You are on page 1of 72

First Year Physics

Laboratory

Handbook

Name ……………………………
Group…………. Subgroup………

Professor D. Evans
d.evans@bham.ac.uk
Head of Year 1 Labs – 2013/14

School of Physics and Astronomy


University of Birmingham
Birmingham, B15 2TT
CHAPTER 1 - INTRODUCTORY NOTEs
1.1 Aims of the Laboratory Course………………………………………………….....4
1.2 Administration ................................................................................................................4
1.3 Assessment .......................................................................................................................5
1.4 Richard Knight Prize ………………………………………………………………….6
1.5 What to do if You Miss a Laboratory Session ..............................................................6
1.6 Text-books .......................................................................................................................6
1.7 Laboratory Note Books ..................................................................................................7
1.8 Formal Reports ...............................................................................................................7
1.9 Experiments.....................................................................................................................7
1.9.1 Standard Experiments .............................................................................................................. 7
1.9.2 Phase 1 experiments ................................................................................................................ 8
1.9.3 Phase 2 experiments ................................................................................................................ 8
1.9.4 Project Experiments ................................................................................................................. 8
1.10 Demonstrators. ..............................................................................................................9
1.11 Marking-off Experiments.............................................................................................9
CHAPTER 2 - GENERAL LABORATORY PRACTICE ......................................... 10
2.1 Safety in Laboratories ..................................................................................................10
2.2 Care of Apparatus ........................................................................................................10
2.3 End of a Laboratory Period .........................................................................................10
2.4 Keeping a Laboratory Notebook .................................................................................10
2.4.1 Starting a New Experiment .................................................................................................... 11
2.4.2 Doing the Experiment ............................................................................................................ 11
2.4.3 At the End of the Experiment ................................................................................................ 12
2.5 Hints on Wiring Electrical Circuits.............................................................................12
2.6 Analysis using Mathcad................................................................................................13
CHAPTER 3 - ANALYSIS OF EXPERIMENTAL RESULTS ................................ 16
3.1 Error in a Single Variable ............................................................................................16
3.1.1 Basic Definitions ................................................................................................................... 16
3.2 Standard Forms and Significant Figures ....................................................................18
3.3 "Rough and ready" methods .......................................................................................19
3.4 Resolution ......................................................................................................................19
3.5 Error in a Function of a Single Variable ....................................................................20
3.5.1 Numerical methods ................................................................................................................ 20
3.5.2 Analytic Formula for Function of a Single Variable ............................................................. 20
3.5.3 F = kx .................................................................................................................................... 21
3.5.4 F = xn ..................................................................................................................................... 21
3.6 Function Of Two Or More Variables........................................................................22
3.6.1 F = x + y + z ......................................................................................................................... 23
3.6.2 F = x y / z ............................................................................................................................... 24
3.6.3 F = xpyq/ zr ............................................................................................................................ 25
3.6.4 Speed-up Hint ........................................................................................................................ 27
3.7 Linear Relations Between Two Quantities .................................................................28

2
3.8 Recording Error Calculations ....................................................................................29
3.9 The Significance Of Discrepancies ..............................................................................30
3.10 Counting Random Events ..........................................................................................30
3.11 Goodness Of Fit, or More About Least-Squares ......................................................31
3.12 Covariance and Correlated Errors ...........................................................................33
3.13 Further Reading ..........................................................................................................35
CHAPTER 4 - PRINCIPLES AND OPERATION OF COMMON INSTRUMENTS
4.1 Oscilloscopes ..................................................................................................................38
4.2 Introduction to Oscilloscopes .......................................................................................38
4.2.1 The Cathode-ray Tube ........................................................................................................... 39
4.2.2. The Y-inputs ......................................................................................................................... 39
4.2.3. The Time-Base ..................................................................................................................... 40
4.2.4. Triggering the Time-base ..................................................................................................... 40
4.2.5 X-Y Display.......................................................................................................................... 41
4.2.6 Co-Axial Cable ...................................................................................................................... 41
4.2.7 High-impedance Probes ......................................................................................................... 41
4.2.8 Digital Storage Oscilloscopes ................................................................................................ 41
4.3 Operation of a CRO ......................................................................................................42
4.3.1 Gould Oscilloscopes OS255/OS300 ...................................................................................... 43
4.3.2 Hitachi Oscilloscope V-212 ................................................................................................... 45
4.4 Digital Meters ................................................................................................................47
4.4.1 Principles of DMM Action .................................................................................................... 47
4.4.2 Absolute and Relative Accuracy............................................................................................ 48
4.4.3 Use and Safety Precautions ................................................................................................... 48
4.4.4 Properties of Particular Meters .............................................................................................. 48
4.5 Power Supplies ..............................................................................................................51
4.5.1 Voltage Control ..................................................................................................................... 51
4.5.2 Current Limiting .................................................................................................................... 51
4.5.3 Setting Current Limit on Farnell L30 .................................................................................... 52
4.5.4 Setting Current Limit on Weir 4000 ...................................................................................... 52
4.5.5 Current Control ...................................................................................................................... 52
4.5.6 Floating Output ...................................................................................................................... 53
CHAPTER 5 – LABORATORY REPORTS .............................................................. 55
5.1 The Purpose of a Report...............................................................................................55
5.2 Structure of a Report ....................................................................................................55
5.3 Submitting Reports .......................................................................................................58
5.4 Sample Report ...............................................................................................................58
CHAPTER 6 - PROJECTS......................................................................................... 70

3
CHAPTER 1 - INTRODUCTORY NOTES
1.1 Aims of the Laboratory Course (20 credit module, 03 19752)
The first year physics laboratory is an integral part of your training to become
competent physicists. The first year laboratory is not designed as an extension or as a
reinforcement of materials met in lectures; its key aim is to provide you with the
opportunity to develop gradually practical and transferable skills. However, most of
the experiments are concerned with topics you will meet in your first year. The
experiments will enable you to broaden and deepen your understanding of the
concepts involved.
The laboratory course has several aims. The major ones are to allow you to learn:
• Manipulative skills and use of instruments
• Skills in data analysis, especially error analysis
• Skills in planning, recording and presenting laboratory work and results
• Physics

1.2 Administration
The laboratory course we offer is continually under review and development. These
notes refer to the situation as it exists at the time of writing, but the organisational
details and the experiments on offer may be changed at short notice. You should
watch the notice board and look for any handouts about differences.
During the first semester, all students except those following TPAM and the non-lab
option within Physics with Theoretical Physics take the laboratory module, 03 19752.
(Students taking Natural Sciences and Physics with Business Management take 5
weeks of lab in semester 2 as part of their Communications Skills module). You will
be expected to have finished the first semester (phase 1) experiments by Christmas,
and your Christmas mark will be based on your first semester module performance.
Students must complete the first semester experiments satisfactorily before they can
continue with the second semester (phase 2) experiments.
The first semester laboratory consists of a common set of eight experiments which
are tackled in weeks 1 to 8 by all students. There is no lab in week 9 in order that
students can prepare a formal report on one of their completed experiments and so
that they can prepare for projects which run during weeks 10 and 11. The report is to
be handed in at the end of week 10 of the first term.
During the second semester, the format of lab changes. During the first 6 weeks,
students following Physics and Astrophysics (and Physics and Space Research) take
Astrolab while all other students have a choice of a range of experiments. (‘Physics
with Particle Physics and Cosmology’ may do some experiments associated their
specialised programme). There is no lab in week 7 so that students can work on their
report (on one of their earlier phase 2 experiments) and also prepare for projects
which run in weeks 8 and 9. PAA and PASR students carry out their projects in
Astrolab. During weeks 7 to 9 inclusive, TPAM and PwT students follow a three
week investigation into examples of chaotic behaviour. The whole class ( but not
TPAM and PWT) merges again in weeks 10 and 11 to tackle two weeks of
electronics: experiments with analogue and digital circuits. These exercises serve as a
bridge to the second year laboratories.
The report on the phase 2 experiments is to be submitted at the end of week 10 of
the second term.

The class is split into two groups (X and Y) for laboratory work. Each group is
expected to attend the laboratory for five hours every week on Tuesday (X group) or
Friday (Y group). You will be supervised by the demonstrators, who are members of
staff and postgraduate teaching assistants. The demonstrators take care of the issuing
and marking of experiments and other administrative matters but they are mainly
present to assist you to become expert in the techniques of experimental physics.
Please consult them freely. The academic member of staff responsible for the first
year teaching laboratory is Dr J.A. Wilson (Room 217 West Physics Building; tel.
ext. 44654; email jaw@hep.ph.bham.ac.uk).

1.3 Assessment
The laboratory work that you do forms part of one module for each semester and in
order to proceed to the second year course, there is a minimum standard you have to
achieve in laboratory work. Laboratory report marks are combined with the
continuously assessed laboratory marks. Each experiment, designed to last one week,
is assigned a weight of one unit while the project has a weight of three units and the
report corresponds to two units. The relative weighting of the components within the
module is then (after rounding!): standard experiments 62%, project 23% and
report 15% for each semester. A pass corresponds to a mark for the module of 40%.
The assessment of lab work is as follows. For each experiment you do, a mark will be
awarded. The definitions of the mark ranges are (approximately):

9 - 10 Very impressive work, showing flair.


7½ - 8½ Outstanding work, more done than expected.
7 - 7½ Good work, everything completed and correct.
5½ - 6½ Competent work, a few mistakes or omissions.
4-5 Satisfactory work, a major mistake or omission.
3 - 3½ Not good enough, many mistakes and omissions.
0 - 2½ Very poor work.

The kinds of things that we are looking for in the lab work are:
• Evidence of planning and forethought in carrying out the experiment.
• Good record keeping.
• Sensible experimental technique.
• Good results, including estimates of the accuracy and precision.
• Appreciation of the limitations of the experiment.
• Understanding the Physics.
The key point is to use your lab book as an “online” record. Jot down notes as you go
along. Your notes must be legible and clearly laid out but we do expect a copperplate
presentation. If you find you have made a mistake, just score out the offending part

5
and revise your notes. This skill of writing up as you go along is difficult to achieve
but it pays off handsomely when you embark on an extended piece of work like a
project when you will not have time to “write up offline” at home.
In your laboratory notebooks the following information should be provided:
• Title of experiment
• The date you started the experiment and bench number
• The key points of the theory behind the experiment, but don't copy the manual
into your book.
• A brief description of your experimental routine
• All experimental observations and errors
• Tables of principal results
• Important graphs
• Brief conclusions
These remarks are rather vague! Although it would be possible to produce a list of
examples which illustrate each point, such a list is not helpful, as one thing you will
need to acquire is the flexibility needed to adapt to novel circumstances, and you can't
specify a new situation in advance! You learn experimental physics by doing it!
Please spend some time reading the relevant manual the day before you attend
the laboratory. The purpose of this preparation is to allow you to do some planning,
and to enable you to keep the whole of the experiment in mind when you are
performing its various parts.

1.4 Richard Knight Prize


Richard Knight was for many years the School Tutor. Richard took great interest in
students’ welfare and progress, in particular in the transition between school/college
and University. After his death, Richard’s family generously endowed the School with
a sum of money and, from this, we are delighted to fund an annual prize:
“the Richard Knight Prize is awarded for the best performance in the Year 1
Laboratory, taking into account all aspects of experimental and project work. The
prize will normally be awarded to one student”.

1.5 What to do if you Miss a Laboratory Session?


Although attendance for laboratory sessions is mandatory, you may be absent from
the laboratory class for a good reason, e.g. illness. When you return, a self
certification form (available from the teaching office) should be sent to the School
Tutor (Dr R C Jones). If you miss the laboratory for a valid reason then you will be
awarded your average mark obtained in that phase of experiments. Alternatively, if
you wish it may be possible to catch up on work missed later in the year.

1.6 Text-books
A book which gives a good introduction to physics laboratory practice (but not related
to the experiments you will do) is:

6
Squires G.L., Practical Physics 3rd edition (Cambridge University Press).
In addition, you should use a good book on error analysis, e.g. one of the following;
Barlow, R.J., Statistics (Addison Wesley).
Lyons L., Data Analysis for Physical Science Students (C.U.P)
Taylor, J, R., An Introduction to Error Analysis (University Science Books)
You are not expected to buy these books. Note that a small number of textbooks, data
books and other material is kept in the laboratory. These are for reference only and
may not be borrowed.

1.7 Laboratory Note Books


In all experimental physics it is essential to keep adequate notes. In research scientists
always use bound notebooks, not loose-leaf ones, for recording details and
experimental results. The school policy is that you must have at least three A4 bound
notebooks, two for ordinary experiments and one for projects. Students will be issued
with three A4 lab books at the beginning of term. Linear graph paper is freely
available during laboratory sessions.
These notebooks are to be your diary of everything that you have actually done in
your experiments. It is not intended for "writing up" experiments. More detailed
advice on what to put in your laboratory notebook is given in chapter 2.
N.B. Taking results on scraps of paper is strictly forbidden. Demonstrators have
been known to tidy up the lab at lunch time. You have been warned!

1.8 Formal Reports


Twice during the year you will be asked to write a "Report" on an experiment, one for
each semester. Notices will be posted towards the end of the semester giving details
when the reports are to be written. You will have a free choice of topic from among
your completed experiments, the only restriction is that the second semester report
must be about a second semester module experiment. Further advice on report
writing is given in chapter 5.

1.9 Experiments
Most of the experimental work that you will do falls into one of two broad categories:
a) Standard experiments
b) Project experiments
The laboratory year is divided into three phases (one for each term). Phase 1 and
phase 2 consist of a set of standard experiments followed by a project. You will work
with a partner in each experiment.

1.9.1 Standard Experiments


For standard experiments, you are provided with the necessary apparatus laid out on a
bench, and given fairly detailed instructions in a manual. Since the time available in
the laboratory is restricted, it is advisable to read through the experiment manual
BEFORE your session. The second semester manuals should not be removed from
the laboratory, but are available on the Web (Canvas) for consultation outside
laboratory hours.

7
1.9.2 Phase 1 experiments

In the first semester, you will do eight compulsory experiments which are designed to
introduce basic laboratory skills. Each of these experiments will take no more than
one session (five hours). At the start of the next laboratory session, you and a partner
(who will be assigned to you each week) will move on to the next experiment, no
matter how much of the previous experiment you have completed. You are
responsible for maintaining your own laboratory notebook, which must be
handed to your demonstrator, responsible for the experiment you have just
tackled, at the end of each session.
The Phase 1 experiments are each worth one unit, accumulating to 8 units over the
term with the project (which occupies the last two weeks of term) being worth 3 units.

1.9.3 Phase 2 experiments


The second semester laboratory has a wide range of more extensive experiments of
which you have a choice. For this module you will work with a partner of your
choice, usually the same student worked with during the first semester project period.
In the list of experiments, each experiment is assigned a number of units. Roughly
speaking, one unit corresponds to a laboratory working day of five hours; so a 1½ unit
experiment should be completed in about eight hours. By the end of the second
semester you should have completed at least 11 units: 6 units of phase 2 experiments,
3 units assigned to the project and the 2 units assigned to the electronics labs. This
electronics lab comprises two sets of exercises introducing you to analogue and digital
electronic, an essential component of the year 2 laboratory.
If you find you are taking significantly more time than this schedule indicates, then
you are working too slowly, and you should consult a demonstrator for advice on
speeding up without losing quality. Conversely, if you are working much faster than
this, then maybe you are rushing things and missing important points.
As in Phase 1, each experiment will be the responsibility of a particular demonstrator;
please give her/him your notebook for marking before proceeding to the next
experiment which you have chosen.

1.9.4 Project Experiments


There are two short experimental projects during the year at the end of each semester
(weeks 10 and 11 of Autumn term and weeks 8 and 9 of Spring term). You will work
with another student on some problem for which little or no instructions are provided.
Partners for the first semester project often decide to work together in the second
semester. For this reason, it may be advisable to choose a partner from the same
degree programme as yourself.
The first semester project list will appear on the notice board during week 7 of
Autumn term. You and your chosen partner will sign up for a project during your
laboratory session in week 8/9. The second semester list is displayed during week 5
of Spring term with the sign up sessions in week 6/7 (Note: PWT students do not take
part in Spring projects; instead with TPAM students, they investigate Chaos in weeks
7 to 9). The sign up for projects is on a ‘first come, first served’ basis, so always be
prepared with a second choice if your first choice has already been reserved. Any
student who does not sign up beforehand will be allocated a project arbitrarily.

8
Note that, although the project is worth three units, it will normally run for two
laboratory sessions; the extra unit is to indicate that we expect you to prepare and plan
your project outside normal laboratory hours.

1.10 Demonstrators
Each experiment which you tackle will have assigned to it a particular demonstrator,
who will be generally responsible for advice, marking, etc. (though if your
demonstrator is not available, do not hesitate to approach any other demonstrator or
ask a fellow student). Serious informal discussion between student and demonstrators
and also between students themselves is a very important aspect of lab – physics in
action! If you are stuck, please don’t soldier on in silence.

1.11 Having Experiments marked.


In the first semester module, you will hand in your laboratory book at the end of each
session. Your demonstrator will look at your notes and, during the next session, will
discuss them with you, write comments as necessary and award you a mark. This
procedure means that you will need two notebooks for your main experiments and
your experimental records will alternate between one notebook and the other.
(Research physicists often have separate notebooks for each project.)

Even though there is no strict time for completing a second semester module
experiment, you must still hand your notebook to the demonstrator when you have
finished (“finished” means that all the data has been analysed and results worked out).
The demonstrator will only then assign you to a new experiment for which you
will use your second notebook.

9
CHAPTER 2 - GENERAL LABORATORY PRACTICE

Practice and demonstrators’ guidance will help you to acquire good working habits for
laboratory investigations, but there are some points which are known to require special
emphasis. Brief notes on these appear below.

2.1 Safety in Laboratories


You will be handed a general leaflet on this subject issued by the University for the guidance
of all departments. There will be a copy available for reference in the laboratory. Read it and
apply its lessons to your work. In the Course 1 laboratory there are some special hazards
associated with e.g. X-rays, lasers and high voltages. Be aware of these hazards; your
experiment may not involve any of them but a neighbouring one might.
Food or drink are also not permitted in the laboratory. If you are feeling thirsty or hungry,
please feel free to pop out for a short break.

2.2 Care of Apparatus


The equipment you use in the laboratory is expensive. On average, each experimental set-up
costs £3000. A very important experimental skill is handling equipment with appropriate
sensitivity. Electrical and optical equipment are particularly at risk. If you meet a piece of
equipment which is unfamiliar to you, you must ask. You must also learn how to read and
interpret manufacturer's instruction manuals.
Most electrical measuring equipment can be damaged by misuse. The cost of repair of a
simple meter is often close to the cost of a new one. Please take the utmost care of the
equipment you are using. Think carefully before connecting power to electrical circuits. If in
doubt about the correctness of your circuits, ASK.
Take great care with optical equipment. NEVER touch diffraction gratings, lenses or mirrors
with your fingers and never attempt to clean a front-silvered mirror.
Accidents do, however, happen. Thus, if in spite of all reasonable precautions you do
damage something, please notify a demonstrator promptly so that repairs may be put in hand
immediately.

2.3 End of a Laboratory Period


You must disconnect your apparatus when you have finished the experiment, or at the end of
the working day. This is particularly important during projects. Please leave the work
environment in a tidy state, in particular leave your bench exactly how you found it; put away
all leads etc. that you have fetched from the laboratory trolley.

2.4 Keeping a Laboratory Notebook


Squires, chapters 10-12 gives a clear account on Laboratory Techniques and how to keep a
laboratory notebook. The crucial test of a good notebook is that it provides you (not the
demonstrator!) with a complete record of everything that you have done so that long
afterwards you can recall all relevant facts. Some years ago, there is a series of law cases
about who should have the income from a patent on all the gas lasers being made. Many
companies were sued, and some went out of business. The crucial evidence was one man's
record of three days’ work as recorded in his laboratory notebook in 1961. Imagine losing

10
millions of dollars just because you never got into the habit of keeping good records!
Incidentally under US law a laboratory note book, properly dated and signed by witnesses, is
accepted as establishing priority for patent applications. Unfortunately, that is not the case in
the UK.

2.4.1 Starting a New Experiment


1. Read the relevant handout before starting work, preferably the day before your lab class.
2. Write down the date, experiment name and bench number. (You should also write
down the date at the start of each days work.)
3. Plan the experiment. Become familiar with the apparatus, to see how it works, and how
the various bits hang together. Record preliminary results (always take preliminary
results!), to give you an idea of the range of readings to be expected, etc.. This will also
help you to have a sensible layout on the bench as things you need to adjust carefully
should be close to your best hand.
4. Think about likely sources of error. Identify the main ones and decide how you are
going to deal with them.
5. Plan how you are going to record data. Thus if you are going to record a quantity, and the
theory shows that you will need the cube, then leave space for a column for the cube. Do
not accumulate large amounts of unprocessed data as this will only lead to disaster.

2.4.2 Doing the Experiment


1. Write observations down in your book while you take them (not on bits of paper: they
will get lost or taken away by a demonstrator).
2. Record actual readings, not derived data. Thus if you measure x but need x3, record the
values of x then calculate the cube separately and write this down as well. If you do not
bother to write down x then you will not be able to unscramble the effects of miskeying
your calculator/computer.
3. Draw graphs at the same time as you take the data - to see how the experiment is
progressing. It does not matter if you get the scales wrong, you can draw a decent graph
later. A "wrong" point will be spotted very quickly then, and you can go back and
investigate while every thing is still set-up. This will mean that the record will not flow
smoothly like a laboratory report, but this does not matter. Once you are sure all the data
are correct, you may then use a computer program to fit your data, but not before!
4. Record everything important e.g. circuit diagram used (even if it is in the manual) as
well as the voltages measured, zero error of micrometer even if it was zero. Also record
apparatus numbers - this way you can be sure the inductor you are using this week is the
same one you carefully calibrated last week!
5. Be neat (consistent with 1!) so you can find your way through afterwards. Neatness helps
to avoid errors, but make it a habit rather than a religion.
6. Spread out your written account a bit; leaving spaces so that comments, calculations or
corrections can be added later if needed.
7. Be as brief as possible consistent with 3. Do not omit units on tables of results or graphs
though. No need to write sentences or perfect prose.
8. Do not reproduce the manual, unless, for example, you find it useful in understanding
the theory, or if a diagram in the manual helps to state what was measured.

11
9. Do not remove results that later prove to be wrong: put a single line through them and
add a note stating why they were wrong. Being able to notice, and correct, one's mistakes
is the sign of a reliable experimenter! Correcting fluid should not be used.
10.Always describe what you saw, as well as recording the numbers. Qualitative
observations are just as important as quantitative.

2.4.3 At the End of the Experiment


1. Your notebook will also contain the analysis of results and any error calculations.
Compare your result with accepted values from data books, if possible, and attempt to
explain any discrepancies.
2. A short verbal conclusion should be added, with comments on how the experiment went,
main difficulties, major sources of error, possible improvements etc.

2.5 Hints on Wiring Electrical Circuits


Figure 2.1(a) shows a variant on a Wheatstone bridge (called a Kelvin bridge). There are
various complicated interconnections in this circuit, and it would be easy to make a mistake
in wiring up. Figure 2.1(b) shows the circuit diagram as it might appear in a lab manual,
showing ways of avoiding these mistakes. The first thing is to number places where
connections to components are made at the component terminals. Then colour code the
various wires, e.g. red for positive voltages, black for negative, green for ground, etc. Write
down the colours you choose by the corresponding line on the diagram.

+10V
red red
1 9
2 10
7 8
blue blue
green green
3 11
4 12
green
blue blue
5 13
6 14
black black
0V

Fig 2.1(a) Circuit Fig 2.1(b) Circuit diagram drawn as a wiring diagram as it might
diagram as it might appear in a (very tidy!) laboratory notebook. Note the numbering
appear in a textbook of the terminals, and the proposed colour coding for the connecting
wires.
Lay out the circuit components sensibly on the bench, so that leads can be kept as short as
possible. Keep meters, resistance boxes etc. that you want to read or adjust often near the
front of the bench, where it is easier to get at them. Things that you do not read or adjust so
frequently should go at the back. Then proceed to wire up the circuit, working through the
numbers systematically, e.g. say to yourself (under your breath!) "Connect 1 to 9 with a short
red lead. Connect 1 to +10 V with a long red lead. Connect 2 to 3 with a short blue lead... ".
You might find it helpful to make a little mark on the circuit diagram by each connection as
you make it.
When connecting integrated circuit chips together, remember that there is a standard

12
convention for numbering the pins, and it helps to write down the pin number at the point
where the connection joins the chip symbol.

2.6 Analysis using MatLab

EasyFit User Guide

EasyFit is the weighted least-squares fitting program used in first year laboratory. You will
use this software at some point during most lab sessions, so understanding it well will be
invaluable.
The program’s input panel is given below, with explanations of each control and its function:

(A) The table to which data is input. This


can be done either manually, using
either the mouse or arrows, or via the
Import Data button.
(B) If your data points all have the same
A
error value, you can type it into this box
and click Apply To All. This will copy
the value to every cell of Y Err adjacent
to filled X and Y cells.
(C) If you wish to carry out a non-weighted
fit without error values, select Calculate
fit without errors. No Chi2 value or
errors will be calculated.
(D) Using this button, data can be imported
from either the clipboard (data cut or
B copied from another program) or a file
F
C (including text and spreadsheet files).
(E) The button will delete all data contained
D E within (A), the data input table.
(F) If select, each of these will multiply the
data in their appropriate column by
G
10(value entered into the adjacent box), during fit
calculation.
(G) This panel allows the selection of the
equation to which the data will be fitted
(e.g. linear). Any additional inputs
required for the chosen fit are shown
H below the drop-down menu.
(H) A graph title and axes labels can be
entered here and will appear on the
I graph.
(I) Axes limits can be defined here to
override those calculated automatically.

13
Tip If you wish to print your data in a well presented table or need to manipulate it prior to
fitting, it is suggested that your type it into Microsoft Excel, which has many excellent
formatting and data manipulation features. Once your data is ready, you can copy the
columns that you wish to use and import them to EasyFit via the Import Data button.

The program’s output panel is given below, with explanations of each section/value:

C
B
D

14
(A) Main plot area - This is where the data
points are plotted along with the fit line.
(B) Residual plot – The x-value of each point
is that of the data point, while the y-value
is the y-axis distance between the data
point and the fit line. If the error bar
crosses the central axis, the point agrees
with the fit within errors. In a good fit,
about 70% of points should agree.
(C) Chi2 per Degree of Freedom – This is a
measure of how good the fit is. A value
of 1 is ideal; a value of much less than 1
may indicate that your estimated errors
are too large; a value much greater than 1
indicates a poor fit.
(D) The name and equation of the fit used.
(E) The coefficients of the fit, corresponding
to the equation, with the error on each.

15
16
CHAPTER 3 - ANALYSIS OF EXPERIMENTAL RESULTS

‘Errors using inadequate data are much less than using no data at all.’
Charles Babbage (1791-1871), British mathematician

3.1 Error in a Single Variable


The purpose of experimental work is to find the best value of a physical quantity. However, a
single number is of limited value without an estimate of its accuracy. You should
immediately get into the habit of considering the accuracy of everything you measure. The
statistical error on a measurement can be reduced by repeating it several times and averaging
the results.

3.1.1 Basic Definitions


Let x1, x2, ... , xi, ... , xn be a set of n independent measurements of a single quantity, which
has a true value µ. i is an index, taking each value from 1 to n in turn, so that xi is the symbol
for the ith reading.
DEFINITION: The best estimate of µ is the mean of the set of results;
i=n

_ ∑x i
i =1
x= . (3.1)
n

DEFINITION: The residual di or the deviation from the mean of the ith reading is the
difference between the ith reading , xi, and the mean;
_
d i = xi − x . (3.2)

DEFINITION: The best estimate of the accuracy of a typical member of the set is the
estimated population standard deviation s;
i=n i=n 2
 _

∑d
i =1
i
2

i =1
 x i − x
 
s= = . (3.3)
n−1 n−1
The standard deviation s gives an indication of the spread in values around the mean, x .
The magnitude of s is a measure of the “width” of the distribution. As more data are taken,
s will be determined more accurately but its value will not change systematically.

DEFINITION: The best estimate of the accuracy of the mean is the standard error in the
mean;

 _ 2 i= n

s
∑ x i − x 
i=1
u= = . (3.4)
n n( n − 1)

17
To summarize, note that s has the same value whatever the value of n (though you will have
statistical fluctuations with different samples). Hence u gets smaller as more readings are
taken so the mean becomes more accurate.
u
DEFINITION: The fractional error in the mean is _
. (3.5)
x
This is often quoted as a percentage. It is only used in intermediate calculations (see later).
_
DEFINITION: The final result is always quoted as x ± u. This is called the standard
form.

Example 1
The following values of the resistance of a resistor were obtained;
R[kΩ] 10.7 10.3 10.5 10.4 10.3 10.5
The sum of these values is = 62.7kΩ
The number of readings n is 6, hence the mean value = 62.7/6 = 10.45kΩ
_
In order to calculate the errors, we need to work out the residuals d i = Ri − R for each
reading, and then square these results;
R[kΩ] 10.7 10.3 10.5 10.4 10.3 10.5
di[kΩ] 0.25 -0.15 0.05 -0.05 -0.15 0.05
di2[10-3(kΩ)2 ] 62.5 22.5 2.5 2.5 22.5 2.5
Note the units for everything.
The sum of the residuals ∑d i is zero, as expected (this is a useful check on arithmetic!). The
sum of the squares of the residuals is 115.0×10-3 (kΩ)2 , hence the population standard
deviation is
6
∑d i
2
115.0 × 10 −3
i =1
s= = = 23 × 10 −3 = 0.15 kΩ.
( n − 1) 5
s 015
.
The error (standard deviation of the mean) is u = = = 0.06 kΩ.
n 6
_
Hence the best value of R is (in standard form) R = (10.45 + 0.06) kΩ.
_
The fractional error in R is u/ R = 0.06/10.45 = 6×10-3 (0.6%).

18
Exercise 1
Complete the following table with headings, and give the mean, population standard
deviation, standard error of the mean and the final result in standard form.
Count rate x[s-1 ] Deviation Deviation2
3.4
2.9
3.6
3.1
2.6
3.4
3.4
2.2
3.5
3.1

Mean =.....................................=........................................
Population Standard Deviation =.....................................=........................................
Standard Error of Mean =.....................................=........................................
Count rate =.....................................=........................................

Notes:
(a) The standard deviation of a set of sample readings is σn = √(di2/n), which is slightly
smaller than our estimate s= √(di2/(n-1)) of the population standard deviation. The
latter may also be denoted σn-1 and is the accuracy of any one reading.
(b) Some pocket calculators have statistics programmes built in. Usually they only
calculate σn and σn -1. The figure we really need, the error in the mean, has to be
calculated by dividing σn -1 by √n.
Check your calculator by entering the data of Example 1 and seeing which keys give
the following results:
0.138... (sample standard deviation).
0.151... (population standard deviation).
Divide the latter by √6 , and check that the error in the mean is as advertised.

3.2 Standard Forms and Significant Figures


_
Your final result must always be written in the standard form x ± u. You may choose the
units to suit your taste, e.g.
either G = (1.436 + 0.004)×1010 N m-2 or G = (14.36 ± 0.04) GN m-2 is perfectly acceptable.

19
Sometimes, particularly in tables, a bracket convention is used. e.g. G = 14.36(4) GNm-2
where the number in brackets is the error in the last figure quoted. If two figures are in the
brackets, then this is the error in the last two figures. This convention is becoming more and
more common, but you should be aware that in some tables the figure in brackets is the
power of 10! The context should make it clear which is meant.
Whatever you do, do not use wrong or inelegant forms like:
G = (1.436×1010 ± 4×107) Nm-2 (inelegant)
G = 1.436×1010 Nm-2 ± 4×107 (wrong)
G = 1.436×1010 Nm-2 ± 4×107 Nm-2 (inelegant)
G = (1.436×1010 ± 0.3%) Nm-2 (inelegant and wrong)
-2
G = (1.43575261 ± 0.0036175) Nm (very inelegant and wrong)
The last example is obviously wrong because there are far too many figures quoted.
Remember that u is an estimate made from a relatively small number of readings, and so is
inevitably inaccurate! Roughly speaking, the error in u is about u/√(2n-2) . This shows that
if you take about 10 readings, the fractional error in u is about 1/√18 = 0.2, so that u is known
to one significant figure at best. So once you have u on your calculator, round up (never
down) the error to one significant figure. Always match the number of significant figures in
the error and in the result. Thus if you have three decimal places in the error (e.g. 0.004),
then the result must be given to three decimal places as well (e.g. 10.738, never 10.7377 or
10.74). For an explanation of significant figures refer to your copy of your first year
textbook.

Exercise 2.
Some of the following results are correctly expressed, some are wrong. Mark the correct
ones with a tick, the incorrect ones with a cross. If possible, correct the wrong results.
12.45 kg ± 0.53 g 357800 ± 200
(345.720 ± 0.004) kg 349.2×1023 ± 4.3×1022 mol.
(4.4673 ± 0.033)×104 µg 45.05 m ± 3%

3.3 "Rough and ready" methods


If the number of observations is too small to do a sensible calculation of the standard
deviation of the items about the sample mean, a cruder method of estimating an error may be
adopted. It is usual to quote one-half the total spread of the observations, and then treat it as if
it were a regular u.

3.4 Resolution
Sometimes you will get the same result for a measurement - e.g. a length measured with a
steel rule - every time you measure it. In this case, the error is not zero but given by one half
of the "reading error" or resolution of the instrument being used. Treat this as a normal error
of the mean, u. See Chapter 4 for more on resolution and accuracy.
DEFINITION: The resolution of a reading is the smallest detectable change in the reading.
Obviously, this depends on the apparatus and how it is used. Thus the resolution of a digital
meter is ±1 in the last digit, while that of a rule could be ±1, ± ½ or ±0.1 of a scale division,

20
depending on the quality of the rule and how carefully the measurement was taken. For a
good steel rule, you can take u to be about 0.1 mm which is the most accurate interpolation of
the 1mm graduations on the rule that you can make.
In principle, if you take a large number of measurements and average them, the error u
(which is ∝ 1/√n) can be made as small as desired. However, once it becomes much smaller
than the resolution of the measuring instrument, one is very vulnerable to systematic errors.

3.5 Error in a Function of a Single Variable


So far we have only described how to calculate the error, u, in a mean value x. The result of
an experiment may not be x itself, but some function F of x, or a function of several different
quantities x, y and z, each of which has an error. We want to calculate the error in the result F
from the errors in the readings. We will denote the error on F as σ(F)

3.5.1 Numerical methods


These are best illustrated by example.

Example 2:
In section 3.1 we obtained the result R = 10.45 ± 0.06 kΩ for the resistance of a resistor.
Suppose now we wish to calculate the best value of R3 and the associated error. Since
10.45kΩ is the best value we have for R, the best value we have for R3 is 1141.17 kΩ3. (Not
all the figures are significant, but as we don't know the error yet, we better play safe and keep
a few guard figures for the moment). Now change R by an amount equal to its error, i.e. add
or subtract 0.06 from R, giving a new value. Cube this, giving an extreme value of R3. The
difference between this value and the best value is the error in R3. Putting in numbers, the
upper value of R is 10.51 kΩ (adding the error), and the corresponding upper value of R3 is
1160.94 kΩ3. The difference is 19.77 kΩ3, which rounds up to 20 kΩ3. Thus R3 =
1140±20kΩ3.

3.5.2 Analytic Formula for Function of a Single Variable


Calculus gives us a method of
F
calculating the change ∆F in a result F
resulting from a small change ∆x in x
(fig. 3.1);
∆F dF
∆x ∆F = ∆x
dx
x For our problem ∆x is the error in
x = σ(x), and ∆F is the resulting error
Figure 3.1 Change in F due to change in x.
in F = σ(F). Hence

dF
σ( F ) = σ( x ) . (3.5)
dx

Example 3
For the case of example 2, F = R3 , so σ(R3) = (dR3dR)σ(R) = 3R2σ(R) = 3×10.452×0.06 =
19.6 kΩ3 = 20 kΩ3 after rounding up.

21
Exercise 3
a) Show that when the error in θ is σ(θ), the error in sinθ is cosθσ(θ). At what angle is the
error in sinθ maximum, and what is this error?
σ(sinθ) = ...................................................................................
σ(sinθ) maximum when θ = ...... , and is = ...... .
b) In Bragg diffraction of X rays, the lattice spacing of the diffracting crystal, d, is inversely
proportional to the sine of the diffracting angle θ. If the error in θ is u(θ), find an
expression for the error in d. What is the error in d at θ = 90°?
σ(d) = σ(1/sinθ) = .......................................................................
When θ = 90°, σ(d) = ...................................................................
c) Show that the absolute error in ln(x) equals the fractional error in x.
σ(ln(x)) = ...................................................................
IMPORTANT: The following rules for error propagation are extremely useful.
You are expected to have these at your finger-tips, so that using
them is instinctive. They must be learned and understood.

3.5.3 F = kx
The result F is obtained by multiplying x by a constant k. dF/dx = k, so the error in F is the
product of the constant and the error in x;
σ(F) = σ(kx) = |k|σ(x). (Rule 1)
Note that the sign of k does not matter, hence the modulus sign.

Example 4
The frequency of the mains was measured to be f = (50.01±0.02) Hz. What is the angular
frequency ω = 2πf? Applying rule 1, ω = 2π×50.01 ± 2π×0.02 = (314.22±0.13) s-1. (It is
probably best to keep the two figures in the error here; rounding down is always wrong, but
rounding up would exaggerate the error by nearly a factor of two, and this is probably too
much. You can think of the last figures as "guard figures", which help avoid gross rounding
errors in subsequent calculations.)

Exercise 4
The first order diffraction from a grating of spacing a = 7.0 µm occurs at an angle φ given by
sinφ = aλ, where λ is the wavelength of the light used. If the angle of diffraction is
5.2°±0.3°, calculate the value of λ. Neglecting any error in a, what is the error in λ?
λ = .....................................................................................
σ(λ) = ...................................................................................

3.5.4 F = xn
The result F is obtained by raising x to some power n. Then dF/dx=nxn-1, and σ(F)=nxn-1
σ(x). Dividing both sides by xn = F gives a simple rule to remember;

22
σ( F )
=
( ) = n σ( x)
σ xn
. (Rule 2)
F xn x
The fractional error in a reading x raised to any power n is the product of the modulus of the
power and the fractional error in x;
Note that the sign of n does not matter. You can use percentage errors in this formula if you
wish.

Example 5.
The reactance X of a capacitor has been measured to be (22.53±0.04) kΩ at mains frequency.
What is the value of the susceptance B = 1/X?
Applying rule 2, since B = 1/X = X-1 = Xn with n = -1,
σ(B)/B = |-1| σ(X)/X = 0.04/22.53 = 1.8×10-3.
B = 1/X= 4.438×10-5 Ω-1.
σ(B) = 1.8×10-3 × 4.438×10-5 = 8×10-8 Ω-1.
Hence B = (4.438±0.008)×10-5 Ω-1.

Exercise 5
Use rule 2 to calculate the error in R3 when R = (10.45±0.06) kΩ.
σ(R3) = .....................................................................................

3.6 Function Of Two Or More Variables


Suppose now that the result F of the
experiment is obtained from several F

different measurements, compounded


together in some way. What we do then is
∆x
to change each variable in turn by an ∆y
∆F x
∆F y
amount equal to its error, keeping the other
values fixed at their best estimates, and x

work out the corresponding change in F.


When all the variables have been dealt with, ∆F x
the final result is obtained by squaring, ∆Fy
y ∆F
adding and squarerooting.
Fig. 3.2 shows how this may be explained.
Figure 3.2 The surface F traced out as a function of
Suppose that we have two variables x and y, the two variables x and y, cutaway to show the
and that the result F is obtained from some changes in F due to independent changes in x and y.
formula involving x and y. Then, plotting F, The inset shows how the two changes are combined
in a similar manner to vectors.
x and y on Cartesian axes, the function
F(x,y) traces out a surface (remember that a
surface is a 2-dimensional line!). Consider any point on the surface. The height of the
surface above the x-y plane is F(x,y). Now change x by a small amount ∆x, keeping y
constant. This means that we move along the surface in a direction parallel to the x-axis to a
new point, where the value of F is F(x+∆x,y) The change in F due to the change in x alone is

23
∆Fx = F(x+∆x,y) - F(x,y) .
Similarly, changing y by ∆y, keeping x constant, corresponds to a movement in a direction
parallel to the y-axis, producing a corresponding change in F of
∆Fy = (F(x,y+∆y) - F(x,y) .
Now, both these changes in F, ∆Fx and ∆Fy, are associated with movements in particular
directions. We can think of them as vectors which are perpendicular to each other
(mathematically speaking, orthogonal). The final change in F due to simultaneous changes in
x and y is therefore given by the vector addition rule;
∆F = √(∆Fx2 + ∆Fy 2) .
It is easy to extend this rule to the general case where F is a function of several variables x, y,
z, ...... . Then
∆F = √(∆Fx2 + ∆Fy 2 + ∆Fz 2 + ..... ) . (3.6)
Equation (3.6) can be approximated by a calculus formula just as in the case of a function of
one variable. The partial derivative ∂F/∂x of a function F(x, y, z, ...) is defined as
∂F F ( x + ∆x , y , z , ....) − F ( x , y , z , ....)
= lim
∂x ∆x→ 0 ∆x
with similar definitions for ∂F/∂y, ∂F/∂z, ...... Since ∆Fx = (∂F/∂x) ∆x equation (3.6) can be
written
2 2 2
 ∂F   ∂F   ∂F 
∆F =   ∆( x) +   ∆( y) +   ∆( z ) + ...
2 2 2 2

 ∂x   ∂y   ∂z 
Expressing this in terms of the errors σ(x), σ(y), σ(z), .... , gives the error propagation rule for
calculating the error in F:
2 2 2
 ∂F   ∂F   ∂F 
σ( F ) =   σ( x) +   σ( y) +   σ( z) +.... . (3.7)
2 2 2 2
`
 ∂x   ∂y   ∂z 
A very important condition that must be met in applying this rule is that the errors in the x's ,
y's, z's, .... are small, random and uncorrelated. i.e. they are not related in any way.
(Mathematically, x, y, z, .... are all mutually orthogonal.) The case where this is not true is
discussed later. (section 3.12).

3.6.1 F = x + y + z
The result F is obtained by adding or subtracting different readings x, y, z, ... . The partial
derivatives ∂F/∂x, ∂F/∂y, ∂F/∂z ,... all equal one, so that equation (3.7) becomes
σ (F) = √[σ(x)2 + σ(y)2 + σ(z)2 + ...] . (Rule 3)
In words: the errors in x, y, z, ... are squared, the sum of the squares is found, and the error in
F is the square-root of the sum. Note that the signs of x,y, z,... do not matter. This is known
as ‘adding in quadrature’.

24
Example 6
Two resistors are measured to be (10.45±0.06) kΩ and (5.84±0.04) kΩ. They are connected
in series, what is the combined resistance?
Using rule 3, the sum of the squares is (0.06)2 + (0.04)2 = 0.0052. The combined resistance =
16.29 ± √(0.0052) = (16.29±0.08) kΩ.

Exercise 6
A third resistor of value (7.95±0.06) kΩ is added in series to the two resistors of example 7.
What is the combined resistance now?
σ(R) = .......................................................
R = ............................ ± ............................

3.6.2 F = x y / z
The result F is obtained by multiplying or dividing different readings x, y, z, ... . The partial
derivatives are
∂F/∂x = y/z ; ∂F/∂y = x/z ; ∂F/∂z = xy/z2 .
These can be used straight away in equation (3.7), but a formula which is easier to memorise
comes from noting that
∂F/∂x = (xy/z) 1/x = F/x ;
∂F/∂y = (xy/z) 1/y = F/y ;
∂F/∂z = (xy/z2) = F/z.
Putting these into equation (3.7) and dividing throughout by F2 gives

 σ( F )   σ( x )   σ( y )   σ( z ) 
2 2 2 2

 F  =  x  + y  + z  . (Rule 4)
       

In words: the fractional errors in x, y, z, ... are squared, the sum of the squares is found and
the fractional error in F is the square-root of the sum.

Example 7.
The capacitance of a capacitor of reactance X is C = 1/(ωX). Using the results of examples 4
and 5, what is the capacitance?
Using rule 4, the sum of the squares of the fractional errors is (1.8×10-3)2 + (4×10-4)2 =
3.4×10-6. Hence the fractional error in C is √(3.4×10-6) = 1.9×103, and C =
(1.413±0.003)×10-7 F = (141.3±0.3) nF.

Exercise 7
In a Wheatstone bridge, the value of an unknown resistance R is given by
R = PQ/S .
The known resistors have values P = (10.45±0.06) kΩ, Q = (5.84±0.04) kΩ, S =
(7.95±0.06)kΩ, What is the best estimate of the value of R?

25
Fractional error in P = .............................................
Fractional error in Q = .............................................
Fractional error in S = .............................................
Fractional error in R = .............................................
R = ..................... ± ..............................

Exercise 8
Suppose that the error in a quantity x is σ(x). Rule 2 tells us that the fractional error in x2 is
σ(x2)/x2 = 2σ(x)/x. But if we think of x2 as x times x, then rule 4 tells us that σ(x2)/x2 =
√2σ(x)/x. Which answer is correct, and what is wrong with the wrong answer? (We get a
student asking this question every year!)
Correct answer: σ(x2)/x2 = ....................................................
The other answer is wrong because
......................................................................................................................................................
......................................................................................................................................................
......................................................................................................................................................
......................................................................................................................................................
......................................................................................................................................................
...............................................................................................................

3.6.3 F = xpyq/ zr
The result F is obtained by multiplying or dividing different readings x, y, z, ... , which occur
to different powers. The quickest way to see the answer is to combine rules 2 and 4, the
fractional errors in x, y, z, ... are multiplied by the power, (ignoring signs, as for rule 2), then
squared. The sum of the squares is then found, and the fractional error in F is the square-root
of the sum (as in rule 4).

 σ( F )   σ( x)   σ( y)   σ( z ) 
2 2 2 2

  = p  + q  + r  . (Rule 5)
 F   x   y   z 
Note that the powers p, q, r, ... go inside the brackets before squaring, otherwise rule 5 won't
give the same result as rule 2 in the special case when there's only one variable.

Example 8.
In the experiment on bending a beam, various lengths a, b, c, d are measured. The final
expression contains these quantities in the form ac2/bd3. For the case where a = (206.2±0.2)
mm, b = (25.42±0.01) mm, c = (494.3±0.2) mm, d = (8.354±0.005) mm, what is the value of
the combination?
Using rule 5, the sum of squares of the fractional errors weighted by the powers is
2 2 2 2
 0.2   0.2   01.   0.005 
 206.2  + 2 494.3  +  25.42  +  3 8.354  = 4.7×10-6.

The fractional error in the combination is √(4.7×10-6) = 0.0022 (0.22%), hence ac2/bd3 =
(3189±7) mm-1.

26
Exercise 9
A voltage V = (26.7±0.1) V is applied across a resistor R = (12.0±0.2) Ω. What is the power
dissipated in the resistor? (Use V2/R)
Fractional error in V2 = ...........................................
Fractional error in R = ..............................................
Power = ................. ± ..........................

Example 9.
This example shows the use of the general rule, equation (3.7), and also how some of the
special rules may be applied.
The resistor and capacitor of examples 2 and 5 are connected in series to give an impedance
Z. The expected value of Z is√(R2 + 1/(2πfC)2) = √(R2 + X2) = 24.836 kΩ. To calculate the
error in Z using equation (3.7), we proceed as follows.
Since Z = √(R2 + X2), ∂Z/∂R = ½(2R/√(R2 + X2) = R/Z .
Similarly, ∂Z/∂X = X/Z .
2 2 2 2
 ∂Z   ∂Z  R X
∴σ(Z) =   σ( R) +   σ( X ) =   σ( R) +   σ( X )
2 2 2 2 2

 ∂R   ∂X  Z  Z 
= [R2σ(R)2 + X2σ(X)2]/Z2 = [(10.45)2(0.06)2 + (22.53)2(0.04)2]/24.842
So σ(Z) = √(1.95×10-3) = 0.044 kΩ. i.e. Z = (24.84±0.05) kΩ (rounding up the error figure).
The calculation can also be done using rules 1 to 5. This is lengthier, and one must be
methodical about it, working from the "inside to the outside"; thus:
Using rule 2: Fractional error in R2 = 2 x fractional error in R = 12x10-3. Error in R2 = σ(R2)
= R2 × fractional error in R2 = (10.45)2 × 12×10-3 = 1.3 (kΩ)2. (N.B. This is not (σ(R))2 nor is
it (2 × σ(R)2)).
Similarly, error in X2 = σ(X2) = X2 × 2 × fractional error in X = 507.6 × 2 × 1.8×103 =
1.8(kΩ)2.
Using rule 3: error in Z2 = σ(Z2) = √(σ(R2))2 + (σ(X2))2 = √(1.32 + 1.82) = 2.2 (kΩ)2.
Error in Z = Z × (1/2) × fractional error in Z2 (Rule 2 again) = 24.84 × (½) × (2.2/(24.84)2) =
0.05kΩ. Hence Z = (24.84±0.05) kΩ again.

Notes:
a) To combine errors in all cases, you must square, add and square-root. Just adding them is
wrong!
b) Distinguish carefully between the cases where the absolute errors are used and those
where fractional or % errors are used. To help you remember, Absolute errors for
Addition, Fractional errors for Fractions (Multiplication is a special fraction!!)
c) Note that x, y, z ... must be independent of each other.

27
3.6.4 Speed-up Hint
Calculations can be often considerably simplified and speeded up by noticing that some of
the contributions (especially when squared) are small enough to be ignored completely.

Example 10.
In example 9 we had to evaluate the following expression for σ(Z):
σ(Z) = √{[(10.45)2(0.06)2 + (22.53)2(0.04)2]/24.842} .
However (always on the look-out for short cuts), remember that σ(Z) is only required to one
significant figure, so it is pointless to use four figures in the calculation.
σ(Z) = 0.01√{[(10)2(6)2 + (23)2(4)2]/252} [rounding to 2 figs]
= 0.01√{[(100 × 36) + (500 × 16)]/600} [232 = 500 to 1 fig]
= 0.1√{[36 + 80]/25} = 0.1√{116/600} = 0.1 × √0.2 [116 = 120!]
= 0.045 kΩ. Call it 0.05 kΩ because all rounding has been down.
No calculator used! See how even quite crude approximations can be good enough.
Notice in example 9 that the error in Z is approximately equal to the error in X. Even though
the error in R (and the fractional error) is bigger, the contribution of σ(R) to the error in Z is
negligible. This is because X is twice as big as R, hence X2 + R2 ≅ X2, (Z is only 10% bigger
than X). One can then see intuitively that the error in Z must be due almost entirely to the
error in X, because R doesn't contribute very much to the value of Z, the fluctuations in the
last figure of R will change Z by a negligible amount. As you gain more experience, this kind
of argument will become instinctive.

28
3.7 Linear Relations Between Two Quantities
100

The relation between two quantities x and y, (either


directly measured or inferred from measurements) can
often be written as y = mx + c, where m and c are
constants. The aim of the experiment is to calculate m or 50

c or both. Clearly a graph of y against x should be a


straight line, with slope m and intercept c on the y-axis.
The first thing to do is to plot such a graph, draw the best
straight line that you can and get preliminary values for
m and c. If the values of x and y were exact, then all 0

points lie on a straight line, and finding m and c is easy. 1 2 3 4 5 6 7 8 9

If the points are somewhat scattered though, judging the


"best" line is not easy . A more objective method is to
make a least-squares fit to the data. 0

The best-fit line to a set of data can be established by


examining the residuals, di, given by di = yi -y, where yi
is the measured value for a particular x, and y is the value
calculated from the formula of the line. The least squares Figure 3.3 Straight line and second-
principle states that the sum of the squares of the order fits to a set of data. The residuals
residuals is a minimum for the best fit. Note that this are plotted below., the curve is those for
applies for any curve, not just a straight line (fig. 3.3) the straight line, the undulating line close
to the axis is that for the second-order
The usual formulae for least-squares fitting to a straight equation. Note that any other straight
line are derived on the assumption that experimental line would give bigger residuals.
errors in the quantity chosen for "x" are negligible
compared with those in "y". This situation is often realised in practice. Slightly more
complicated methods exist for use in cases where errors in the two quantities are comparable.
(Read the clear description in Mayer’s book – see refs).
You must have a preliminary look at the data and have convinced yourself that they are
reasonable, and that a straight-line fit is a sensible thing to do before trying a computer fit.
Several different software packages include both linear and non-linear fits.
Note that some calculators have the facility for calculating the "regression line" of y on x.
This is the same as the least-squares procedure, but unfortunately, the calculators do not give
any indication as to the errors in the gradient or intercept.

29
Example 11
Figures 3.4 and 3.5 show the results of fitting a straight line to very similar sets of data.
However, the residuals show that although the data of fig. 3.4 is fitted reasonably well by a
straight line, that of fig. 3.5 needs a second order term to fit i.e. an expression of the form,
y = ax2 + bx + c .

30
30

20
20

10

10

1 3 5 7 9

1 3 5 7 9

Figure 3.4 Data which the residuals


show is fitted reasonably well with a Figure 3.5 Data which the residuals
straight line show the need of a term in x2 to fit .

3.8 Recording Error Calculations


It makes it much easier for you (and the marker!) to see mistakes in your error calculations if
you bring them together at the end of the experiment in something like the following form :
[Quote Formula]
2x 3 y
G = 1
where m is the slope of the graph
2
z m
[Quote Error Formula]

σ ( y) 
2
 3σ ( x )  σ ( z)   σ ( m) 
2 2 2
σ (G )
=   +  y  +  2z  +  m 
G  x       

[Put in numbers previously calculated]


σ(G) = G√{(3×0.05)2 + (0.5)2 + (1.2/2)2 + 22} = 2.2%
[Quote final result in a standard form]
G = (100±2) S.I. units.

30
Note: It's very easy to make numerical mistakes in these calculations. Be very suspicious of
your working if you appear by mathematical methods to have multiplied or reduced
your error by a factor of 50! In the above case, the dominant error in the result was
due to the slope of the graph, and this comes through the calculation not much altered.

3.9 The Significance Of Discrepancies


Suppose we have measured a quantity and come up with the value x ± σ(x), and someone else
has measured what they believe to the same quantity (or calculated it from theory) and come
out with the value y ± σ(y). Of course, the value of x will not generally be the same as the
value y; how do we decide whether the discrepancy is important, or just due to random
errors?
Calculate the error in (x-y) by the usual rules of calculating errors; √{(σ(x))2 + (σ(y))2}.
Compare the discrepancy (x-y) with the error. If the discrepancy is several times larger than
the error, it is "significant" because it is unlikely to have arisen by chance, and someone,
somewhere has made a mistake either in experiment or theory. If the discrepancy is
comparable with the error, the two results should be regarded as consistent, with the
difference only due to random errors. A useful rule of thumb is that the difference is
significant if it is bigger than three times the error in the difference.

Example 12
A data book value of the Young modulus for brass is 100 GPa, with no indication of the error
(data books are very bad at telling one just how good the data are!). You may assume that the
error is the resolution (±1). Your measurements give (104.3±0.5) GPa. The difference is
(4.3±1.1) GPa, which is just big enough to be significant, and you need to look for reasons
why there should be a systematic difference. (The first thing to do is to check your work for
mistakes. The second thing to do is to check the book for mistakes.)

3.10 Counting Random Events


Many processes, like the decay of a radioactive nucleus or the counting of X-ray photons,
occur at random. This means that if a repeat measurement of a count is made, even under
identical conditions, there will inevitably be a random fluctuation in the recorded counts.
Events like these follow Poisson statistics, for which there is a particularly simple relation
between the count and the standard error of the results. If N counts occur in a time interval t,
then the standard deviation in the counts is √N. The counting rate is

N N
± . (3.8)
t t

N N
Note that the error in the counting rate is , not .
t t

Exercise 10
A student makes 10 readings of the counts in 20 seconds from a γ-ray source. The count rate
is calculated, and tabulated below. Complete the table with headings, and give the mean,
population standard deviation, standard error of the mean and the final result in standard
form.

31
Count rate x[s-1 ] Deviation Deviation2
3.4
2.9
3.6
3.1
2.6
3.4
3.4
2.2
3.5
3.1

Mean .........................= .........................


Population Standard Deviation .........................= .........................
Standard Error of Mean .........................= .........................
Count rate = .........................± .........................
Since the counts are random, the student decides to check that the error in the count rate is in
fact given by equation (3.8)
Total number of counts = .......................................................................
Total time = .......................................................................
Count rate = .......................................................................
Error in count rate = .......................................................................
Count rate = .................... ± ....................

3.11 Goodness Of Fit, or More About Least-Squares


In this section and the next, we consider some more advanced statistical ideas.
Suppose we have some data points (x,y), which we have fitted to a straight line. The question
then arises: does the data really fit a straight line? Deviations from a straight line occur for
two reasons:
a) random deviations because of experimental error.
b) systematic deviations because of some effect ignored in the theory (e.g. the breakdown of
the approximation sinθ = θ in the theory of the pendulum) or the experiment (e.g.
calibration error).
These can be distinguished by studying the residuals
di = yi - mxi - c , (3.9)
where di is the deviation or residual for the ith point, xi is the observed value of x, yi the
corresponding value of y, and m, c the slope and intercept found by the fitting procedure.

32
If the deviations are random, then di will have random variations in sign and magnitude. If,
however, di varies systematically (e.g. from large negative values through to large positive
values and back) then the data points lie on a curve, not a straight line. Sometimes you can
see this from the graph - that is why it is always so important to plot one!
A more objective measure of the goodness of fit is given by χ2 (chi-squared, χ pronounced
"chi", where the ‘ch’ has the sound found in a Scottish loch. English people who cannot
produce the correct sound usually say ‘ki’.) In order to use χ2 , it is necessary to have some
idea of the typical error in the y-values, for example by repeating readings for a particular x
and calculating the population standard deviation s of the y's. Then

χ 2
=
∑d i
2

. (3.10)
s2
In words, χ2 is the ratio of summed squared deviations to the squared error in a single reading
(note the sum sign in the numerator only!). Clearly, the more points you take, the bigger χ2
will be. The least squares method in fact minimises χ2.
For a good fit to a straight line, χ2 should be about N-2, where N is the number of data points.
The ratio of χ2/(N-2) is called the "goodness of fit parameter". The following rule-of-thumb
gives a measure of how good a straight line fit is.
If χ2/(N-2) ≈ 1, the fit is probably satisfactory. Very roughly, we can interpret
χ2
≈ 1 as indicating that on average each point lies a
N −2
distance from the fitted line of about one standard deviation s.
Therefore overall the deviations are consistent with the fitted
shape being correct.
If χ2/(N-2) > 1.5 - 2, the fit is probably poor; either the accuracy of the points is
less than you estimated (s2 is too small) or a straight line is not
the appropriate fit. You have to decide which is the case by
inspecting your results closely.
If χ2/(N-2) < 0.5 the fit is probably too good! Your estimate of s2 is too high,
and the data may be better than you think. Again you need to
think carefully about your results. Have you overestimated
your errors?
Note the use of the word probably in the rule. Sometimes, by chance, perfectly good data
may give a χ2/(N-2) > 1.5. The χ2 rule is not infallible, but must be regarded as a guide.

Example 13.
A mass M is hung on a spring, and the extension e is measured. The reading error in e is
about 1 mm and so the error in e is taken as ±0.5 mm. The error in M is negligible. The
following results were obtained:

33
M[g] e[mm] ecalc[mm] e - ecalc[mm] (e - ecalc)2[mm]2
10 5 5.4 -0.4 0.16.
12 7 6.3 +0.7 0.49
14 7 7.2 -0.2 0.04
16 8 8.1 -0.1 0.01
Total 0.70

The least-squares fit to the data in columns 1 ("x") and 2 ("y") gives a slope of (0.45±0.11)
mm gm-1 and an intercept of (0.9±1.5) mm. The best-fit straight line is e = 0.45 M + 0.9.
From this we calculate the expected value of e for each value of M and then find the
differences, (columns 3 and 4). The sum of the squares of the differences is then found
(column 5), and χ2 calculated:
χ2 = 0.7/(0.5)2 = 2.8 χ2/(N-2) = 2.8/2 = 1.4
This is less than 1.5, so the straight line fit is probably satisfactory. (Actually, with so few
points, a satisfactory fit could well give χ2/(N-2) greater than 1.5. The critical value of 1.5 is
"typical" for graphs with 10-20 points on them).

3.12 Covariance and Correlated Errors


Earlier in these notes we quoted a relation (equation (3.7)) between the error in variables x, y,
z, ... and the resulting error in the function F(x,y,z, ...). If there are two variables x and y, then
the relation may be written in the form
2 2
 ∂F   ∂f 
σ(F) =   σ( x) +   σ( y) .
2 2 2

 ∂x   ∂y 
This equation is valid provided x and y are uncorrelated; that is, values obtained for y are
independent of the x value. For most measurements this is true, but there are times when this
is not the case. An important example of this is when straight line fits to data are obtained.
Here the value of y obviously depends on the value of x chosen, and so when we calculate the
slope and the intercept, the intercept value will be to some extent related to the slope. If the
slope is too high, the intercept may be too low, for example:

34
Example 13.
To show the kind of problems that this can cause, consider the data of example 13. Suppose
we wish to calculate the extension for a mass of 13g. We would expect the error in the
calculated value to be rather less than 0.5 mm, since the error in any one e is about this, we
are considering some sort of "average" of the y's, and the x value chosen is in the middle of

Figure 3.6 Graph of e against M, showing point for M = 13 g. The error is clearly an overestimate!

Fit To Data Plot of Residuals of Fit


10

9 2

Residuals
e

7
10 12 14 16

5 2

4
10 12 14 16 18
M
M

the bunch. (If we are far away, then we would expect larger errors, of course. Look at the
error on the intercept, for example.) Using e = mM + c = 0.45M + 0.9, we get e = 6.75 mm.
Now use the above equation to estimate the error in e caused by the errors in slope m and
intercept c.
2 2
 ∂e   ∂e 
σ(e) =   σ( m) +   σ( c)
2 2 2

 ∂m   ∂c 
= M2σ(m)2 + σ(c)2 = (13)2 × (0.11)2 + (1.5)2 = 4.29 mm2
σ(e) = ±2.5 mm.
This value for the error is obviously far too big, the range is bigger than the whole range of
e's covered (fig. 3.6).
The mistake is not to allow for the fact that errors in m and c are correlated. The correct
expression to use in this case is:

2 2
 ∂F  ∂F ∂F  ∂F 
σ(F) =   σ( x) + 2 σ( x , y) +   σ( y) . (3.11)
2 2 2

 ∂x  ∂x ∂y  ∂y 

35
where σ(x,y) is called the covariance of x and y (often shortened to cov(x,y)). σ(x)2 is called
the variance of x, because it comes from the variation in the values of x. Similarly σ(y)2 is
the variance of y. The extra term allows for the variation in x (or y) due to corresponding
variations in y (or x); it measures how the variations in x go with (hence the prefix co-"with")
the variations in y, i.e. how x is correlated with y. If these variations are uncorrelated, then
σ(x,y) is zero.

Figure 3.7 Graph of e against M, showing point for M = 13 g. The error bar calculated taking account
of the covariance is much smaller than in fig 3.6.

Fit To Data
9 Plot of Residuals of Fit
1

7
Residuals

10 12 14 16
e

5 1

4
10 12 14 16 18
M
M

Example 14.
For the data in example 13, the covariance is not zero, but equals -0.163 (the minus sign
shows that the slope and intercept in this case are anti-correlated, i.e. if the slope is too low,
the intercept is too high and vice versa. A positive sign would show that either both were too
high or both were too low). Using the modified expression,(3.10), we find that
2 2
 ∂e  ∂e ∂e  ∂e 
σ(e)2 =   σ( m) + 2 σ( m, c) +   σ( c)
2 2

 ∂m  ∂ m ∂c  ∂c 
= M2σ(m)2 + 2Mσ(m,c) + σ(c)2
= (13)2× (0.11)2 + 2×13×(-0.163)+(1.5)2 = 0.05 mm2
giving σ(e) = ±0.2 mm, a much more satisfactory error figure (fig 3.7). Many least-squares fit
programs will print out the elements of the so-called error matrix. The diagonal elements are
the variances in m and c, and the off-diagonal elements (which are equal - the matrix is
symmetrical about the diagonal) give the covariance between m and c.

3.13 Further Reading


The recommended text for the Experimental Statistics course is
Lyons L. Data Analysis for Physical Science Students (Cambridge University Press)

36
Other books on this subject include:
Barford, N.C. Experimental Measurements, Precision, Error and Truth (2nd edition) (John
Wiley and Sons, London).
Barlow, R.J. Statistics (Addison Wesley)
Bevington, P.R. Data Reduction and Error Analysis for the Physical Sciences (McGraw Hill,
New York).
Mayer, S.L. Data Analysis for Scientists and Engineers.
Squiers, G.L. Practical Physics (3rd edition) (Cambridge University Press).
Taylor, John R. An Introduction to Error Analysis (University Science Books)
Topping, J. Errors of Observation and their Treatment (Chapman and Hall)

37
38
CHAPTER 4 - PRINCIPLES AND OPERATION OF COMMON
INSTRUMENTS
4.1 Oscilloscopes
We have several different types of oscilloscope in the laboratory, all of which work on the
same general principles, but which differ in details of construction and in the names
identifying their controls. Section 4.2 of these notes describes the operation and main
facilities which are common to all; section 4.3 gives specific instructions for setting up the
individual models. The instructions are very brief, their aim being simply to teach you the
proper routine for obtaining suitable working conditions. For further explanation and
information you should consult a demonstrator or read the appropriate pamphlet in the
library.

4.2 Introduction to Oscilloscopes


A cathode ray oscilloscope (CRO) is used mainly to display fluctuating potential differences
("wave-forms"), as functions of time. It can therefore be thought of as a form of recording
voltmeter and compared with an instrument consisting of a moving coil voltmeter in which
the pointer is attached to a pen that makes a trace on a moving chart (pen recorder). There
are however several important differences between these two instruments.
a) The moving coil meter would take a time of at least about a tenth of a second to respond to
a change in voltage, whereas the oscilloscope can respond in times considerably less than
a microsecond.
b) The input to the oscilloscope is through an amplifier which draws very much less current
than a moving coil meter, although it is certainly possible to design a pen recorder with an
amplifier input. In oscilloscopes the input impedance is commonly about l MΩ in parallel
with 30 pF.
c) The screen on which the oscilloscope trace is displayed is of limited length, in contrast to
the effectively infinite length of the paper chart in a pen recorder. The use of an
oscilloscope is therefore usually restricted to the study of signals for which the waveform
can be drawn on the screen over and over again, each trace being superimposed on the
previous one. In order to achieve this, all repetitions of the trace must be made to start at
equivalent points on the input signal. This is done by using the input signal itself (or one
related to it) to "trigger" the time base, i.e. to make the trace start to move across the
screen at the desired instant.
d) Oscilloscopes are frequently equipped with a
wide range of additional facilities which make Signal to
them very versatile. For example, two Input Y-plates
Y-amp C.R.T
different signals may be applied to the X and
Y inputs, giving Lissajous figures, and the
brightness of the trace may be modulated by Ramp to
X-plates
yet a third signal if desired. Trigger
Fig 4.1 shows a block diagram of an Pulse
Trigger Time-base
oscilloscope. The various functions are Generator
discussed below.
Figure 4.1 Functional block diagram of CRO

39
4.2.1 The Cathode-ray Tube
The cathode ray tube (CRT) X-plates
consists of three main parts Y-plates
contained in an evacuated glass
G
tube as shown in fig. 4.2. The
electron gun G contains a heated Brilliance and
cathode which emits electrons Focus Electron Beam S
that are accelerated and focused
by a series of anodes to form a Figure 4.2 Schematic view of CRT
narrow beam. This beam passes
between two pairs of deflector plates onto the fluorescent screen S where it produces a spot of
light. When potential differences are applied between the X pair of deflector plates, the point
where the beam strikes the screen is deflected horizontally by an amount proportional to the
applied p.d.. Applying a p.d. to the Y-plates gives a similar vertical deflection. The decay
time of the phosphorescence plus persistence of vision means that if the motion is fast enough
the spot of light appears to become a line called the trace.

4.2.2. The Y-inputs


The input to the CRO is usually via a BNC type coaxial socket (see section 4.2.6). The signal
then enters an input amplifier of variable gain, the output of which is fed to the Y-deflector
plates. The amplifier is calibrated so that, for each gain setting, the vertical deflection on the
screen corresponds to a known voltage at the input. The accuracy with which this voltage
can be determined is limited by the stability of the amplifiers and by the precision with which
readings can be taken from the screen. The absolute accuracy of most oscilloscopes is only
about 5%, unless careful calibration procedures are adopted at the time of measurement.
Note that most Y-amplifiers have an additional variable gain control which is uncalibrated. If
you need accurate voltage measurements, then this control must be set to "CAL".
A switch near the Y input enables A.C. A.C.
you to choose either A.C. or D.C.
operation and, on some models, to
switch off one of the channels or to Input Gnd Gnd ToCRO
disconnect the signal and earth the
amplifier input. In the D.C. position
as shown in fig. 4.3, the input signal
is fed directly to the Y input
amplifier; in the A.C. position it is D.C. D.C.
fed via a capacitor. Use of the A.C.
position serves to block off any D.C. Figure 4.3 Input circuit of CRO. The two switches are
component of the signal, but will ganged, so that the different connections are made
simultaneously.
cause distortion and attenuation of
low-frequency signals (typically < 5 Hz). In the GND (Ground) position, the external circuit
is disconnected and the input to the Y amplifier is connected to earth (0V). This enables you
to zero the trace.
A dual trace tube enables two signals to be displayed simultaneously. Electronic switching is
used to feed the Y1 (or channel 1) and Y2 (channel 2) signals from their respective amplifiers
alternately to the Y-plates. Persistence of vision means that the switching action cannot be
seen, and it looks as if the two signals are present simultaneously. The oscilloscopes in the

40
lab are dual trace. (In dual beam oscilloscopes the two traces are achieved by providing two
entirely separate electron guns and deflection plates: these are genuinely simultaneous.)

4.2.3. The Time-Base


B
The fluorescent spot is made to
travel horizontally across the screen
at constant speed by means of the
internal time-base circuit. This A C A’
supplies to the X-plates a potential
difference that rises linearly with Figure 4.4 Time-base waveform (ramp)
time and then falls suddenly to its
initial value, as shown in fig. 4.4. During the interval AB the spot moves uniformly from left
to right and during BC it returns to its starting point, where it remains until the next time-base
cycle starts at A'. To avoid seeing the confusing "fly-back" trace, the electron beam is
blanked off for the interval BCA'. The rate at which the spot moves across the screen can be
selected by one of the front-panel controls.
Many oscilloscopes possess an additional variable time base control which is uncalibrated. If
you need accurate timing measurements, then this control must be set to "CAL".

4.2.4. Triggering the Time-base


Usually, if you switch an oscilloscope on and
feed a signal into one or both of the Y inputs,
you will not obtain a steady trace. This is due
to the time-base "free-running"; as soon as one
sweep has been completed, another will start of
its own accord without any reference to the
phase of the input signal (fig. 4.5). If, however,
the time-base is adjusted so that it will start only
when it receives a triggering pulse from one of Figure 4.5 Free running timebase. Note the
the Y-inputs, the time-base and input signal will jumbled signal as seen on the screen.
keep in step; the time-base is said to "lock-on to
the input" (fig. 4.6).
Achieving this condition is an important part
of the setting-up procedure; practical details,
which differ according to the type of
oscilloscope that you are using, are given in
section 4.3. With all oscilloscopes you can
select either positive or negative trigger pulses
and choose between "automatic" and "normal"
(adjustable level) triggering; these terms are Figure 4.6 Triggered time base. Similar parts
explained in section 4.3. It is possible also to of the waveform are continually displayed,
trigger the time-base from an external pulse giving a steady trace.
which usually has a fixed time relation to the
input signal. Many oscillators generate a fixed-amplitude signal for this purpose, which is
very convenient to use when the amplitude of the Y input is too small to act as a reliable
trigger.

41
4.2.5 X-Y Display
All CRO's have a facility where an external signal may be applied to the X-plates instead of
the time base. This can be useful when studying the response of a circuit to an imposed
signal, particularly if phase information is required (read up Lissajous' figures to see how this
is achieved). Often the X-signal is applied using one of the Y-amplifiers, which is then
connected to the X-plates by a switch.

4.2.6 Co-Axial Cable Outer Polythene


Cover Insulator
Co-axial cable ("co-ax") consists of an inner
conductor separated from the outer copper braid
by polythene insulation, the whole unit being
enclosed in a black insulating sleeve (see figure
4.7). At least one end is terminated by a metal
connector (BNC type). The outer part of this Earth Inner
Braiding Conductor
connector is joined to the copper braid and the
fine central contact is soldered to the inner
Figure 4.7 Co-axial cable
conductor of the co-ax cable.
We have fitted the other end of the lead with 4mm banana plugs, spade terminals or hook
terminals; these are colour-coded; the red plug is soldered to the inner conductor while the
other (either black or green) is connected to the braid.
The cables are normally used with oscilloscopes, most of which have BNC type sockets. The
centre pin of the BNC socket leads to the Y-input circuit, while the outer is connected to the
CRO ground (0V), which is mains earth. This is a big drawback of CRO's; all voltages
displayed are automatically measured with respect to ground. When you plug the cable into
the CRO socket, the braid and the corresponding terminal are automatically earthed via the
oscilloscope housing and mains lead. Make sure that no other point on your circuit is
earthed. If you are using both channels of the oscilloscope, the outer of both cables must go
to the same point of the circuit. Remember, there is only one Earth!

4.2.7 High-impedance Probes


For some purposes the input impedance of the oscilloscope is too low, so that it draws
excessive current from the circuit it is being used to study. The high impedance probe can be
used to raise the input impedance from 1MΩ in parallel with 30pF to 10MΩ in parallel with
8pF. When this probe is in use the voltage sensitivity of the input is reduced by a factor of 10,
i.e. all the gain settings have to be multiplied by 10 to obtain the true voltage. The crocodile
clip at the side of the probe is an earth connection to the circuit, but it is not essential to use
this at low frequencies if the circuit is already connected to the oscilloscope earth.

4.2.8 Digital Storage Oscilloscopes


The CRO as described above is an "analogue real-time device", displaying a varying signal as
it is received. As soon as the signal is past, it is lost. It is useful to have a way of recording
signals for later analysis, particularly if it is a "one-off" transient, like a spark. One way is to
photograph the screen of a CRO, the other is to use a storage scope. The storage or digital
scope is a conventional CRO as far as the input amplifier and trigger controls are concerned.
However, instead of sending the signal direct to the screen, the signal is digitised by an
analogue-to-digital converter (ADC), and is then stored in an electronic memory like a
computer (Indeed, a computer fitted with suitable input circuits can be used as a storage
scope.) The memory is read continually, and a digital-to-analogue converter (DAC) produces
the voltages needed for the CRT display.

42
A storage scope can be used in two ways; either continual sweep, when it behaves just like an
analogue CRO, or in "single-shot" mode, where a single sweep of the timebase is stored, and
then replayed repeatedly for study, either visually on the screen or by outputting to another
device such as an X-Y recorder or a computer. The continual sweep is used for setting up,
and then the storage scope is set for single shot, and the waveform captured. If rare events
are to be studied, then the trigger level is set appropriately, and the CRO can then be left to
"baby-sit"! It is possible to store that part of the signal occurring just before the trigger, so
the build-up to the event can be studied. The price for all this clever electronics is that
storage scopes tend to be rather slow, and cannot be as fast as analogue CRO’s. They are
also much more expensive!

4.3 Operation of a CRO


All modern CRO’s have the following basic controls, though the front panel layout will
differ. Controls are usually grouped together as follows.

1. Trace controls:
Brilliance or Intensity (often combined with mains power on/off switch): controls
brightness of trace
Focus: Controls sharpness of trace

2. Channel 1 or Y1
Input: Usually co-axial socket. Signal is on the inner pin, the outer is grounded. See
the note on Co-axial Cable.
Volts/cm: Switched calibrated fixed gain control.
Gain or Var. Sens (Variable sensitivity): Variable gain control. Must be set to CAL
(calibrated) position for correct voltage readings. (Not always present.)
↕ or Position: Sets vertical position on screen.

3. Channel 2 or Y2
Controls for channel 1 are duplicated. The following extra controls are often present.
CH2 invert: All plus signals become minus and vice-versa.
CH1 - CH2: the trace displayed is the difference between the channel 1 and channel
2 signals. (Useful if you want to see distortions or to measure a p.d. between two
points, neither of which are at ground potential.)

4. Time-base
Time/cm: Switched calibrated fixed settings.
Var. Sweep or Variable: Variable adjustment of time base rate. Must be set to CAL
position if accurate timings required.
↔or Position: sets horizontal position on screen.

5. Triggering Controls
Trig. Level or Trigger Level: Sets exact position on input waveform which starts the
time-base.

43
Trigger or Trigger Select: Array of switches which select source of trigger pulse.
The important ones in initial setting up are whether the trigger is derived from channel
1, channel 2 or an external source, and the selection of automatic or manual
adjustment. For details see individual scope descriptions.
+/-: Decides whether the trigger is on a upward (+) slope or rising edge of the signal,
or on a downward (-) slope or falling edge. Most useful when dealing with pulses,
which can be +ve or -ve going.

6. Other controls
The following features are present in all CRO's, though manufacturers differ in how
they are presented. For details, see the individual CRO descriptions.
Channel selection: Ch. 1 and/or Ch. 2 may be switched off or on.
X-Y: The control selecting this feature is positioned in many differing places!
CAL: A square wave output signal useful for checking the working of the CRO.
Finally, four DO NOT's to remember!
DO NOT keep switching the mains on and off.
DO NOT use an unnecessarily bright trace, and never leave a bright spot stationary
on the screen. (It will burn off the phosphor).
DO NOT adjust the screwdriver controls without consulting a demonstrator.
DO NOT bring a magnet near a CRO. (It magnetises the structure and permanently
distorts the trace).

4.3.1 Gould Oscilloscopes OS255/OS300 (no longer in general use in the year1 lab but
the description of its controls may still be applicable to other types of oscilloscope).
Plug the oscilloscope into the mains, but do not yet connect it to any other circuit. Switch on
by turning the INTENSITY control slightly clockwise, just far enough to operate the click
switch. Check that the indicator lamp lights. Give the oscilloscope a few minutes to warm-
up while setting other controls as follows:-
FOCUS Mid-position
↕ and ↔ Mid-position
CH1 and CH2 (Slide-switches) Gnd.
Var Sens and Var Sweep Cal. (fully clockwise)
BRIGHT-LINE ON (button out)
VOLTS/CM lV
TIME/CM l ms
MODE
Dual (both channels on).
Two horizontal traces should appear on the screen as the INTENSITY control is advanced.
Adjust FOCUS to obtain sharply defined traces. Adjust X-shift ( ↔) so that the beginning of
the traces is on the screen. If you cannot see the traces even with the intensity control turned
fully clockwise, try rotating the ↕ knobs slowly: it is possible the traces maybe off the top or

44
bottom of the screen. If you still cannot obtain traces, turn the intensity down, do not turn
off, and consult a demonstrator.
A suitable signal to investigate in the first instance is the nominal 1 kHz square wave
obtainable from the CAL output. Set the slide switches controlling CH1 to A.C. Plug a
coaxial cable into the input, and connect the inner to the CAL output. The wave form should
now appear, though it may not be stable. Trigger from CH1 by pushing in the CH1 button in
the "Trigger" group of controls. With a simple A.C. input, and internal trigger, the position
of the +/- button is irrelevant; it is important only if you have a unipolar (pulse) triggering
signal. To obtain a stable trace, rotate the TRIG LEVEL control. Experiment with the
TIME/CM, VOLTS/CM, and the Var Sens and Var Sweep controls. Repeat for CH2 .
The Bright-line control corresponds to the automatic trigger facility. When the button is out,
the time base will free-run in the absence of an input; this helps in finding the trace. If the
input signal has a frequency greater than 40 Hz, it will take control of the triggering when the
TRIG LEVEL control is appropriately set, and a stable trace will result. If, however, the
frequency of the input is less than 40 Hz it may be difficult, or impossible, to lock the signal
to the time-base. In this situation the "BRIGHT-LINE OFF" mode must be selected. The
trace then will only appear when the TRIG LEVEL control is correctly set. Provided that the
amplitude of the signal is sufficient to give a deflection of about 0.5 cm or more on the
screen, a stable display should result; and, by turning the rotary switch of the TRIG LEVEL
control, the display can be triggered from any point on the waveform.
Other controls featured on this CRO are;
X-Y
This is selected by the MODE switch. In this position, CH1 is Y, CH2 is X.
Inv CH2
If the Inv CH2 button is in, the channel 2 signal is inverted.
Add
When the MODE switch is set to Add, the signals from CH1 and CH2 are added. If the Inv
CH2 button is in, the CH2 signal is inverted, and then CH2 signal is subtracted from the CH1
signal. This feature is rarely used, but can be useful as a way of reducing pickup.
External Trigger
A separate trigger source fed into the Ext Trig socket is selected when both CH1 and CH2
Trig. buttons are in.
TV/AC/DC
Normally the AC button is depressed. To get reliable triggering from low frequency signals,
the DC button is depressed. If both buttons are depressed, then conditions are correct for
triggering off TV line synchronisation pulses (we rarely use this feature).
Mag.
When in, the time base scale is magnified 5 (or 10) times giving a maximum sensitivity of
0.1µs/cm. If you need to study signals which are this fast, then using a faster CRO is strongly
recommended.

SOCKETS AT REAR OF INSTRUMENT


Ext Intens Input

45
A signal in the range from +2 to +30 V (d.c. + a.c. peak) will decrease the brilliance of the
trace. Thus the intensity may be externally controlled or modulated.

4.3.2 Hitachi Oscilloscope V-212 (no longer in general use in the year1 lab but the
description of its controls may still be applicable to other types of oscilloscope).
Plug the oscilloscope into the mains, but do not yet connect it to any other circuit. Turn
INTENSITY knob full counter-clockwise, and then switch on by pushing the POWER button
A red pilot light should come on. Give the instrument a few minutes to warm up, setting the
other controls as follows:
FOCUS Mid-position
POSITION Mid-position *
POSITION Mid-position
VOLTS/DIV 0.1 volt
VAR (co-axial with VOLTS/DIV) Full clockwise (CAL) *
AC/GND/DC GND
(input) MODE (rotary switch) CH1
INT TRIG CH1
TIME/DIV. 0.5 ms
SWP/VAR Full clockwise (CAL) *
SOURCE INT
LEVEL Mid-position *
(trigger) MODE (lever switch) AUTO
Make sure that all knobs marked with a * in the above list are depressed. Then turn the
INTENSITY knob clockwise until the trace is visible. Adjust FOCUS until a sharp line is
obtained. With the trigger MODE switch set to AUTO and no signal the timebase is free
running, i.e. triggering is produced by pulses produced internally. If you do not see a trace on
the screen, adjust the POSITION knob. If you cannot see the trace even with the intensity
control turned fully clockwise, try rotating the POSITION knobs slowly: it is possible the
trace may be off the top or bottom of the screen. If you still cannot obtain a trace, turn the
intensity down, do not turn off, and consult a demonstrator.
Adjust focus and intensity and if necessary rotate POSITION ↔until the beginning of the
sweep is visible near the left-hand side of the screen. Set CH1 switches to AC. Connect CH1
input to the CAL 0.5 tip-terminal or to an A.C. signal from an external circuit; adjust the time
base speed (TIME/DIV) and VOLTS/DIV switch if necessary. Provided that the amplitude
of the signal is sufficient to give a deflection of about 0.5 cm or more on the screen, a stable
display should result. By turning the knob of the LEVEL control, this display can be
triggered from any point on the waveform. If the input signal is too small to actuate the time-
base, or should the signal be "lost", a free-running reference trace will appear. Normally, a
stable trace is given by leaving the trigger MODE switch set to AUTO, but for low frequency
inputs (i.e. less than 25 Hz), you will find it necessary to set the trigger MODE to NORM.
TV is rarely used.
To observe signals on CH2 alone, set the input MODE rotary switch and the INT TRIG
switch to CH2.

46
Other controls featured on this CRO are
Dual Beam Switching
The V-212 oscilloscope achieves its double beam by electronic switching; that is, the CH1
and the CH2 signals are fed alternately into the Y output amplifier. The switching mode is
controlled by the CHOP and ALT positions on the input MODE switch. In the ALT position,
the CH1 and CH2 signal are displayed alternately on successive sweeps of the time base.
This setting is preferred for dual trace work at high sweep speeds. In the CHOP position, the
CH1 and CH2 signals are switched to the Y output amplifier at a frequency of 250 kHz, i.e.
many times within a single sweep of the time base. This setting is preferred for dual trace
work at slow sweep speeds.
Vert. Mode Switching
Triggering may be from either CH1 or CH2 as selected by the INT TRIG switch.
Occasionally, you may wish to observe signals of two different frequencies from different
sources. Normally, only one display will be properly triggered, but by setting the INT TRIG
lever switch to VERT MODE, triggering signals are taken alternately from CH1 and CH2, so
both displays are stable. This only works when the input MODE switch is set to ALT.
X-Y
This is selected by setting the TIME/DIV switch to X-Y. In this position, CH1=X, CH2=Y.
PULL INVERT
If the CH2 POSITION knob is pulled out, then the CH2 signal is inverted.
ADD
If the input MODE switch is set to ADD, the signals from CH1 and CH2 are added. If the
CH2 POSITION knob is pulled out as well, then the CH2 signal is subtracted from CH1.
This feature is rarely used, but can be useful as a way of reducing pickup.
PULL x 5 GAIN
If the VAR knobs (small ones co-axial with the VOLTS/DIV switches) are pulled out, an
extra gain of approximately x5 is given to the appropriate channel, to give a maximum
sensitivity of lmV/cm.
LEVEL PULL (-) SLOPE
The rotary TRIGGER LEVEL control is pushed in for triggering from rising edges (+ve-
going) of signals. It is pulled out for falling or negative-going edges.

PULL x10 MAG


When the POSITION knob is pulled out, the time base scale is magnified by approximately
x10, giving a maximum sensitivity of 0.02 ms/cm. If you need to study signals that are this
fast, then using a faster scope is strongly recommended.
SOURCE
When set to INT, triggering is from the signal applied to CH1 or CH2, as selected. When set
to LINE, triggering is from the mains supply (this is useful when looking for mains pickup or

47
mains related noise, e.g. from a fluorescent lamp). When set to EXT, a separate triggering
source is used, supplied via the TRIG IN coaxial socket.

SOCKETS ON REAR OF INSTRUMENT


EXT BLANKING INPUT
A signal +5 to +30 V (dc + ac peak) will decrease the brightness of the display. Thus the
trace brightness may be controlled or modulated from an external source.
CH1 OUTPUT
The CH1 signal is available here (amplified if necessary).

4.4 Digital Meters


In electrical circuit measurements, you should always remember that voltmeters draw a small
current and that there is necessarily a voltage drop across ammeters. Thus, when using such
instruments, you should always try to find out how their properties will influence your results
and connect them so that the effects are minimised. Most electrical measurements in this lab
use general-purpose digital multimeters (DMM). These notes summarise the functions and
use of the three types that are available.

4.4.1 Principles of DMM Action


DMM’s of the type in general use are dc voltmeters. For voltage measurements they can be
considered to be an ideal infinite-impedance voltmeter in parallel with a large value resistor.
The current drawn by this resistor is the current leakage through the voltmeter. Generally the
resistance is 1 to 10 MΩ. The voltmeter consists of an amplifier followed by a analogue-to
digital converter (ADC). The output of the ADC is displayed on a digital display. The
various voltage ranges are obtained by varying the gain of the amplifier.
AC voltages are measured by rectifying the signal, and measuring the mean direct voltage
with the ADC as above. The amplifier gain is adjusted so that the displayed voltage is 1.110
times the mean voltage, where 1.110 is the ratio of the RMS voltage to the mean voltage for a
sinusoidal waveform. The meter is said to be "mean sensing, RMS calibrated". You should
note that this conversion factor is only valid for sinusoidal waves, and the reading is
misleading for other waveforms such as square, pulse or triangle. We have multimeters
which read true RMS values irrespective of waveform, these are available for project work if
required.
BEWARE! These meters will only record correctly over a limited range of frequencies.
They will give readings for frequencies outside this range, but these will be incorrect.
Direct and alternating currents are measured by passing them through a small resistance built
into the meter and measuring the voltage drop across the resistor. The amplifier gain is
usually kept constant, so the current corresponding to full scale reading is equal to a constant
voltage (called the voltage burden) divided by the value of the resistor. Scale changing is by
changing the resistor to one of a different value. A separate input for high currents is often
provided. A fuse is used to provide protection from current overload.
The meters measure external resistance by internally generating a pre-set current, which
passes through the resistor when connected, and measuring the potential drop across the
resistor. Scale changing is by changing the amplifier gain except for small resistance values
where the pre-set current is increased.

48
4.4.2 Absolute and Relative Accuracy
If you set the meter to read a stable current or voltage, its reading will only vary by ±1 in the
last digit at most. If it were reading 190.0 mA this would represent 1 part in 1900, which is
only 0.05%. This is the meter resolution. However, this is not the absolute accuracy of the
meter: if you replace it with another one or if you change ranges, it is likely that the reading
will alter by more than 1 digit; it could be as much as 15 digits. This because the absolute
accuracy of the instrument is determined by voltage and resistance standards built into its
circuitry, and these vary from meter to meter and range to range. In other words, the absolute
accuracy is a measure of how well the meter was calibrated by the manufacturer. The
absolute accuracy is usually defined as a percentage of full scale reading + the number of
digit reliability in the least significant digit (usually 1, sometimes more). Usually one of these
error figures is much greater than the others, so you will only need to quote one of them. If
they are comparable, then square, add and square root as usual.
The relative error when you are comparing readings on the same meter set on the same scale
are usually good to the resolution of ±1 digit, so it is good experimental technique to find a
method for comparison. The absolute error indicates the manufacturer's tolerance in setting
up a range of instruments, so if you calibrate the meter, then the error is limited by the
precision of the scale.

4.4.3 Use and Safety Precautions


a) Use extreme care when working near high voltage sources. This means any voltage
source greater than 200 V dc or RMS ac
b) Do not attempt to measure any voltage differences when the lower voltage is appreciably
above ground potential. For the Beckmann T100B and the Lascar LMM 100 the limit is 1
kV peak, for the Thurlby 1503 it is 500 V peak.
c) Never remove the test leads from the circuit if it connected to a live source of high voltage.
d) When using the current range, always make sure the meter is in series with the load
carrying the current to be measured. Never connect the multimeter across a voltage source
when in a current mode.
e) Never apply a voltage to the instrument that exceeds the maximum allowable voltage for
that range.
f) When making measurements on a circuit containing capacitors, always turn the power off
and make sure the capacitors are discharged fully before connecting or disconnecting the
meter.
g) Do not change range when measuring current in a circuit containing large inductors.
h) When measuring an unknown voltage or current, always set the meter to the highest range,
and reduce until a satisfactory reading is obtained.

4.4.4 Properties of Particular Meters


The following table summarises the main properties of the three meters we use commonly.
The tables give an indication only, and the manufacturer's instruction book must be consulted
for fuller details.
Beckmann Lascar Thurlby
T100B LMM 100 1503

49
Display (no of digits) 3½ 3½ 4 3 /4
Input sockets
Common/low potential COM LO LO
Voltage V-Ω HI HI
Resistance HI HI
Current (up to) A (2A) HI (2A) HI (0.8A)
High Current 10A 20A 10A
Frequency --- --- 3.5 mm jack
Function selection
Voltage button V button V
Resistance common button Ω button Ω
Current rotary button A V & Ω together
AC/DC select switch button button
Range Selection buttons buttons
Direct Voltage Measurement
Ranges 200.0 mV 200.0 mV 320.00 mV
- 1000 V - 1000 V - 1200.0 V
Resolution (+ LSD) 1 1 1
Absolute Accuracy
% reading 0.5% 0.1% < 0.1%
+ no. of digits 1 (all ranges) 1 (all ranges) 6 (320 mV)
1 (1.2 kV)
2 (others)
Input impedance 10 MΩ 10 MΩ 10 MΩ

Overload (max. dc ac peak) 1 kV 1 kV 370V/1.2 kV *

Direct Current Measurement


Ranges 200.0 mA 200.0 mA 80.00 mA
- 10.00 A - 2.000 A - 10.00 A
Resolution (± LSD) 1 1 1
Absolute Accuracy:;
% reading 1% 0.25% < 0.3%
(2% (2% 2 A) (2% 10 A)
2 & 10 A)

50
+ no. of digits 1 1 2
3(2 & 10A) 1(10 A)
4 (80 µA)
Voltage burden 250 mV ** 800 mV
(750 mV 2 & 10A) (1.2V 0.8A)
(150mV 10A)
Overload (max. dc+ac peak) Fused up to 2A Fused 2 A Fused up to 2A
only only
Alternating Voltage Measurement
Ranges 200.0 mV 200.0 mV 3200.0 mV
- 750 V - 1000 V - 750.0 V
Resolution (± LSD) 1 1 1
Absolute Accuracy:
% reading 1.25% 0.75% 0.5% (2% 750
V)
+ no. of digits 4 5 15
Input impedance 10MΩ//100µF 10MΩ//100µF 10MΩ//40µF
Overload (max. dc+ac peak) 1 kV * 1kV 370 V /1.2 kV*
Frequency range 40 Hz - 1 kHz 45 Hz - 5 kHz 45 Hz - 10 kHz
Alternating Current Measurement
Ranges 200.0 mA 200.0 mA 800.0 mA
- 10.00 A - 2.000 A - 10.00 A

Resolution (± LSD) 1 1 1

Absolute Accuracy;

% reading 1.5%(2.5%) 1% 0.4% typical


+ no. of digits 3 5 20
4 (2 & 10A) 1 (10A)
Voltage burden 250 mV ** 800 mV
(0.7 V 2 & 10 A) (1.2V 0.8A )
(150mV 10A)
Overload (max. dc+ac peak) As for direct current As for direct As for direct
current current
Frequency range 40 Hz - 1 kHz 45 Hz - 5 kHz 45 Hz - 10 kHz

51
Resistance Measurement
Ranges 200.0 Ω 200.0 Ω 320.00 Ω
-20.00 MΩ -20.00 MΩ -32.000 MΩ

Resolution (+ LSD) 1 1 1
Absolute Accuracy;
% reading 0.75% 0.2% <0.01% typical
(2% 20MΩ) (0.05% 320 Ω)
+ no. of digits 1 1 30
5 (20 MΩ) 3 (200 Ω) 160 (32 MΩ)
4 (200Ω) 50 (3.2 MΩ)

Open Circuit Voltage (2.6 V 200 Ω)


Overload (max. dc+ac peak) 250 V 350 V 370 V
* See manufacturer's instruction book for further details.
** Not quoted in manufacturer's instruction book.

4.5 Power Supplies


There are portable power supplies of two main types: fixed output voltage and variable output
voltage. The first are usually dedicated to a particular experiment, and will not be discussed
any further. These notes refer specifically to the Farnell L30 series, Farnell E30 series and
the Weir 4000 series power supplies. There are other models available, and if you need to
use one of these, then you MUST read the manufacturer's instructions before use, else
considerable damage to equipment and personnel (i.e. you!) could occur.

4.5.1 Voltage Control


A constant voltage can be supplied to a circuit, whatever the resistance of the circuit, unless
the circuit resistance is so low that the supply won't push through any more current. In this
case, the output voltage will drop (c.f. "lost volts" in a battery). On all models, this voltage
can be set from 0-30V (E30: 0-15 V as well, depending on push-button selection). The
voltage is set by using the coarse and fine control knobs. The terminal voltage can be read by
the analogue meter on the Farnell supplies if the selector switch is set to VOLTS, and is
displayed on the upper digital meter on the Weir 4000 if the DVM button is pushed IN. For
all supplies, if an accurate value of the supply voltage is required, it is as well to measure it
with a separate meter.

4.5.2 Current Limiting


If the external load has too low a resistance, then current limiting sets in. The output voltage
falls until it reaches a value such that the current through the load equals a pre-set maximum.

52
This maximum current can be set by you (except for the E30, where it is pre-set) to suit
circumstances. This is useful to prevent burning out of components.
The procedure for setting the current limit on the L30 and the 4000 units is slightly different.
For the E30 and the L30, if an accurate value of the current is required, it is best to measure it
with a separate meter.

4.5.3 Setting Current Limit on Farnell L30


1. Before turning the main switch on, make sure the load is disconnected by setting the
output ON/OFF switch OFF.
2. Set the voltage controls to maximum, and turn the CURRENT LIMIT to minimum.
3. Switch the meter to CURRENT.
4. Connect a short length of copper wire across the output terminals.
5. Turn mains ON, turn the output ON/OFF switch on, and raise the current slowly until the
meter reads the desired limit. A red LED shows if current limiting is in action; it is below
the output switch. If you switch the meter to VOLTS, you will see that the terminal
voltage drops to a very low level.
6. Turn the output switch off, remove the short and connect the circuit.
7. Turn the voltage control knobs to provide the output voltage needed, and turn the output
switch on. If the resistance of the load is high enough, then the current is below the limit,
and control is by the voltage control knobs. If the resistance is too low, then current
limiting has control, the terminal volts will drop, and control can only be regained if the
voltage control knobs are turned to reduce the volts.

4.5.4 Setting Current Limit on Weir 4000


1. Before turning the main switch on, make sure the load is disconnected by setting the
"OUTPUT" button "OUT" so that the supply will not give any current.
2. Set the "DVM" button "IN", so that the display will show the voltage being supplied.
3. Turn up the voltage adjust to maximum.
4. Switch on with the red button.
5. The current to be supplied can now be set with the lower knob, its value being given by
the lower display.
6. Turn the voltage control knobs to provide the output voltage needed, and turn the output
switch on. If the resistance of the load is high enough, then the current is below the limit,
and control is by the voltage control knobs. If the resistance is too low, then current
limiting has control, the terminal volts will drop, and control can only be regained if the
voltage control knobs are turned to reduce the volts.

4.5.5 Current Control


With low resistance circuits it is often easier and better to control the current rather than try to
control the voltage. For example, if the circuit resistance is about 1 then a voltage of only 1V
(1/30th of full scale on the Weir 2000) is required to drive the maximum current of 1A. It is
clearly less difficult to set a given current using the whole range of the current control knob
than to use a small range of the voltage control knobs.

53
To use this feature, follow the instructions for current limiting, but leave the voltage control
knobs at their maximum setting. With the Weir supply, when you push the "OUTPUT"
button "IN" the current set will now be supplied to the circuit, and the voltage will adjust as
necessary. The current can be changed simply by setting the lower knob. The whole range of
the knob is used to vary the current from 0 to 1A. Current control on the L30 can be
achieved in a similar manner.
The I terminal on the L30 is used for a current supply controlled using the voltage control
knobs. See the instruction book for details.

4.5.6 Floating Output +


+ 0
All models have floating output, which +
Gnd
means that either terminal may be connected Gnd
- -
to ground. Fig 4.8 shows some arrangements - 0
to give +ve supply, -ve supply, and + +
symmetrical supplies. The Farnell E30 and Gnd
+ -
the Weir 4000 have a ground terminal +
provided on the front panel. The Farnell L30 Gnd 0 + 0
needs a separate ground connection. Gnd
- - - -
Floating outputs also mean that power
supplies may be connected in series to give Figure 4.8 Various circuits for power supplies
higher voltages, though the maximum
voltage permitted is limited. See the manufacturer's instruction book if you wish to do this.
Remember that circuit connections are only made to the red and black connections; the
green terminal is connected to ground only!

54
55
CHAPTER 5 – LABORATORY REPORTS
5.1 The Purpose of a Report
The laboratory notes taken during the course of an experiment will rarely be suitable for
communicating the aims, procedures and results to another scientist. Nearly all scientific
investigations therefore have to be presented as papers in scientific journals if the information
obtained is to be of use to people other than the experimenters. The writing of such papers is
not easy. The laboratory reports written about Course 1 experiments are therefore mainly
intended as an exercise in such writing, although you will often also find them valuable in
clarifying your thoughts about experiments. As an exercise try and find one of your tutor’s
recent publications to give you an idea of what is required from a professional physicist.
A report is intended to convey specific information to a specific person or class of people.
Before writing it is important to decide on the purpose(s) of writing the report (e.g. to say that
g = 9.817895 m s-2 and/or to report on a novel way of measuring g and/or to indicate some of
the difficulties involved in measuring g). Be consistent about these aims from the title
through to the conclusions. While writing it is also necessary to have in mind the reader for
whom the report is intended so that you know what knowledge you can assume on his/her
part. For the purpose of the laboratory report exercise you should assume that the reader will
be a capable undergraduate at about the same level as yourself, but who has never seen or
heard about the Course 1 laboratory in which you are working.
A report represents your final considered opinion on an experiment, and in real life it is the
only thing that most readers will ever see of your scientific work. It is essential that it is
absolutely correct. Mistakes that are apparent to the reader may lead him to treat the work as
unreliable, and mistakes that are not apparent may be misleading and might have very serious
consequences.
The report should be a continuous piece of prose. Additional information such as tables,
figures and appendices may be included provided that each item ("table 1", "figure 6", etc.) is
drawn to the attention of the reader at the relevant point in the text. Set them out on separate
pages and include them wherever you and the reader find it convenient.
The amount of scientific work being published these days is so large that the need for
conciseness is very great. Laboratory reports should therefore be as short as possible without
omitting essential information.
There are certain conventions in writing scientific papers, some of which arise from the needs
of presentation in typescript or in print. As far as possible laboratory reports should be
written in the same way, with the exception that we ask for a full explanation of the estimate
of precision.
Choose to write a report on an experiment that worked well and has a clear theme and
purpose - it's hard to make a satisfying "story" if you end up with an unexplained "wrong"
result. You should also be sure you understand the experiment! You may care to write up a
project. The marker will probably find it more interesting because it's more novel! You may
decide to only write up part of an experiment. This is permissible but only if the report forms
a self-contained whole.

5.2 Structure of a Report


Here are some of the main points to be observed in laying out a report.

56
(a) Title
This must indicate clearly what the report is about; do not use the brief titles used in the
laboratory to identify the experiments: "Young Modulus" does not even tell the reader
whether you have performed an experiment or done a theoretical calculation! On the other
hand, a title longer than two lines is definitely too long!

(b) Abstract
An abstract is normally not more than one paragraph in which the aims, results and principle
conclusions are set within a few sentences. It is a summary of the whole report, including
numerical results and verbal conclusions. This will involve repetition of parts of other
sections of the report, but in real life the abstract will be the most widely read part of your
report! The abstract is often the last part of the report to be written (and the most difficult),
even though it appears first! This is because it must be a summary of what you have actually
said, not what you intended to say. Often the "story line" gets altered as you write. Upon
completion, the author(s) of a paper must decide if the abstract is suitable for direct inclusion
in an abstracting service.

(c) Introduction
This sets the experiment in the context of Physics as a whole. Do not just state baldly the
"aims of experiment" here, though it might play a part. Do not, in this section, rush into
describing the apparatus in detail, though you may wish to include a statement of the general
principles of the method. Again, the introduction is often written late in the process; you
need to angle it to point in the direction you actually went! You should describe the purposes
of the experiment and give enough background to ‘set the scene’.

(d) Theory

(e) Description of Apparatus

(f) Procedure
Which first? The purpose of the theory may not be clear unless the apparatus is described
first. The operation of the apparatus may not be clear unless the theory is described first.
This is a problem you have to solve, perhaps by giving a verbal summary of the theory within
the procedure section, or by presenting these sections in a different order. It may be that one
of these sections is superfluous, and the material fits naturally into another section.
These sections should assume an intelligent reader who knows nothing about the experiment.
It is very hard to check that you have told the reader everything she needs to know in a
logical order. You will probably have to rewrite them to get the balance right for handing in.
An appendix usually helps.
It used to be the firm convention that all experiments were written up in the passive voice and
past tense (e.g. "the diameter was measured"). Although stilted, it is still by far the most
common, because "I measured" sounds too pushy! Non native-English speakers also find the
stilted style easier to comprehend. Do not write a recipe: "the length must be measured".
Describe what was done in the past tense, not what is or what ought to be done.
Mathematical equations should be set out as in good books, i.e. they should be included into
the grammatical structure of a normal sentence thus (note the comma!) :
"The area, A is given by

57
b

A= ∫ ydx
a
, (6)

where a and b are the x-co-ordinates of the two wires."


The number (6) is included if it is needed to refer to the equation later in the report. Theory
should be summarised, and details of the full theory either put in an appendix (see (j) below)
or a "reference" given (see (i) below).There is no need to derive well-known equations.

(g) Results
The results section is not a list of numbers. Instead it is a passage written in your very best
English. It should include a description of how the results are analysed, as well as containing
critical values. Results of error analysis should also be presented. Details of purely
arithmetical operations (such as substituting numbers into a formula and calculating the
answer) should not be included.
All numerical data should be given in S.I. units using the correct abbreviations.
Although they wouldn't appear in such detail in a proper scientific paper, we ask you to copy
out your original results into neat tables in your report. This allows the marker to check that
you have analysed the results in the correct way. Also give full details of the calculation of
errors (again unlike a scientific paper).
All the uncertainties involved in an experiment and their relative importance should be
incorporated into an overall uncertainty for the final result(s). When giving results do not
quote more significant figures than are realistic, e.g. 30±3 NOT 29.87185±2.95621.
All readings relevant to the final result must be given. This includes calibrations, checks on
assumptions etc., but preliminary trials etc. can be left out unless they have some special
interest.
Graphs, figures and tables should have captions. Number all figures consecutively,
remember for this purpose that graphs are figures too. Graphs should, of course, have scales
marked on the axes. If you draw a straight line through the points, it should be the computer
fitted line, not one drawn by eye.

(h) Conclusions
Give your final numerical results here along with their uncertainties. Compare if possible
with values obtained by other authors. Make substantiated comments on any discrepancies.
Give verbal conclusions too about other relevant information gained by the experiment, main
difficulties, thought-out suggestions for improvements. Did the experiment achieve it’s
original aim? This is the other section of your paper that, in real life, would have a large
readership - those too busy to read the whole report!

(i) References
Give references to text books, papers etc. which you use in your work. A well known piece
of theory need not be repeated, instead, give a reference so that readers who do not know it
can look it up. References are usually collected at the end of the text, before the appendices.
The reference must have the following style:
Serial number. This is the number given where the relevant material appears in the text,
usually as a superscript, e.g. (5).

58
Author(s), Surname and initials
Title of book (or journal name if referring to a paper). If not the first edition, give the edition
number.
Publisher and where published (this helps the reader's library to get it if it is not in stock)
Date of publication
Specific pages referred to (helps in ordering photocopies!)
Example:
(8) A.F. Kip, Electricity and Magnetism, 2nd Ed., McGraw Hill, London, 1975, pp. 120-123.

(j) Appendices
To avoid breaking the flow of the main report, it is usual to relegate some material (e.g.
algebraic details of theory) to an appendix. In the main text, you must refer to the appendix
(e.g. "see appendix 1 for details") at the appropriate point so that the reader realises that the
material is available.

5.3 Submitting Reports


Reports should be word processed, and presented in a plastic binder available from Stanford
and Mann or the Guild of Students. Print on one side of the paper only and number the pages.
Credit will be given for neatness and the general quality of presentation of the whole report.
Please keep a copy of your report, either a photocopy or on floppy disk, in the unlikely event
of your report being mislaid.
Reports should be handed in at the laboratory by the deadline. You will be asked to sign a
list to verify that your report was handed in on time. Late reports should still be handed in,
but bear in mind that lateness without a valid reason may result in a penalty.
On the front of the report a title, your name, your group (X or Y), and your tutor’s name
must be given.

5.4 Sample Report


This sample report is written to give some idea of the style used in such scientific papers.
Study it in conjunction with section 5.2 above. Note how the authors have attempted to solve
the various problems in communication, not always successfully! Do not attempt to copy the
lay-out too closely when writing on a different experiment for which other sub-titles etc. may
be more appropriate. Here are some questions to bear in mind as you read the different
sections. Other points may well occur to you.
Title: Does it summarise the experiment? Can it be shorter?
Abstract: Is it an accurate summary of the experiment? Is the abstract complete by
itself? Should it state the suspected cause of the possible systematic error?
Should it quote the accepted value of λ?
Introduction: Does it start at the right level of knowledge for a C1 student? Does it end with
a clear idea of the purpose of the experiment?
Theory: Is it understandable? Did the introduction give enough detail of the layout so
that variables used and how they relate to each other are understandable?
Have any details have been left out, should they be put in for completeness? Is

59
the derivation of equation (1) needed if it is in an optics text-book? Should the
way the lens is used be explained in the procedure section?
Apparatus: Could the apparatus be set up from the description given? Are the references
to fig. 2 satisfactory? Would this be better before the theory section?
Procedure: Could the experiment be repeated by following the procedure? Should more
detail about the alignment check be given? Are all the tables of results
properly referenced? Is this section in the right place in the report? Would it
be better to merge this with the apparatus section?
Results: Is this in connected English?, Has the author lapsed into note style anywhere?
Are all the tables properly referenced and captioned. Are all the essential
results present? Are there any data which should not be there?
Conclusion: Are the results examined critically? Are all the possible sources of errors
considered? Is the comparison with previous work satisfactory?
Overall: Does the author tell a good story? Does the author understand the experiment?
Has the author presented the argument in an easily understandable way? Has
the author remembered that the reader has not seen this experiment at all? Are
the final result and conclusions believable? Could the whole report be
shortened without losing too much of the flavour?
Overleaf, there is a copy of the check sheet that we use for student reports. You might like to
fill it in for the sample report, and also to give the report itself a mark out of 10.

60
COURSE 1 LABORATORY REPORTS
To the marker: Please indicate below your opinion of the quality of the report by marking the appropriate
squares. Use the final column if you wish to draw attention to particular points. Please also make constructive
suggestions for improvements by writing directly on the report itself. Record an overall grade at the end of
the report. Sign your name at the end too.
To the student: The demonstrator marking your report will have made comments as s/he sees fit and will have
assigned a grade to your work. The chart below is intended to summarise the quality of the main parts of your
report. The assessment of reports is a subjective affair and you may not agree with some of the criticisms made.
However, if your report created a bad impression with even one reader, it is well worth considering the reasons.
Please discuss your report with the marker if the comments made are not clearly justified. Note that the final
grade is based on the overall impression of your report and is not necessarily linearly related to the marks
on the chart.

Very Good Average Poor Very Brief Comment


Good Poor

Abstract

Introduction

Presentation of Theory

Description of Apparatus
or Database (catalogues
or images etc.)
Discussion of Method
(including special
precautions
Presentation of Results

Sources of
error/uncertainty and
their treatment
Presentation of Tables
(including headings and
captions)
Presentation of diagrams
(including captions)

Conclusions

General Clarity

Conciseness

61
SCHOOL OF PHYSICS AND
ASTRONOMY

PHYSICS COURSE 1 LABORATORY


First Semester Report
2014

The Use of the Fresnel Biprism to Determine


the Mean Wavelength of the Sodium D-Lines

by T.W. O'Lecturers
(Group X)

Tutor: Dr. Ivor Pain

62
The Use of the Fresnel Biprism to Determine the Mean Wavelength of the
Sodium D-Lines
Abstract
The separation of the interference fringes produced when a Fresnel biprism was illuminated
with a narrow source of sodium light was determined. The major source of systematic error
was tentatively identified and a suggestion made for improving the equipment. The final
value obtained from the mean wavelength of the D-lines was (576 ± 4) nm.

Introduction
It is well known that when light from two coherent sources is brought together, the two wave
trains superpose to give constructive and destructive interference, which is observed as a
pattern of alternating light and dark bands. A familiar example of how this is achieved is by
Young's slits arrangement (Fig 1a). Here the two coherent sources are the two narrow slits S1
and S2 illuminated by the slit source S.
The Fresnel biprism (Fig. 1b) is a less familiar way of producing two coherent sources (1)(2).
It relies on the fact that light passing through a thin prism is deviated through a small angle,
and appears to come from a source displaced slightly from the original. Two prisms
cemented base to base will produce light that appears to come from two sources S1 and S2,
these being virtual images of the slit source S.
By measuring the separation of the interference fringes, the separation of the sources and the
distances between the sources and the screen, it is possible to deduce the wavelength of the
light emitted. In this experiment, it is the wavelength of the yellow emission line from
electronically excited sodium atoms which is determined.

Theory
For a two-source interference pattern produced as described above, it is easy to show(3) that
the separation of the fringes y is given by
λD
y= ,
s
where λ is the wavelength of light, D the distance from the observing plane to the sources
and s is the separation of the sources. For the case of the biprism, D = D1 + D2 , where
D1, D2 are the source-biprism and biprism-observation plane distances respectively (fig. 1).
Although it is possible to measure D1 and D2 , we shall show that it is unnecessary. Let c be
the value of D2 when the microscope is placed at an arbitrary point. If the microscope is
displaced a distance x from this point, and the fringe spacing is measured at this position, then
D2 = c + x and
λ( D1 + D2 ) λ( D1 + c ) λx
y= = = . (1)
s s s
If x is regarded as an independent variable and y as the corresponding dependent variable,
then equation (1) is the equation of a straight line, the slope of which is λ/s. Thus measuring
values of y for different values of x enables one to find λ/s without having to measure D1, D2
or c directly.

63
(a)
H

S
1

S
2

(b)
H

S
1

S
2

Figure 1. (a) Young's slits (b) Fresnel's biprism. Interference occurs where the light from the two
sources S1 and S2 overlap. The plane HK marks the object plane of the microscope used to observe the
fringes.

64
In order to find λ, it is necessary to find s. A convenient method for determining s is to use a
converging lens to produce real images of S1 and S2. If a lens is placed such that when it is a
distance u from the two sources, the images are a distance v from the lens, then the separation
of the images is ∆ = vs/u(4). There is a conjugate position where the distance between the
image and object is the same, and that is when the object distance is v and the image distance
is u. The separation of the two images in this case is δ = us/v. Thus by measuring ∆ and δ, it
is possible to deduce a value for s, using
s =√(∆δ) (2)
From the known values of λ/s and s, it is now possible to deduce λ.

Apparatus
The equipment was arranged on an optical bench as shown schematically in Fig. 2. The
vertical slit S was illuminated by a sodium vapour discharge lamp Na. The biprism P was
placed a short distance from the slit on the other side from the lamp. A microscope M was
mounted on a slide which allowed movement along the bench and also in the transverse
direction. The transverse motion was provided by a micrometer screw gauge G. The slide
carried a vernier V which was aligned with a scale C on the optical bench. The auxiliary lens
L used to measure the source separation is not shown in Fig. 2, but was inserted in a lens
carrier between M and P when required.
G

Na P
S
M

Figure 2. Schematic arrangement of apparatus. The letters labelling the various components are
explained in the text.

Procedure
The microscope was oriented so that its optic axis lay, as nearly as possible, along the axis of
the optical bench; and on this common line were set the slit and the refracting edge of the
biprism. A transverse drive on the saddle carrying the biprism enabled this adjustment to be
made. Adjustments for height were carried out in the first instance with a millimetre scale,
and finally by sighting through the microscope. Interference fringes appeared when the
biprism was rotated about a horizontal axis to make its refracting edge parallel to the length
of the slit. Exact parallelism was assumed to occur when maximum contrast in the fringes
was observed. The final check on the alignment of the equipment consisted in observing the
centre of the interference pattern as the microscope was withdrawn from the slit. Any shift
across the field of view indicated a divergence of the slit-biprism line from the axis of the
bench, an error in the original attempt to set up the optic axis of the microscope or a
combination of both causes. The arrangement was considered to be satisfactory when the
misalignment was less than one milliradian.

65
The microscope was then set at an arbitrary point on the bench. The mean distance between
the interference fringes was determined by setting the cross-wires in the microscope eye-
piece on the centres of successive dark fringes and recording the corresponding screw gauge
readings. The position x of the microscope on the bench was noted, and observations of
fringe separations were repeated for several such positions. The detailed readings for one
scan across the interference field, and a summary of the subsequent ones, are presented in
table 1. The separation s of the virtual sources was deduced with the aid of the auxiliary lens
L. Several positions of the microscope were used and a mean value for s was calculated. The
observations are recorded in table 2.
0.6

0.5

0.4

0.3

y/mm
0.2

0.1

0
0.7 0.9 1.1 1.3

x/m

Figure 3. Fringe spacing y for different positions x of the plane of observation.

A second value for λ was obtained by repeating the experiment with a different slit-to-
biprism distance; that is, with a different separation of the virtual sources. The results are
summarised in table 3. The distance between the slit and the plane in which the fringe
measurements were made ranged from approximately 150 to 650 nm. At larger distances the
fringes became rather faint; at smaller ones they were somewhat too closely spaced for
convenient measurement. The slit-to-biprism distances were approximately 40 and 100 mm.

Results
For each value of x, a set of micrometer readings Y on the dark fringes were taken. The graph
of Y against number of fringe should be a straight line, the slope of which would give the
distance y between consecutive fringes. The laboratory computer programme LEAST was
used to fit a straight line to the data, and the mean value of y together with its error was
computed. A sample set of these values is entered into Table 1, the other sets taken have
been omitted for brevity.
Equation 1 predicts that the graph of y against x is a straight line with slope = λ/s and Fig. 3
shows that this is correct. A fit to the line gives λ/s = (29.21±0.19)x10-5. The measurements
of ∆and δ are given in Table 2, and from these values, the mean value of s is found to be
(1.965±0.005) mm, giving a value of λ of (574±4) nm. The measurement from the repeat

66
run with a different value of s were treated similarly. The data of Table 3 gives λ/s =
(8.07±0.05)x10-4, while the mean value of s for this case was (0.723±0.007) mm, giving λ=
(584±7) nm. The weighted mean value is λ = (576±4) nm.

Comments
Although the two values of λ obtained agree to within their respective statistical errors, the
mean differs from the published value (5) of 589.3 nm by about three times the error. It is not
clear what the source of the discrepancy is, but it is suspected that it may lie in the
determination of s.
The location of the exact position of focus depended on the subjective assessment of the
observer. The relatively low scatter in the values of s show that the observer could be
regarded as consistent, but the possibility of a systematic error in judgement of focus cannot
be precluded. Repeating the measurements of s with different observers would show how the
results are affected by observer bias. Replacing the lens with a better quality one might
improve the accuracy, as the images formed by L were rather ill-defined, and it was difficult
to decide on the exact position of focus.
Apart from this, we have found that the Fresnel biprism is a convenient way of producing
interference fringes, and it is possible to determine wavelengths of visible spectral lines to a
statistical accuracy of about 1%.

References
(1) Longhurst R.S., Geometrical and Physical Optics, (Longman) (1957) p 125
(2) Hecht E, Zajac A., Optics, (Addison Wesley, Reading, Mass) (1974) p 284
(3) Longhurst, loc. cit. p 123
(4) Longhurst, loc. cit. p 13-14
(5) Kaye G.W.C., Laby T.H., Tables of Physical and Chemical Constants, 14th Edition
(Longman) 1973, p 88.

67
Table 1 Measurements of fringe separation
x = microscope vernier reading on bench scale
Y = micrometer reading on dark fringes
y = distance between consecutive fringes
x[mm] Fringe No Y[mm] Fringe No Y[mm] Fringe No. Y[mm]
1304.0 1 12.349 8 10.914 15 9.650
2 12.134 9 10.721 16 9.339
3 11.916 10 10.546 17 9.134
4 11.708 11 10.323 18 8.940
5 11.500 12 10.106 19 8.472
6 11.326 13 9.930 20 8.549
7 11.127 14 9.732

1200.5 * * * * * *
1100.4 * * * * * *
1000.6 * * * * * *
900.5 * * * * * *
800.0 * * * * * *
{* In the report you submit, these columns would be filled in as for the first row. It has not
been done here for economy of space.}

Table 2 Summary of y results


x[mm] Mean y[mm] Standard error in y[mm]
1304.0 0.16760 0.00085
1200.5 0.13879 0.00025
1100.4 0.10968 0.00037
1000.6 0.07980 0.00017
900.5 0.07980 0.00017
800.0 0.05157 0.00026

68
Table 3 The distance, s, between the virtual sources
Y1 and Y2 are micrometer readings on the two components of the magnified image.
∆ = (Y2 - Y1 ).
y1 , y2 and δ are the corresponding quantities for the diminished image.
s = √(∆/δ)
Y1[mm] Y2[mm] ∆/[mm] y1[mm] y2[mm] δ[mm] s[mm]
11.824 14.665 2.841 12.486 13.845 1.359 1.965
10.988 13.795 2.807 11.662 12.998 1.336 1.937
10.139 13.205 3.066 11.130 12.401 1.271 1.974
10.119 13.216 3.097 11.129 12.410 1.281 1.992
10.174 13.210 3.036 11.179 12.435 1.256 1.953
10.178 13.582 3.404 11.360 12.500 1.140 1.970
10.067 13.509 3.442 11.313 12.430 1.117 1.961
9.841 13.681 3.840 11.398 12.413 1.015 1.974
9.380 13.968 4.588 11.304 12.138 0.834 1.956

Table 4 Measurements of Fringe Separation for the Second Determination of λ


Symbols have the same meaning as in table 1.
x[mm] Mean y[mm] Standard error of
mean[mm]
1250.3 0.5120 0.0028
1150.3 0.4205 0.0026
1050.2 0.3434 0.0018
950.5 0.2618 0.0017
850.4 0.1824 0.0010
750.2 0.10156 0.00076

Each mean value of y is deduced from measurements of 10 consecutive fringes.


The mean value of s for this determination is given in the text.

69
70
CHAPTER 6 - PROJECTS
Two laboratory sessions in each semester are occupied by a project. For these, there are no
detailed instructions. A few weeks before you are scheduled to do a project, choose a partner
to work with. Note that the partner you choose for the first semester project will be the
person you will be working with during the second semester (unless he/she is studying a
different course from you). Some details about the projects will be displayed a few weeks
before the start of projects. Then sign on for the project you have chosen. Project allocation
is strictly on a ‘first come, first served’ basis so you may have to revert to a second choice.
One of the main aims in projects is to encourage the planning of experiments. During the
weeks before the project session start, you and your partner should meet and draw up a
"game-plan" to get you off to a flying start. Some of the things you should think about are:
1. Use skills developed in workshops. Estimate how big is the effect you are going to
measure. Produce some numbers (e.g. If you are going to have to measure the resistance
of a wire, look up resistivity values in a book of constants, and find out what length and
cross-section would give a suitable value of resistance.)
2. What theory do you need to know? Look at an appropriate text book, or work it out
yourself before the lab. session. Then you will know what measurements you need to
take, and how accurately you will need to make them.
3. Think of ways to make the measurements you need. Consider more than one method - the
first one you think of may not be the best (e.g. For resistance measurements; use of a
DMM on the Ω range, measure V and I and use Ohm's law, use a four-point method with a
standard resistor, use a Wheatstone bridge).
4. Find out what equipment is available. Standard year 1 lab equipment is relatively freely
available, but there are some general purpose instruments, (e.g. pressure transducers).
available as well. Some projects have special equipment.
5. Has the equipment chosen got the accuracy and sensitivity you need? (e.g. How many
figures do different digital meters display? What voltage or current is represented by the
least significant digit on each meter? Do you need this sensitivity? What is the largest
current that can be passed through the resistance?).
6. Decide on the initial construction, tests or measurements you are going to start with on the
first day.
Jot down your ideas on 1 to 6 in your lab. book. You should be prepared to explain (and
defend!) them before you start, even though they will certainly have to be modified once you
start work. Don't stick rigidly to your game plan; be flexible!
Between the first and second weeks you should consider the progress you have made, and
what you intend to do next. You will need to be able to justify these ideas.
At the end of the last day, you will have to hand your notebook to your demonstrator and get
your project marked off as for a normal experiment. Remember: projects count as three
units, although they only run for two sessions.

71


You might also like