
correction equation ie (Vs – Vx)(ra – 1.2) (see paragraph
6.4) In this example the volumes have not been measured
and the uncertainty in the volume difference is based on the
values given in the OIML recommendation R111 (if the
volumes are measured the uncertainty can be less) but the
value of air density has been measured.

53
Ds Uncorrected The uncertainty quoted on the mass standard’s calibration
drift of the certificate will not include any contribution for drift in its
standard mass value. The evaluation of this effect is normally the
responsibility of the weight’s owner as they are best placed
to evaluate how much its mass changes between
calibrations. Drift is usually determined by considering how
much a particular artefact has changed its mass value over
a recent period and extrapolating the figure to cover the
period up to its next calibration. In this example no previous
calibration knowledge is assumed and the uncertainty in the
current mass value calibration is also used to estimate the
limits of drift.
dId Digital rounding Each reading is subject to a rounding error. It is taken to be
error ± half the resolution of the comparator. Such errors occur in
the comparator reading of the standard mass and the
unknown mass.
WR Repeatability This is an uncertainty component which is a measure of the
‘spread’ of the repeated readings. It is estimated by
determining the experimental standard deviation of the
mean (see Further reading). In this example a previous
evaluation of repeatability of the measurement process
(from ten comparisons between a mass standard and an
unknown mass) were used to establish a standard deviation
of 0.00017 mg which was then divided by Ön, where n is the
number of readings in the current measurement – in this
case three).

6.3.4 Probability distributions

A probability distribution is a statistical description of how results behave. There are


three distributions commonly used in mass uncertainty budgets: normal (sometimes
called Gaussian), rectangular (sometimes called uniform) and triangular. The following
graphs illustrate these distributions where the zero value on the horizontal axis
represents the mean value of a number of readings, the vertical axis is the probability of
a particular value occurring and the broken lines represent plus and minus one standard
deviation (k=1) - encompassing approximately 68% of the measurement values (ie 68%
of the area under each curve).

54
6.3.4a Normal distribution
Normal distribution
This represents a group of measurements
where the values are more likely to fall 0.45

0.4
closer to the mean value than further 0.35
away from it. Repeated measurements 0.3

are an example of this type of distribution. 0.25

The graph shows normally distributed data 0.2

with a mean value of zero and a standard 0.15

0.1
deviation of ±1. 0.05

-4 -2 0 2 4

6.3.4b Rectangular distribution

Rectangular distribution
This represents a group of
measurements where the values are
0.12 evenly spread between two limits and
never fall outside these limits. An
0.1
example is when using an assumed air
0.08 buoyancy correction (as opposed to a
0.06 measured or calculated value). The
graph shows a rectangular distribution,
0.04
again with a mean value of zero, but
0.02 limits of ±5. In this case one standard
0 deviation is ±2.89.
-6 -4 -2 0 2 4 6

6.3.4c Triangular distribution


Triangular distribution

This is the distribution you get when adding 0.25


two rectangular distributions. An example is
the uncertainty due to rounding errors. The 0.2

graph shows a triangular distribution, again


0.15
with a mean value of zero and limits ±5. In
this case one standard deviation is ±2.04. 0. 1

0.05

0
-6 -4 -2 0 2 4 6

55
6.3.5 Divisor

In order to eventually sum all the individual input quantities they must be quoted with the
same confidence level. This is done by establishing a number by which the input
quantity uncertainty value is divided to convert it to one standard deviation and is
dependent on the distribution as shown in the table below.

Distribution Divisor
Normal 1 or 2
Rectangular Ö3
Triangular Ö6

The normal distribution has a divisor of either 1 or 2 depending on the confidence level
of the value quoted. For example on a certificate of calibration the uncertainty might be
quoted as ‘k = 1’ (~68%) or ‘k = 2’ (~95%) in which case the divisor is the ‘k’ number.
Other values are sometimes used, for example k = 3.

6.3.6 Sensitivity coefficient (ci)

This is a multiplication factor which converts the uncertainty in the value of an input
quantity to a corresponding uncertainty in the output quantity (it sometimes has to
convert both quantity – such as temperature or pressure - to mass and also the right
units). In the example being discussed all the input quantities are already expressed in
the quantity mass and using the sub-unit milligram so the sensitivity coefficient is 1.

An explanation of the air buoyancy correction calculation is in paragraph 6.4.

6.3.7 Standard uncertainty in units of measurand, ui(W i)

In order to add all the components together we need them in the same units and the
values for this column are simply calculated from

Value ¸ Divisor x Sensitivity coefficient (ci)

6.3.8 Degrees of freedom vi

The number of degrees of freedom is “…in general the number of terms in a sum minus
the number of constraints on the terms of the sum”[1]. Before considering this further it
is necessary to first appreciate that measurement uncertainties are considered to fall
into one of two categories, known as Type A and Type B.

56
· Type A uncertainties are those that are evaluated by statistical methods. For
example, uncertainty due to less-than-perfect repeatability of a measurement can
be reduced by calculating a mean value from several measurements.

· Type B uncertainties are evaluated by other means. They cannot be reduced by


taking more measurements – the uncertainty quoted on a certificate of calibration
cannot be made lower by repeatedly reading the certificate for example!

The degrees of freedom, vi, for individual uncertainty contributions are given by:

Type A vi = n-1 where n is the number of measurements used to


evaluate the type A contribution
Type B vi is usually taken to be infinite

6.3.9 Adding it all up

6.3.9a Combined standard uncertainty u(Wx)


To obtain an uncertainty in Wx, the mass value of the unknown weight, the components
have to be added to obtain a combined standard uncertainty. As it is unlikely that all the
errors will have been at their maximum value in any one measurement it is inappropriate
to add them in arithmetically. The recognised way to address this issue is to
arithmetically add the squares of the standard uncertainties and then take the square
root of the result – this process is know variously as taking the root sum of the squares
(RSS) or quadrature summation.

In the example we are considering the summation is:

u (Wx) = 0.0250 2 + 0.0115 2 + 0.0216 2 + 0.0289 2 + 0.0020 2 + 0.00012


= 0.0454 mg

The resulting probability distribution will be a normal distribution unless one rectangular
distribution is much larger than the other components.

6.3.9b Effective degrees of freedom veff

In general the effective degrees of freedom, veff, will not need to be calculated if the type
A uncertainty is less than half of the combined standard uncertainty, there is only one
type A component and at least three measurements have been taken. Otherwise the
effective degrees of freedom will have to be calculated to ensure that the k-factor of 2
will indeed give a confidence level of ~95%.

57
The effective degrees of freedom for the combined standard uncertainty will depend on
the magnitude of the degrees of freedom for the type A contributions in relation to the
type B. If the type B uncertainties are all taken to have infinite degrees of freedom the
relationship is shown using the simplified Welsh-Satterwaite equation

u c4 (y )
v eff =
æ u i4 (y ) ö
çç ÷÷
v
è i ø

where uc(y) is the combined standard uncertainty


ui(y) is the individual type A uncertainty
contribution
vi is the degree of freedom in ui(y)

therefore in our example

0.0454 4
veff =
æ 0.00014 ö
çç ÷
è 9 ÷ø

veff = 3.8E+11

veff in this example is a very large number which can be taken to be infinity. If this value
had been less than 100 a k-factor would have been calculated from a distribution other
than a normal distribution. More information about this can be found in the further
reading list but for our example the k-factor is two.

6.3.9c Expanded uncertainty

The expanded uncertainty U(Wx) is the combined standard uncertainty u(Wx) multiplied
by a k-factor which will give an uncertainty value with a confidence level of
approximately 95%, in this case 2.

The final uncertainty budget for our example is shown below.

58
Symbol Source of Value Probability Divisor Sensitivity Standard vi or
uncertainty ±mg distribution coefficient uncertainty veff
± mg
Ws Calibration of 0.0500 Normal 2 1 0.0250 ¥
standard
weight
dC Comparator 0.0200 Rectangular Ö3 1 0.0115 ¥
linearity
Ab Air buoyancy 0.0216 Normal 1 1 0.0216 ¥
Ds Uncorrected 0.0500 Rectangular Ö3 1 0.0289 ¥
drift of the
standard
dId Digital 0.0050 Triangular Ö6 1 0.0020 ¥
rounding
error
WR Repeatability 0.0001 Normal 1 1 0.0001 9
u(Wx) Combined Normal 0.0454 >500
standard
uncertainty
U(Wx) Expanded Normal k=2 0.0908 >500
uncertainty

6.4 Air buoyancy uncertainty budget

In order to calculate the uncertainty in the air buoyancy correction for entry into the main
uncertainty budget an additional uncertainty budget has to be completed. Air buoyancy
(Ab) is dependent on the volumes of the weights and the air density with the following
relationship:

Ab = (Vs – Vx)(ra – 1.2)

where (Vs - Vx) is the difference in volume between the standard and the unknown
weight
(ra - 1.2) is the difference between measured density of the air and the standard
air density

There are two ways to calculate an uncertainty value for Ab, either working in relative
values or calculating the sensitivity coefficient directly. In our example the volumes have
not been measured so the value of (Vs - Vx) is taken to be the largest difference
possible according to the OIML recommendations [4] when comparing E2 and F1
weights – that is 1.3 cm3 with an uncertainty of ±1.3 cm3. This uncertainty is treated as a
rectangular distribution because the real value may lie anywhere between these limits
and thus the standard uncertainty (u(V)) is equal to ±(1.3 ¸ Ö3). In this example the air

59
density, ra, has been measured as being 1.22 kg/m3 with an uncertainty of ±10% - thus
the uncertainty in (ra – 1.2), that is u(r), is ±0.012 kg/m3. Thus:

Ab = (1.3)(1.22 – 1.2) = 0.026

6.4.1a Relative uncertainty

The relative uncertainty in the air buoyancy, u(Ab)/Ab, is given by

2 2
u ( Ab) æ u (V ) ö æ u ( r ) ö
= çç ÷÷ + çç ÷÷
Ab è (Vx - Vs ) ø è a ( r - 1.2) ø

inserting the values gives

2
æ 1.3 ¸ 3 ö æ 0.012 ö 2
u ( Ab) = Ab çç ÷ +ç
÷ ÷
è 1.3 ø è 1.22 - 1.2 ø

u ( Ab) = 0.0216

This method of calculation works well for an equation where the only operators are
multiplication or division.

6.4.1b Partial differentiation

The other method, partial differentiation, sounds more complicated but is actually quite
straightforward for this type of equation.

Using the same equation, Ab = (Vs - Vx)(ra - 1.2), and the same values as above, we
calculate the sensitivity coefficients - the numbers by which values of u(V), expressed
here in cm3, and u(r), expressed here in kg/m3, should be multiplied to calculate their
effect on the output quantity expressed in grams.

The partial derivative of a simple equation, such as the one we are looking at, is simply
the multiplier for the term for which we wish to calculate the partial derivative. For
example, if our equation is A = B ´ C then the partial derivative of B is C and the partial
derivative C is B.

In our air density problem the partial derivative of (Vs - Vx) is (ra - 1.2) and the partial
derivative of (ra - 1.2) is (Vs - Vx). In the correct terminology this is expressed as :

60
¶ (Ab)
= ( r a - 1.2) = 0.02
¶ (Vs - Vx )

¶ (Ab)
= (Vs - Vx ) = 1.3
¶ ( r a - 1 .2 )

The values 0.02 and 1.3 are entered directly into the sensitivity coefficient column of the
uncertainty budget and the remaining calculations are the same as for the main
uncertainty budget.

Symbol Source of Value Probability Divisor Sensitivity Standard vi


uncertainty distribution coefficient uncertainty or
± mg veff
Vs – Vx Difference in 1.3 Rectangular Ö3 0.02 0.0150 ¥
volumes
ra – 1.2 Difference in 0.012 Normal 1 1.3 0.0156 ¥
air density
Combined Normal 0.0216
standard
uncertainty

The combined standard uncertainty is the same as in the previous method.

Partial differentiation is more difficult when the equation is more complex [1].

6.5 Reporting the results

In the worked example the result would be reported in the form:

100.000 71 g ±0.10 mg

and would be accompanied by a statement explaining how the uncertainty and its
confidence level are calculated such as:

The reported expanded uncertainty is based on a standard uncertainty


multiplied by a coverage factor k = 2, providing a level of confidence of
approximately 95%. The uncertainty evaluation has been carried out in
accordance with UKAS requirements.

This statement was taken from the UKAS document [2].

The uncertainty has been rounded to 0.10 mg; uncertainties should always be rounded
up rather than down to ensure that the value remains within the 95% confidence limit.

61
6.6 Further reading

[1] BIPM, IEC, IFCC, ISO, IUPAC, IUPAP, OIML. Guide to the Expression of
Uncertainty in Measurement, International Organisation for Standardisation,
Geneva. ISBN 92-67-10188-9

[2] UKAS publication M3003. The Expression of Uncertainty and Confidence in


Measurement, 1997.

[3] NPL Best Practice Guide No 11. A Beginner’s Guide to the Uncertainty of
Measurement. (included in the accompanying documentation)

[4] International Organisation of Legal Metrology (OIML), International


Recommendation No 111:1994 Weights of classes E1, E2, F1, M1, M2, M3.

62

You might also like