Maths For Science PDF
Maths For Science PDF
for
Science
This book forms part of Open University teaching materials. Details of Open University
modules can be obtained from the Student Registration and Enquiry Service,
The Open University, PO Box 197, Milton Keynes MK7 6BJ, United Kingdom
www.open.ac.uk
Maths
for
Science
Sally Jordan, Shelagh Ross and Pat Murphy
Published by Oxford University Press, Great Clarendon Street, Oxford OX2 6DP in association with
First edition published by The Open University 2002 reprinted 2003, 2005, 2008. This edition first
published by Oxford University Press in association with The Open University in 2013.
above should be sent to the Rights Department, Oxford University Press, at the
address above.
Printed and bound in the United Kingdom by Latimer Trend and Company Ltd., Plymouth.
3.1
Contents
Introduction 1
1.1 Numbers 3
1.2 Fractions 11
Chapter 5 Algebra 89
Acknowledgements 398
Index 399
Introduction
Introduction
Welcome to Maths for Science. There are many reasons for studying maths
and a compelling motivation for many people is that it provides a way of
representing and investigating the nature of the real world. Real-world
contexts could include population statistics, or economics, or engineering.
Here, the context is science in its broadest sense.
Much of science is couched in the language of mathematics. Nearly all
courses in science will assume some mathematical skills and techniques. Some
may also require you to learn elements of maths alongside the science, which
can be hard going. The aim of this book is to equip you with the knowledge
and skills to tackle the mathematical aspects of science courses with
confidence.
It is clearly not possible in the space of one book at this level to discuss all
the mathematical techniques you might need to pursue your study of science
to degree level but, by working through this book, you will acquire a good
array of basic mathematical tools and confidence in using them. Equally
importantly, you will have a foundation that should make it much easier to
learn further mathematics if and when required.
It is no accident that the word ‘confidence’ has appeared in each of the last
two paragraphs. Maths is in some sense a language with its own alphabet,
vocabulary and rules of grammar. With any language the only route to fluency
is use and practice, so that eventually the process of constructing or
understanding sentences becomes automatic and you can then concentrate
wholly on the message behind the words. You should aim to develop a similar
confidence and fluency in carrying out certain important mathematical
operations. There are few shortcuts: the route requires practice, practice and
more practice! Keep paper, a pencil and your calculator to hand as you study,
and use them constantly. You may find it helpful to write out notes and even
to rework some of the examples given in the text as you go along. There are
questions for you to try throughout the book; you should work through each
question as you reach it. Answers and comments are given at the end of the
book, but don’t be tempted to look at these until you have made a serious
attempt at working out the answer for yourself. If you have solved all parts of
a question successfully on your own, then you are ready to move on. If not,
you would probably benefit from trying further examples on the same topic
from the Additional exercises, available with solutions on the accompanying
website. Whilst studying this book, you should expect to spend about half of
your time actually doing maths rather than just reading about it.
Since there are so many makes of scientific and graphics calculators on the
market, each operating differently, it is impossible to state the exact sequence
of keystrokes you will need in order to carry out particular calculations.
Whenever you meet a new type of mathematical operation, you should
therefore check that you know how to perform it on your own calculator. A
calculator symbol (as shown in the margin) will alert you to the points at
which you particularly need to carry out this kind of check. You should refer
1
Maths for Science
2
1 Starting points
Starting points
1
The point to start from is always what you already know. It is assumed that
you are familiar with the everyday usage of the basic arithmetic operations of
addition, subtraction, multiplication and division, and the use of a calculator
to carry them out, the use of decimal notation (e.g. for money) and the
interpretation of information on a chart or graph (of the kind that might, for
instance, accompany a television news item about economic trends). Many of
the early chapters begin with a brief revision of ideas and skills that you will
probably already have met. This chapter, which concentrates on ideas about
numbers – including fractions, percentages and powers – is slightly different
from later ones in that it covers concepts that are the basis for what is to
follow in the rest of the book, so more of it may constitute revision. Please
note that because of the special nature of this chapter, it contains fewer
science examples than subsequent chapters.
If the points covered in Chapter 1 are completely familiar, you need not spend
very long on them, but they are worth checking out thoroughly as they are the
foundation of much that is to follow. Even if it is only for the sake of
revision, you should test your own skills by doing the numbered questions. If
any of the material is new to you, time spent mastering it now will pay rich
dividends later.
1.1 Numbers
‘Numbers rule the universe’
(Pythagoras)
−5 −4 −3 −2 −1 0 1 2 3 4 5
3
Maths for Science
4
1 Starting points
In case hearing about all these different types of numbers leads you to think
that straightforward ‘counting numbers’ hold little interest for scientists,
Box 1.1 shows how a series of numbers, which mathematicians find
interesting in their own right, have also been found to describe intricate
patterns of plant growth.
5
Maths for Science
8 13 55
89
(a) (b)
(c) 13 21
Figure 1.4 (a) In this pine cone there are 8 parallel rows of bracts spiralling
gradually and 13 parallel rows of bracts spiralling steeply. (b) In this
sunflower head, seeds spiral out from the centre: 55 rows anticlockwise and
89 rows clockwise. (c) In this pineapple, scales also follow spirals: 8 rows
spiral gradually, 13 rows spiral a little more steeply and 21 spiral very
steeply.
6
1 Starting points
Note from this example how brackets can be used to make it clear how
numbers and signs are associated. In this case (−150) is being added to
(−100).
There are several simple rules which can be applied when you need to carry
out arithmetic operations involving negative numbers. These rules are given
below, with examples of each. Note that, in line with normal convention,
positive numbers are not preceded by a + sign.
5 + (−3) = 5 − 3 = 2
5 − (−3) = 5 + 3 = 8
If you multiply or divide two numbers which have the same sign, the
answer is positive.
If you multiply or divide two numbers which have different signs, the
answer is negative.
7
Maths for Science
Before reading on, test your understanding of the rules by doing Question 1.1.
Question 1.1
One important feature of both addition and multiplication is that both these
operations are commutative. This is just the mathematical way of saying that
if you add two numbers then the result (called the sum) is identical whichever
number is written first. For example:
5 + 3 = 8 and 3 + 5 = 8
5 × 4 = 20 and 4 × 5 = 20
5 − 3 = 2 but 3 − 5 = −2
and
1
8 ÷ 4 = 2 but 4 ÷ 8 =
2
The commutativity of addition and multiplication may seem rather obvious
when applied to the counting numbers, but it is worthy of attention because of
its importance in the algebraic equations that will be introduced in Chapter 3.
8
1 Starting points
to subtract the higher temperature from the lower, i.e. −72 °C − 56 °C, which
gives a difference of −128 °C; Oymyakon is colder than Death Valley by
128 °C.
This example shows that in scientific calculations involving negative numbers
it is important to keep the physical situation in mind.
Question 1.2
6 + (−8) = −2
4 − (−3) = 7
5 × (−3) = −15
(−8) ÷ (−2) = 4
and make sure that you can carry out each of these on your calculator,
obtaining the correct sign on the display of the answer. With some makes of
calculator you will be able to enter the expression on the left-hand side more
or less as it is written, with or without brackets. With other makes you may
have to use a combination of the arithmetic operation keys and the +/− (or on
some makes ±) button.
When you are confident that you can input negative numbers in association
with the arithmetic operations, test your skill with Question 1.3.
Question 1.3
Making sure you input all the signs, use your calculator to work out the
following:
(a) 117 − (−38) + (−286)
(b) (−1624) ÷ (−29)
(c) (−123) × (−24)
9
Maths for Science
There is, however, one case in which the calculator does not fully deal with
signs, and that case concerns square roots. The square root of 9 is defined as
the number that multiplied by itself gives 9. One such number is 3:
3×3=9
and if you use your calculator to work out 9 you will indeed obtain the
answer 3. However, it is also true that
(−3) × (−3) = 9
So the square root of 9 is either +3 or −3. It is a mathematical convention that
the notation 9 means the positive value of the square root of 9, and this is
what your calculator displays. In cases in which the negative value of the
square root might be relevant, this is indicated by use of the sign ± (plus or
minus) before the square root sign, i.e. ± 9 .
In Section 1.1.1, the number 2 was given as an example of an irrational
number. Check that you can use the square root button on your own calculator
to get
5
= 1.290 994 449
3
10
1 Starting points
1.2 Fractions
Fractions are used less in everyday life than was the case 50 years ago. In the
UK, people would measure small lengths in eighths and sixteenths of inches
and then add or subtract these. Nowadays, even if you need to add or subtract
fractions or calculate a percentage of something, most modern calculators will
do this for you. However, the ability to add, subtract, multiply and divide
using numerical fractions (without simply entering the fractions into your
calculator) is an extremely important mathematical skill, because it is the basis
for manipulating algebraic fractions which will be discussed in Chapter 5.
1 2 4
2 4 8
Figure 1.5 exemplifies the most fundamental rule associated with fractions:
11
Maths for Science
In the case of the half cake, the numerator and denominator have been
multiplied by 2 to get the equivalent two quarters and again to get the
equivalent four eighths. In the following example of equivalent fractions,
other multiplying and dividing numbers have been used:
6 2 8 10
= = =
9 3 12 15
2
3
is the simplest form in which this fraction may be expressed, i.e. the one in
which the numerator and denominator have the smallest possible values.
The process of dividing the top and bottom of a fraction by the same quantity
is often referred to as cancellation, because it is commonly shown by striking
through the numbers being divided. For example, 155 can be simplified by
dividing the numerator and denominator by 5, and this may be shown as
1
5
15 3
12 7 19
+ =
16 16 16
Note that at this stage you should add the numerators not the denominators.
12
1 Starting points
3 1 3 × 32 1 × 2
+ = +
2 32 2 × 32 32 × 2
96 2
= +
64 64
98
=
64
49
98
=
64 32
This cannot be simplified any further, so
3 1 49
+ =
2 32 32
13
Maths for Science
Question 1.4
2 1
(a) −
3 6
1 1 2
(b) + −
3 2 5
5 1
(c) −
28 3
Question 1.5
Take any fraction, say 164 , and evaluate it as a decimal, using your
calculator if necessary. Now apply each of the following operations to your
original fraction and, in each case, write the result as a decimal:
(a) choose any integer and add it to the numerator and denominator
(b) subtract the same integer from the numerator and denominator
(c) square the numerator and the denominator (i.e. multiply the numerator
by itself, and multiply the denominator by itself)
(d) take the square root of the numerator and the square root of the
denominator.
The results you obtained for Question 1.5 confirm that, for example, adding
the same non-zero number to the top and bottom of a fraction changes its
value, as do operations such as taking the square root of the numerator and of
the denominator. The experience of all calculations of this type can be
generalised by saying that excluding operations involving the integer zero:
In terms of numerical fractions, this rule may seem fairly obvious. But
forgetting it once the numbers are replaced by symbols is the root cause of
many errors in algebra.
14
1 Starting points
So
3 7 3 × 7 21
× = =
4 8 4 × 8 32
Multiplying three fractions together is done by a simple extension of the
method used in the previous examples:
7 7 3 7 × 7 × 3 147
× × = =
16 8 4 16 × 8 × 4 512
3
In the first few sections of this book, all fractions have been written as, for x 12
4
example, 34 . However, in some maths and science books, you will find that the
alternative form, 3/4, is also used, so you will need to become equally
comfortable with both systems and to be able to swap between them at will. 3 x 12
The way in which fractions are written in printed text can also cause some 4
confusion when multiplying, for example, in finding 34 ×12 it is the 3 (the
numerator) that should be multiplied by 12, not the 4 (the denominator). It is
3
important to make this clear when writing fractions by hand, so the first two
4 x 12
expressions in Figure 1.6 are both correct but the third is not.
15
Maths for Science
this in terms of circles. Each circle contains two half-circles, and 4 circles
therefore contain 8 half-circles. So
1
4÷ = 4× 2 = 8
2
Similarly, 12 ÷ 14 asks how many quarters there are in a half. Figure 1.8
illustrates that:
So
1 1 1 1 4 1× 4 4
÷ = ×4 = × = = =2
2 4 2 2 1 2 ×1 2
So
12
4 5 4 9 4 × 9 36 36 12
÷ = × = = = =
3 9 3 5 3 × 5 15 15 5 5
Here the cancellation has been done by dividing the numerator and the
36
denominator of 15 by 3.
16
1 Starting points
However, cancellation could equally well have been carried out at an earlier
stage,
4 9 3 4 × 3 12
× = =
31 5 1× 5 5
Note that divisions involving fractions are commonly written in several
different ways; the example above might equally well have been expressed as
4 5 4/3
or
3 9 5/9
It is important to remember that an integer is equivalent to a fraction in which
the numerator is equal to that integer and the denominator is equal to 1. For
example, the integer 3 is equivalent to the fraction 13 . So dividing by the
integer 3 is equivalent to dividing by the fraction 13 , and that, according to the
general rule about how to divide by a fraction, is the same as multiplying by
the fraction 13 . Thus
1 1 3 1 1 1×1 1
÷3 = ÷ = × = =
2 2 1 2 3 2×3 6
In this context, it may be helpful to restate the general rule in terms of a
specific example:
1
Multiplying by 2
is equivalent to dividing by 2.
1
Dividing by 2
is equivalent to multiplying by 2.
The specific example above and the cartoon both describe the relationship
between 12 and 2, but the number 2 could of course be replaced by any other
integer: it is equally true to say that dividing by 101 is equivalent to
multiplying by 10.
Question 1.6
Work out each of the following, leaving your answer as the simplest
possible fraction:
2 5 1/ 6 3 7 2
(a) ×3 (b) ÷7 (c) (d) × ×
7 9 1/ 3 4 8 7
1.2.6 Percentages
A percentage means a ‘number of parts per hundred’, so is equivalent to a
fraction in which the denominator is 100, for example
50 1
50% is the same as or
100 2
■ Express 35% as a fraction of the simplest possible form.
17
Maths for Science
□ 35
35% is the same as 100 . The numerator and denominator of the fraction
can both be divided by 5 and doing this gives
7
35 35 7
= =
100 100 20 20
Question 1.7
25
1
(d) Express as a percentage.
75 75 300
× 300 = × = 75 × 3 = 225
100 100 1
18
1 Starting points
So 75% of 300 is 225. Note the way in which the cancellation of the zeros
has been indicated in this calculation, with each cancellation representing a
division by ten.
Question 1.8
2 7
(a) of 20 (b) of 24 (c) 15% of 120 (d) 60% of 5
5 8
Percentages are frequently used to indicate the extent to which something has
increased or decreased. For example, if you buy something for £64 and sell it
for £60, what is your percentage loss? Note that percentage calculations of
this type are always performed relative to the initial value.
In this case, the loss is £64 − £60 = £4. Ignoring the units (£) for now – units
are very important but are considered in Chapter 2 – the loss can first be
expressed as a fraction of the initial value, then converted into a percentage
using one of the methods discussed in Section 1.2.6.
4
= 0.0625 and 0.0625 ×100% = 6.25%
64
The percentage loss is therefore 6.25%.
Question 1.9
increase?
(Note: although only one question of this type has been given in this book,
accompanying website.)
19
Maths for Science
Exponent 1 2 3 4
Power of 5 51 52 53 54
Value 5 25 125 625
If you read this table starting at the right and stepping to the left, each time
you take a step you are subtracting 1 from the number in the top row and
dividing the number in the bottom row by five. On the basis of this pattern,
mathematicians extend this table further to the left by continuing to apply the
same rule for each step, giving:
Exponent −3 −2 −1 0 1 2 3 4
−3 −2 −1 0 1 2 3
Power of 5 5 5 5 5 5 5 5 54
1 1 1
Value 125 25 5
1 5 25 125 625
20
1 Starting points
1 1
N 0 = 1 N −a = Na =
Na N −a
where N represents any base number and a represents any exponent.
Quantities such as those represented by the symbols N and a, which can take
any value, are called variables.
The notation of base numbers and exponents is particularly useful when
applied to powers of ten, because then it ties in with the normal system for
writing decimal numbers, as shown in the table below:
In the next chapter, you will see how useful this powers of ten notation can
be in scientific work.
Question 1.10
1 1 1
(a) 2−2 (b) (c) (d)
3−3 40 104
21
Maths for Science
Your calculator probably has an x2 button, and either an x−1 or a 1/x button,
but to evaluate other powers you will have to use a special ‘powers’ button.
On some calculators this is marked xy, on others it has the symbol ^. To input
a negative exponent, you may have to use a button marked (−) or +/− .
Make sure at this point that you can operate your own calculator to obtain
correctly:
Question 1.11
1
(a) 29 (b) 3−3 (c)
42
Box 1.2 presents a little anecdote about powers.
22
1 Starting points
The process is of course not limited to powers of ten. It works for any base
number. For example,
22 × 24 = (2 × 2) × (2 × 2 × 2 × 2) = 26
Again, the exponent of the result (6) is the same as the sum of the two
original exponents (2 + 4).
The process also works for negative exponents. For example, since
1
5−2 =
5
2
1
53 × 5−2 = (5 × 5 × 5) × =5
5×5
5 can be written as 51, and it can be seen that adding the exponents (3 + (−2))
again gives the exponent of the answer (1).
The fact that adding the exponents gives the exponent of the product can be
written in the general form:
N a × N b = N a +b (1.1)
where N represents any base number and a and b represent any
exponents.
The rules can be extended to cover situations in which one power is being
divided by another. The previous example was:
53 × 5−2 = 53+(−2) = 51 = 5
But as you will remember from Section 1.2.5, multiplying by a fraction is the
same as dividing by that fraction turned upside down (i.e. its reciprocal). So
multiplying by 5−2 is the same as dividing by its reciprocal (52), which gives:
53 ÷ 52 = 53−2 = 51 = 5
This time, instead of being added, the second exponent has been subtracted
from the first. More generally,
N a ÷ N b = N a −b (1.2)
where N represents any base number and a and b represent any
exponents.
23
Maths for Science
Question 1.12
102
(a) 230 × 22 (b) 325 × 3−9 (c)
103
of the answer: 3 × 2 = 6.
More generally,
( N m ) n = N m× n (1.3)
where N represents any base number and m and n represent any
exponents.
Equation 1.3 applies for all values of N, m and n whether positive or negative.
So for example:
3
⎛ 1 ⎞ −20 3 (−20)×3 = 10 −60 = 1
⎜ 20 ⎟ = (10 ) = 10
⎝ 10 ⎠ 1060
This is equivalent to saying that
3
⎛ 1 ⎞ 13 1 1
⎜ 20 ⎟ = = 20 ×3 = 60
⎝ 10 ⎠ 20
(10 ) 3 10 10
Question 1.13
24
1 Starting points
n
The positive nth root of a number N can be written as either N or
as N1/n.
In practice, the first type of notation is only used when n = 2 or n = 3,
and square roots are written simply as N not 2 N .
Question 1.14
The rules expressed in Equations 1.1, 1.2 and 1.3 all apply to fractional
powers, as Worked example 1.4 illustrates.
25
Maths for Science
So
(21/ 2 )7 27/ 2
=
27/2 7−3
3/2
= 22 2
2
4
= 22
= 22
=
4
Equation 1.3 can also be used to bring meaning to a number such as 272/3.
Since 23 = 13 × 2 , applying Equation 1.3 shows that 272/3 = (271/3)2, i.e. the
square of the cube root of 27. The cube root of 27 is 3, so 272/3 is equal to
32 or 9.
Question 1.15
3+2×5=
should you do the addition first or the multiplication first? Try entering this
expression into your calculator exactly as it is written. Do you get the
answer 13?
If so, your calculator (in common with most modern scientific calculators)
knows the convention adopted by mathematicians everywhere that
multiplication takes precedence over addition. The calculator has remembered
the 3 until it has worked out the result of multiplying 2 by 5 and has then
added the 3 to the 10. According to the rules all mathematicians follow, if you
wanted to add the 3 and the 2 first and then multiply that result by 5 you
would have to write
(3 + 2) × 5 = 25
The brackets mean ‘do this first’. Again, check that you can use the bracket
function on your calculator to enter this expression exactly as written on the
left-hand side of this equation and that you obtain the correct answer.
26
1 Starting points
There are similar rules that govern the order of precedence of other arithmetic
operations, which are neatly encapsulated in the mnemonic BEDMAS.
(12 ÷ 3) + 2 = 6
There is nothing wrong with adding such ‘redundant’ brackets – they are
simply there for clarity and can even be entered into your calculator (try it).
Far better to have a few additional brackets than to be confused about the
order in which the calculation should be carried out!
A square root sign and a horizontal line used to indicate division can both be
thought of as containing invisible brackets, i.e. the square root sign applies to
everything within the sign and the division applies to everything above the
line. For example, consider
9+7
and 3 × 9 + 7
8
In each case, the addition (9 + 7) should be carried out first, giving
9 + 7 16
= =2
8 8
and
3 × 9 + 7 = 3 × 16 = 3 × 4 = 12
27
Maths for Science
There is one final quirk associated with the use of brackets. In mathematics,
the multiplication sign is often left out (though its presence is implied)
between numbers and brackets, and between brackets and brackets. So
2(3 + 1) = 2 × 4 = 8
and
(1 + 1) (4 + 3) = 2 × 7 = 14
Some calculators ‘understand’ this convention and some do not. Check your
own calculator carefully using the two examples above.
The next operation in precedence after brackets involves exponents. If there
are powers in the expression you are evaluating, deal with any brackets first,
then work out the powers before carrying out any other arithmetical
operations.
■ Evaluate 2 × 32 and (2 × 3)2.
□ In the first case, there are no brackets so the exponent takes precedence:
2 × 32 = 2 × 9 = 18
In the second case, the operation in the brackets takes precedence over
the exponent
(2 × 3)2 = 62 = 36
Question 1.16
28
1 Starting points
Question 1.17
29
2 Measurement in science
Measurement in science
2
Observation, measurement and the recording of data are central activities in
science. Speculation and the development of new theories are crucial as well,
but ultimately the predictions resulting from those theories have to be tested
against what actually happens and this can only be done by making further
measurements. Whether measurements are made using simple instruments
such as rulers and thermometers, or involve sophisticated devices such as
electron microscopes or lasers, there are decisions to be made about how the
results are to be represented, what units of measurements will be used and the
precision to which the measurements will be made. This chapter will consider
these points in turn. Then Chapter 3 goes on to consider how measurements
of different quantities may be combined, and what significance should be
attached to the results.
31
Maths for Science
symbols are:
≥ greater than or equal to (e.g. a ≥ 4 means that the quantity a may take
with the power of ten, then carry out the multiplication or division.
1 6.4
(c) 6.4 ×10−3 = 6.4 × = = 0.0064
1000 1000
Question 2.1
32
2 Measurement in science
8.31
(c) 0.831 = = 8.31 ×10−1
10
In Worked examples 2.1 and 2.2, all the steps in converting from one form of
a number to another have been written out in full. You may be able to manage
with fewer steps in your own calculations – just use as many or as few as you
feel comfortable with in order to get the right answer.
Question 2.2
It is only too easy to lose track of the sizes of things when using scientific
notation, so you should make a habit of thinking carefully about what the
numbers mean, bearing in mind that numbers may be positive or negative. For
example:
33
Maths for Science
−5 × 102
−500 −400 −300 −200 −100 0 100 200 300 400 500
−5 × 100 5 × 101
−5 × 100 −5 × 10−1
−10 −8 −6 −4 −2 0 2 4 6 8 10
−5 × 10−1 5 × 10−2
−1 −0.5 0 0.5 1
Figure 2.1 Portions of the number line, showing the positions of a few large and
small numbers expressed in scientific notation.
34
2 Measurement in science
Most of the seven SI base units are now defined in terms of unchanging
properties of nature. The SI base unit of time, the second, is defined as
the period over which the waves emitted by caesium atoms under
specific conditions cycle exactly 9192 631 770 times. Then the SI base
unit of length, the metre, is defined by stating that the speed of light in a
vacuum, which is a constant throughout the Universe, is exactly
299 792 458 metres per second.
The SI base unit of mass, the kilogram, is the only fundamental unit that
is still (in 2012) defined in terms of a specific object. The metal cylinder
which constitutes the world’s ‘international prototype kilogram’ is kept in
France (Figure 2.3).
Note that the kilogram is actually the standard unit of mass, not of
weight. In scientific language, the weight of an object is the downward
pull on that object due to gravity, whereas its mass is determined by the
amount of matter in it. When astronauts go to the Moon, where the pull
of gravity is only about one-sixth of that on Earth, their mass remains
the same but their weight drops dramatically! And in zero gravity, they
experience a condition known as ‘weightlessness’.
The SI base unit of temperature is the kelvin, which is related to the
everyday unit of temperature, the degree Celsius:
35
Maths for Science
Figure 2.3 The international prototype kilogram (IPK) is stored under three
vacuum jars in an environmentally monitored safe in the basement of a
building belonging to the Bureau International des Poids et Mesures on the
outskirts of Paris.
You will have noticed that while the base unit of length is the metre, not the
kilometre, the base unit of mass is the kilogram, not the gram.
It is important to realise that, although in everyday usage it is common to say
that you ‘weigh so many kilos’, there are two things wrong with this usage
from the scientific point of view. First, as noted in Box 2.1, the kilogram is
not a unit of weight, but a unit of mass. (The SI unit of weight, the newton,
will be discussed in Chapter 3.) Secondly, in scientific language, ‘kilo’ is
never used as an abbreviation for kilogram, in the sense of the everyday
phrase ‘he weighs so many kilos’. In science, kilo is always used as a prefix,
36
2 Measurement in science
37
Maths for Science
The following data may help to illustrate the size implications of some of
the prefixes:
. the distance between Pluto (which lost its status as a planet in 2006)
and the Sun is about 6 Tm
. a century is about 3 Gs
. eleven and a half days contain about 1 Ms
. the length of a typical virus is about 10 nm
. the mass of a typical bacterial cell is about 1 pg.
Although scientific notation, SI units and the prefixes in Box 2.2 are universal
shorthand for all scientists, there are a few instances in which other
conventions and units are adopted by particular groups of scientists for
reasons of convenience. For example, the age of the Earth, about
4.6 × 109 years, could be written as 4.6 ‘giga years’, but geologists find
millions of years a much more convenient standard measure. They even have
a special symbol for a million years or a million years ago: Ma (where the ‘a’
stands for annum, the Latin word for year). So in Earth science texts you will
commonly find the age of the Earth written as 4600 Ma. It won’t have
escaped your notice that the year is not the SI base unit of time – but then
perhaps it would be a little odd to think about geological timescales in terms
of seconds!
A few metric units from the pre-SI era also remain in use. In chemistry, you
may come across the ångström (symbol Å), equal to 10−10 metres. This was
commonly used for the measurement of distances between atoms in chemical
structures, although these distances are now often expressed in either
nanometres or picometres. Other familiar metric but non-SI units are the litre
(symbol l), which is further discussed in Chapter 4, and the degree Celsius
(symbol °C), introduced in Chapter 1.
There are also some prefixes in common use, which don’t appear in Box 2.2
because they don’t conform to the ‘multiples of 1000’ rule, but that when
applied to particular units happen to produce a very convenient measure. One
you will certainly have used yourself is centi (hundredth): rulers show
centimetres (hundredths of a metre) as well as millimetres, and standard wine
bottles are marked as holding 75 cl. One less commonly seen is deci (tenth),
but this is routinely used by chemists in measuring volumes and
concentrations of chemicals, as you will see in Chapter 4. In the next section
you will also come across the decibel, which is used to measure the loudness
of sounds.
38
2 Measurement in science
Answer
From the definition of the prefix ‘nano’, 1 nm = 10−9 m.
From the definition of the prefix ‘pico’, 1 pm = 10−12 m, so
1
1m = pm = 1012 pm
10−12
then
Question 2.3
The easiest way to work out the order of magnitude of a quantity is to express
it first in scientific notation in the form a × 10n. Then if a is less than 5, the
order of magnitude is 10n. But if a is equal to or greater than 5, the power of
ten is rounded up by one, so the order of magnitude is 10n+1. For example,
the diameter of Mars is 6762 km. This can be written as 6.762 × 103 km, and
because 6.762 is greater than 5, the diameter of Mars is said to be ‘of order
104 km’.
39
Maths for Science
■ What is the order of magnitude of the mass of Jupiter, 1.9 × 1027 kg?
□ Mass of Jupiter ~ 1027 kg (since 1.9 is less than 5, the power of ten
remains unchanged).
The phrase ‘order of magnitude’ is also quite commonly used to compare the
sizes of things, for example a millimetre is three orders of magnitude smaller
than a metre.
40
2 Measurement in science
Question 2.4
Question 2.5
example 2.4.):
(i) how many times taller Mount Everest is than a four-year-old child
(ii) how many typical viruses laid end to end would cover the thickness of
a piece of paper.
In Figure 2.4, a logarithmic scale has been used for ease of display, and the
power of ten for the multiplying factor (102) was chosen because it was the
one that best fitted the page (in drawing diagrams and graphs, you are always
free to choose the scale divisions). Logarithmic scales are used in a number of
fields to measure quantities that can vary over a very wide range. In such
cases, an increase or decrease of one ‘unit’ always represents a
tenfold increase or decrease in the quantity measured. Box 2.3 gives two
examples.
41
Maths for Science
1028
distance to the edge of the
1026 observable Universe
1024
1022 diameter of the
our Galaxy:
1020 Milky Way Galaxy
the Milky Way
1018 distance to the nearest star
1016 (Proxima Centauri)
1014
average distance to Pluto
1012 our Solar System
1010
length in metres
42
2 Measurement in science
sound level in dB
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140
Earthquakes
The Richter scale describes the magnitude of earthquakes. An instrument
called a seismometer is used to measure the maximum ground movement
caused by the earthquake, and a correction factor is applied to this
reading to allow for the distance of the seismometer from the site of the
earthquake. Seismometers are very sensitive and can detect minute
amounts of ground movement (they have to be shielded from the effects
caused just by people walking near them), but some earthquakes can
produce ground movements millions of times greater than the minimum
detectable limit. To cope with this huge variation, the Richter scale is
logarithmic: an increase of one unit on the scale implies a tenfold
increase in the maximum ground movement. A magnitude 2 earthquake
can just be felt as a tremor. A magnitude 3 earthquake produces 10 times
more ground motion than a magnitude 2 earthquake. Damage to
buildings occurs at magnitudes in excess of 6. The devastating
earthquake close to the coast of Japan in March 2011 had a magnitude
of 9.0.
80 dB − 20 dB = 60 dB
43
Maths for Science
10 dB + 10 dB + 10 dB + 10 dB + 10 dB + 10 dB
and each 10 dB increase corresponds to multiplying the intensity by 10.
So the intensity of a shout is (10 × 10 × 10 × 10 × 10 × 10) = 106 times
greater than a whisper.
Question 2.6
44
2 Measurement in science
would be quoted as 1.3 m to two significant figures – the first digit removed
is 6, so the 2 is rounded up to 3.
The reason for rounding up when the first digit removed is 5 or greater is
clear if you bear in mind that the number that is midway between 1.2 and 1.3
is 1.25. So all numbers between 1.25 and 1.3 are closer to 1.3 than they are to
1.2. Therefore, it makes sense to round up the last remaining digit whenever it
has been followed by a digit between 5 and 9. Note that, by convention, the
digit 5 is rounded up.
It is worth noting one common pitfall of rounding. If you were asked to round
3.749 m to two significant figures, the answer would be 3.7 m. It is clear that
3.749 m is less than 3.75 m so do not be tempted to round twice, i.e. to round
first to 3.75 m and then to 3.8 m. However, sometimes a legitimate ‘chain
reaction’ occurs when rounding. Imagine that you want to express 89 999.7 m
to five significant figures. The sixth digit is a 7 so you need to round up, so
the fifth digit should be increased by one. However, as the fifth digit is 9,
adding one to it turns it to 10, which means that you have to add one to the
fourth digit, and so on. So 89 999.7 m to five significant figures is 90 000 m.
Sometimes rounding will produce a zero as the final digit; for example both
1.803 m and 1.798 m round to 1.80 m to three significant figures. Don’t be
tempted to ignore the final zero in cases such as this; it contains important
information. In the example, the zero makes it clear that the number is given
to three significant figures (1.8 m implies two significant figures). In general,
trailing zeros after a decimal point ‘count’ as significant digits when
determining the number of significant figures in a value.
Question 2.7
Leading zeros (as in 0.082 m) also require care. These initial zeros only tell
you about the size of the number and not about the precision to which it is
known. The first significant digit in this value is 8, i.e. the number is
expressed to two significant figures. In general, leading zeros do not count as
significant digits when considering the precision of a value.
The temperature quoted above as 16.222 °C (five significant figures) is likely
to actually lie between 16.2215 °C and just less than 16.2225 °C (16.2214 °C
would round down to 16.221 °C and 16.2225 °C would round up to
16.223 °C). In contrast, the temperature quoted as 16.2 °C is likely to lie
between 16.15 °C and just less than 16.25 °C, i.e. it is much less precisely
known. Now consider the value 90 000 m, which was arrived at earlier in this
section by rounding 89 999.7 m to five significant figures. If 90 000 m had
been given with no indication of the number of significant figures, there
would be no way of telling whether the value was given to five significant
figures (i.e. actual value between 89 999.5 m and just less than 90 000.5 m)
or to one significant figure (actual value between 85 000 m and just less than
95 000 m), or any level of precision in between. One way round this
ambiguity is to state clearly the number of figures that are significant; for
45
Maths for Science
= 6.762 ×10(3+3) m
= 6.762 × 106 m
Question 2.8
Care needs to be taken not to use unjustified precision when the results of
measurements are used as the basis for calculations. This will be discussed
further in the following chapter. Chapter 3 will also discuss the units of
calculated values and the points to be noted when using your calculator for
calculations involving scientific notation.
46
2 Measurement in science
Question 2.9
47
3 Calculating in science
Calculating in science
3
There comes a point in science when simply measuring is not enough and it is
necessary to calculate the value of a quantity from values for other quantities
that have been measured previously. Take, for example, the piece of granite
shown in Figure 3.1. If you have measured the lengths of its sides and its
mass, with a little calculation you can also find its volume, its density, and the
speed at which seismic waves will pass through a rock of this type following
an earthquake.
This chapter looks at several scientific calculations and, in the process,
considers the role of units, significant figures, scientific notation and
estimating when calculating in science. In addition, it introduces the use of
formulae and equations.
8.4 cm
5.7 cm
4.8 cm
49
Maths for Science
2 The values for length and width are each given to two significant figures,
Care needs to be taken when multiplying together two lengths which have
been measured in different units. Suppose, for instance, that you wanted to
find the area of a 1 cm by 4 m rectangle. Units of cm × m are meaningless;
you would need to convert the units into the same form before proceeding,
and if in doubt it is best to convert into SI base units. Since 1 cm = 0.01 m,
this gives an area of 0.01 m × 4 m = 0.04 m2.
Question 3.1
Note: the symbols used for SI units are as given in Box 2.1.
50
3 Calculating in science
This is similar to the concept of a chain only being as strong as its weakest
link. An answer is only as precise as the least precise value in the calculation.
For example, in the calculation 3.4 ÷ 2.34 above, the least precise number is
3.4, which has two significant figures. Thus the answer does not justify any
more than two significant figures and should be given as 1.5. Similarly, the
result of the multiplication 8.4 cm × 5.7 cm (used in finding the area of the
top of the granite specimen) should be given as 48 cm2, again to two
significant figures.
■ To how many significant figures should the answer to each of the
following calculations be given:
(a) 5.7 × 4.123
(b) 4.50 × 0.0567
(c) 44.12 ÷ 0.9
(d) (1.23 × 105) ÷ (4.5 × 103)
□ (a) Two significant figures
(b) Three significant figures
(c) One significant figure
(d) Two significant figures.
You should note that the rule is essentially saying that you should identify the
value that has the lowest number of significant figures, and use that to define
the number of significant figures to use in the answer. However, there is one
situation that can catch people out. That is when calculations have exact
numbers in them. For example, if the length of a piece of string is 111 mm,
how long would a piece of string be that was exactly double this length? You
can immediately see that the answer is 222 mm. However, formally, the
calculation here is 2 × 111 mm and, if you had looked for the value with the
lowest number of significant figures (i.e. the 2), you would have concluded
that the answer merited only one significant figure. In fact, the 2 is an integer,
an exact whole number. It is not ‘about 2’ or 2.0 or 1.999 9999; it is exactly
2. In this respect, an integer is the most precise type of number that you can
imagine. Thus, in your simple calculation, the 2 is absolutely precise and the
111 mm is the least precise term. Thus, the answer should be quoted to three
significant figures, i.e. as 222 mm.
51
Maths for Science
For example, if the mass of a bag of potatoes was measured as 12.8 kg, and
you added to it a potato with mass 0.33 kg, the calculated value of the total
mass would be 13.13 kg. However, since the first mass was measured to only
one decimal place, you should quote the total mass to one decimal place too,
i.e. as 13.1 kg (which happens to be three significant figures).
Note that this rule assumes that either the numbers are not written in powers
of ten notation, or the numbers all involve the same power of ten. So if you
are required to add 1.12 × 103 kg and 2.34 × 102 kg, you should write the
calculation either as
1.12 × 103 kg + 0.234 × 103 kg = 1.354 × 103 kg
The base number should then be quoted to two decimal places (since the
1.12 in 1.12 × 103 kg is given to two decimal places), i.e. the answer is
1.35 × 103 kg.
or as
11.2 × 102 kg + 2.34 × 102 kg = 13.54 × 102 kg
The base number should then be quoted to one decimal place (since 11.2
in 11.2 × 102 kg is given to one decimal place), giving an answer of
13.5 × 102 kg or 1.35 × 103 kg, the same as previously.
Question 3.2
6.
732
(i)
1.51
52
3 Calculating in science
of the dangers of rounding errors, consider again the example from the
beginning of Section 3.1.2:
1.452 991 453 × 5.9 = 8.572 649 573 = 8.6 to two significant figures
However, using the intermediate answer to two significant figures gives:
Work to at least one more digit than is required in the final answer, and
just round at the end of the whole calculation.
In the example above, the final answer should be given to two significant
figures, which means that you should work using the result of the first
calculation to at least three significant figures (1.45):
Answer
The addition and subtraction should be carried out first. (There are ‘invisible
brackets’, as described in Section 1.4, around the (3.3 + 1.21) and the
(3.3 – 1.21).)
3.3 + 1.21 4.51
=
3.3 − 1.21 2.09
Since 3.3 is given to one decimal place, the results of the addition and
subtraction are only justified to one decimal place, i.e. as 4.5 and 2.1,
respectively. In considering the division, you then need to think in terms of
53
Maths for Science
significant figures. Both 4.5 and 2.1 have two significant figures, so the final
result should be given to two significant figures too. However, you should
follow the advice to work to at least one additional digit, only rounding at the
end of the calculation:
Note that if you had rounded too soon, and divided 4.5 by 2.1, a rounding
error would have resulted in an incorrect final answer of 2.1.
Question 3.3
(a) ⎛⎜ 4.2 ⎟
⎞ (b)
⎝ 3.1 ⎠ 9.5
54
3 Calculating in science
Question 3.4
8 × 104
(b)
4 × 10−1
(c) (3.00 × 108)2
(e)
To enter a number such as 5 × 10−16 into your calculator, you may need to
use the button labelled something like (−) or +/− (as used in Section 1.1.3) in
Figure 3.3 Examples (a–e) of
order to enter the negative exponent. how various calculators would
display the number 2.5 × 1012.
55
Maths for Science
To enter a number such as 108 into your calculator using the scientific
notation button, it can be helpful to remember that 108 is written as 1 × 108 in
scientific notation, so you will need to key something like 1 EXP 8.
Question 3.5
(2.4 ×104 ) 2
(a)
1.44 ×10−3
However, to get a feel for the relative sizes, you only really need to estimate
the answer. If an estimate is all that is required, it is perfectly acceptable to
work to one significant figure throughout (indeed, working to the nearest order
of magnitude is sometimes sufficient) and since the final answer is only
approximately known, the symbol ‘≈’ (meaning ‘approximately equal to’) is
used in place of an equals sign.
56
3 Calculating in science
It is important that you write out your mathematical calculations carefully, and
one of the functions of the worked examples scattered throughout this book is
to illustrate how to do this. There are three particular points to note from
Worked example 3.2.
57
Maths for Science
Question 3.7
The average distance of the Earth from the Sun is 1.50 × 1011 m and the
distance to the nearest star other than the Sun (Proxima Centauri) is
3.99 × 1016 m. Working to one significant figure throughout, and without
using a calculator, estimate how many times further it is to Proxima
Centauri than to the Sun.
58
3 Calculating in science
Note that this use of the word ‘formula’ is different from the one that
chemists use in writing, for example, CO2 for carbon dioxide.
x+3=8 (3.1)
y=x+4 (3.2)
In Equation 3.1, x can only have one value, i.e. it is a constant. In this case x
has the value 5. In Equation 3.2, x and y are variables which can each take an
infinite number of values, but y will always be 4 greater than x. The values
(of x and y, etc.) which satisfy a particular equation are known as solutions
and if you are asked to solve an equation you need to look for solutions.
In both Equation 3.1 and Equation 3.2, x and y represent pure numbers.
Equations in science are often rather different. Rather than representing pure
numbers, the symbols usually represent physical quantities and will therefore
have units attached.
C = 2πr (3.3)
where C is the circumference of a circle of radius r.
A = πr2 (3.4)
where A is the area of a circle of radius r.
V = 43 πr 3
(3.5)
where V is the volume of a sphere of radius r.
F = ma (3.6)
where F is the magnitude of a force on an object, m is its mass and a is
the magnitude of its acceleration.
59
Maths for Science
E = mc2 (3.7)
where E is energy, m is mass and c is the speed of light.
μ
vs = (3.10)
ρ
where vs is the speed of an S wave travelling through rocks of density ρ
and rigidity modulus μ.
P = ρgh (3.11)
where P is the pressure at depth h in a liquid of density ρ, and g is the
acceleration due to gravity.
Ek = 12 mv 2
(3.12)
where Ek is the kinetic energy of an object with mass m and speed v .
v = fλ
(3.13)
where v is the speed of a wave, f is its frequency and λ is its
wavelength.
q = mcΔT (3.14)
where q is the heat transferred to an object, m is its mass, c is its specific
heat capacity and ΔT is the change in its temperature.
vi + vf
vav = (3.15)
2
where vav is average speed, vi is initial speed and vf is final speed.
v x = u x + a xt
(3.16)
where ux, vx and ax are respectively initial speed, final speed and
acceleration, all in the direction of the x-axis, and t is time.
60
3 Calculating in science
sx = u xt + 12 axt 2
(3.17)
where sx, ux and ax are respectively distance, initial speed and
acceleration, all in the direction of the x-axis, and t is time.
m1m2
Fg = G (3.18)
r2
where Fg is the magnitude of the gravitational force between two objects
of masses m1 and m2, a distance r apart. G is a constant called Newton’s
universal gravitational constant.
1/ 2
⎛ 2GM ⎞
vesc = ⎜ ⎟ (3.19)
⎝ R ⎠
where vesc is the escape velocity, i.e. the speed with which an object
must be fired from the surface of a planet of mass M and radius R in
order just to escape from it. G is Newton’s universal gravitational
constant.
d = [ L (4πF )]1/ 2
(3.20)
where d is the distance at which light from a star of luminosity L has a
flux density of F.
The symbol chosen to represent something is often the first letter of the
quantity in question, for example m for mass, t for time and l for length, but it
isn’t always so simple. Greek letters are also frequently used as symbols, for
example λ (lambda) for wavelength in Equation 3.13 and ρ (rho) for density in
Equations 3.9, 3.10 and 3.11. In a sense it doesn’t matter which symbol you
use to represent a quantity, since the symbol is only an arbitrarily chosen
label. For instance, Einstein’s famous equation (Equation 3.7) is usually
written as E = mc2, but the equation could equally well be written using any
symbols you wanted to use, for example p = qr2, provided you also made it
clear that p was used to represent energy, q was used to represent mass and r
was used to represent the speed of light. However, the use of conventional
symbols, such as E for energy, saves scientists a lot of time in explaining their
shorthand. This book follows convention as far as possible in its use of
symbols. Sometimes the reason for the choice of symbol will be obvious but
unfortunately this is not always the case.
Sometimes a subscript is used alongside a symbol in order to make its
meaning more specific, as in vi , vf and vav used in Equation 3.15 to mean
initial, final, and average speed, and ax in Equations 3.16 and 3.17 used to
mean acceleration along the x-axis. Note that although ax, for example, uses
two letters, it represents a single physical entity; note also that ax is not the
same as ax. The symbol Δ (the Greek upper case delta) is frequently used to
represent the change in a quantity, so ΔT in Equation 3.14 means a change in
temperature T; again a single physical entity is represented by two letters.
61
Maths for Science
A few letters have more than one conventional meaning, for example c in
Equation 3.7 represents the speed of light, but in Equation 3.14 the same letter
represents specific heat capacity. The intended meaning should be clear from
the context. Other letters have two meanings but lower case is conventionally
used for one meaning and upper case for the other, for example v for speed
(or velocity, discussed further in Section 9.4) and V for volume, or t for time
and T for temperature. Care needs to be taken to use the correct case.
Unfortunately some Greek letters look rather like everyday English ones; for
example ρ (rho), used for density, can look rather like the English lower case
p. Some textbooks use lower case p for pressure (this book uses capital P)
and Equation 3.11 (p = ρgh) can then appear to have the same quantity on
both the left- and right-hand sides of the equals sign, especially when written
out by hand. In reality, this formula has pressure on the left-hand side and
density (and other things) on the right-hand side. A similar confusion can arise
because the letter l can look like the number 1.
A final possible source of confusion stems from the fact that the same letter
may sometimes be used to represent both a physical quantity and a unit of
measurement. For example, an object with a mass of 6 kilograms and a length
of 2 metres might be described by the relationships m = 6 kg, l = 2 m, where
the letter m is used to represent both mass and the units of length, metres. In
this book and most other textbooks, letters used to represent physical
quantities are printed in italics, whereas those used for units are not.
Rules of arithmetic, such as the fact that addition and multiplication are
commutative, and the BEDMAS order of operations, apply when using
symbols too.
62
3 Calculating in science
Although the order in which multiplications are written doesn’t matter, various
conventions are generally applied. Note that in Equation 3.3 (C = 2πr), the
number 2 is written first, then the constant π, then the variable r. This order
(numbers, then constants, then variables) is the one that is generally applied.
Similarly, E = mc2 (Equation 3.7) could be written as E = c2m, but it
generally isn’t! Variables that are raised to a power tend to appear at the end
of equations.
It follows from the BEDMAS rules (Section 1.4) that operations within
brackets take precedence, i.e. operations inside brackets should be evaluated
before those outside the brackets. When working with symbols, this means
that an operation applied to a bracket applies to everything within the bracket.
⎛ 2GM ⎞ 1
So in Equation 3.19, the whole of ⎜ ⎟ is raised to the power 2 .
⎝ R ⎠
Equation 3.20 uses two sets of brackets (different styles of brackets have been
used to avoid confusion). The inner, round brackets ( ) are used to indicate
that L should be divided by the whole of (4πF) and the outer, square brackets
[ ] are used to indicate that the whole of L/(4πF) should be raised to the
power 12 .
As discussed in Section 1.4, a square root sign and a horizontal line used to
indicate division can both be thought of as containing invisible brackets. So,
⎛μ⎞
in Equation 3.10, the square root applies to the whole of ⎜ ⎟ and in
⎝ρ⎠
Equation 3.15 the whole of ( vi + vf ) should be divided by two. Throughout
this book, brackets are sometimes used for added clarity even when this is not
strictly necessary. In addition, you are encouraged to add your own brackets
whenever you think doing so would make the meaning of an equation clearer.
It follows from the ‘E’ in BEDMAS that exponents take precedence over
divisions and multiplications, so in Equation 3.7 (E = mc2) the c must be
squared before being multiplied by m. This means that it is only the c that is
squared, not the m. For clarity you could write this as E = m(c2) and it is very
important to remember that mc2 ≠ (mc)2, i.e. that mc2 ≠ m2c2, where the
symbol ≠ means ‘is not equal to’.
The mnemonic BEDMAS also serves of a reminder that multiplications should
be carried out before additions and subtractions, so in Equation 3.16
( vx = u x + axt ), the ax and the t should be multiplied together before the ux is
added.
Finally, note that all of the rules discussed in Chapter 1 for the writing and
manipulation of fractions and powers apply when using symbols, in exactly
the same way as they do when using numbers. So, Equation 3.17 could be
written as
a xt 2
s x = u xt + instead of sx = u xt + 12 axt 2
2
63
Maths for Science
L
d= d = [ L (4πF )]1/ 2
4πF
Question 3.8
Which two pairs of equations given below are equivalent? You should be
able to answer this question by just looking at the equations, but you might
like to check your answer by substituting values such as x = 3, y = 4,
z = 5.
(i) a = x(y + z) (ii) a = xy + z (iii) a = (y + z)x
(iv) a = x + yz (v) a = z + yx
Question 3.9
Two of the equations given below are equivalent. Which two? Again, you
should attempt this question initially by simply looking at the equations.
bac 2 b 2c 2 bc 2
(i) m = (ii) m = a (iii) m = a
d d d
abc 2 b 2 a 2c 2
(iv) m = (v) m =
ad d
64
3 Calculating in science
V = 43
πr 3
4 × π × (6.38 × 106 m)3
=
3
= 1.0878 ×1021 m3
= 1.09 ×1021 m3 to three significant figures
i.e. the Earth’s volume is 1.09 × 1021 m3.
65
Maths for Science
= 5.
4973 ×103 kg m −3
Note, from Worked example 3.3, the following point about the handling of
units.
Units have been included next to values at all times, and the units in the
final answers are both consistent with the working and what you would
expect the units of the final answer to be.
Question 3.10
(a) The granite block shown in Figures 3.1 and 3.2 has a length of
8.4 × 10−2 m, a width of 5.7 × 10−2 m and a height of 4.8 × 10−2 m. Use
the word equation volume = length × width × height to calculate the
volume of the block.
(b) Use your answer to part (a) and the fact that the mass of the block is
0.620 kg to calculate its density. You will need to use Equation 3.9 from
Box 3.1.
ux = 1.5 m s−1 Give your answers in SI units, scientific notation and to an appropriate
Worked example 3.4 illustrates the way in which units should be handled in a
more complicated example.
ax = 9.81 m s−2
Worked example 3.4
Use Equation 3.16 from Box 3.1
v x = u x + a xt
(3.16)
to find the speed reached after 0.45 s by a stone thrown downwards from a
cliff with initial speed 1.5 m s−1. This situation is illustrated in Figure 3.4.
Figure 3.4 A stone being
thrown from a cliff.
You can assume that the magnitude (size) of the acceleration is 9.81 m s−2,
where m s−2 is the SI unit of acceleration.
66
3 Calculating in science
Answer
Equation 3.16 states that vx = u x + axt , and you are trying to find vx . The
question tells you that
ux = 1.5 m s−1 ax = 9.81 m s−2 t = 0.45 s
Thus
The units of axt are found by multiplying the units of ax and the units of t:
m m× s m
×s = = = m s −1
s 2 s× s s
Then, returning to the calculation
vx = 1.5 m s −1 + 4.4145 m s −1
= 5.9 m s −1 to two significant figures
i.e. the speed after 0.45 seconds is 5.9 m s−1.
In Worked example 3.4, values were input with units of m s−1 for initial
speed, units of s for time, and units of m s−2 for acceleration, and the units for
final speed worked out to be m s−1. The units for final speed were not simply
assumed to be m s−1, but rather they were calculated at the same time as the
numerical value. Handling units in this way ensures that the answers are
expressed as physical quantities (with units), not just numbers. It also gives an
easy way of checking a calculation. If the final units in Worked example 3.4
had come out as m2 s−1, you might have realised that, since these are not
units of speed, you must have made a mistake.
It is good practice to work out the units in this way in all your scientific
calculations. To enable you to do this, Box 3.3 explains a little more about
some of the derived units that you will encounter in your study of science.
67
Maths for Science
likely to end up with a unit named after you! The units in Table 3.1 are
named after Sir Isaac Newton, James Prescott Joule, James Watt, Blaise
Pascal and Heinrich Hertz, respectively.
Physical quantity Name of unit Symbol for unit Base unit equivalent
force, such as weight newton N kg m s−2
energy joule J kg m2 s−2
power watt W kg m2 s−3
pressure pascal Pa kg m−1 s−2
frequency hertz Hz s−1
1J=1N×1m
1J
1W =
1s
1N
1 Pa =
1 m2
The following data may help to illustrate the sizes of the units:
. An eating apple has a weight of about 1 N on Earth.
. An athlete with mass 75 kg, sprinting at 9 m s−1, has a kinetic energy
(energy of movement) of about 3000 J.
. A domestic kettle has a power rating of about 2500 W.
. Atmospheric pressure at sea-level is about 105 Pa.
. A human heart beats with a frequency of about 1.2 Hz when at rest.
The unit prefixes, introduced in Box 2.2, are also used with derived
SI units, for example
1 MW = 106 W
A unit prefix not introduced previously, hecto (abbreviated h and
meaning one hundred) is sometimes used when discussing atmospheric
pressure:
1 hPa = 100 Pa
Worked example 3.5 makes use of the fact (from Table 3.1), that
1 J = 1 kg m2 s−2 and Worked example 3.6 makes use of the fact that
1 N = 1 kg m s−2.
68
3 Calculating in science
Ek = 12 mv 2 (3.12)
to find the kinetic energy of the stone shown in Figure 3.4, which has a mass
of 0.175 kg, when its speed is 7.8 m s−1.
Answer
m = 0.175 kg
v = 7.8 m s−1
Substituting into Equation 3.12 gives
Ek = 12
mv 2
= 12
× 0.175 kg × (7.8 m s −1) 2
=
2
= 5.3235 kg m 2 s −2
Now consider the significant figures and the units. Since the speed is given to
two significant figures, whilst the mass is given to three significant figures, the
answer should be given to two significant figures. Since 1 kg m2 s−2 = 1 J, the
units have worked out to be joules, as expected for units of energy.
The answer should be written as Ek = 5.3 J to two significant figures. The
stone has an energy of 5.3 J when the speed is 7.8 m s−1.
69
Maths for Science
1
then, since 1 N = 1 kg m s−2 and kg−2 =
kg 2
1/ 2
⎛ 2 × 6.673 ×10−11 × 5.98 ×1024 kg m s −2 m 2 kg ⎞
vesc = ⎜ ⎟
⎝ 6.38 ×106 m kg 2 ⎠
1/ 2
⎛ 2 × 6.673 ×10−11 × 5.98 × 1024 kg m s −2 m 2 kg ⎞
=⎜ ⎟
⎜ 6.38 ×106 m kg 2 ⎟
⎝ ⎠
= (1.2509 ×10 m s )
8 2 −2 1/ 2
Taking the square root of both 1.2509 × 108 and m2 s−2 gives
vesc = 1.12 × 104 m s−1 to three significant figures.
The escape velocity is 1.12 × 104 m s−1, with units of m s−1, as expected for
a speed.
Question 3.11
Use F = ma (Equation 3.6 from Box 3.1) to find the force experienced by a
mass m = 2.1 kg subject to an acceleration a = 4.6 m s−2.
Question 3.12
70
3 Calculating in science
Question 3.13
71
4 Unit conversions
Unit conversions
4
Units of measurement were introduced in Chapter 2 whilst Chapter 3
emphasised the importance of working out units as well as numerical values
in calculations. You have learnt to convert, say, from picometres (pm) to
metres (m) but the more complicated units introduced so far in this book, for
quantities such as speed, density and energy, have all been in the form of
SI units, expressed or expressible in terms of the base units of metres,
kilograms and seconds. However, quantities are not always expressed in SI
units. This short chapter considers methods for converting values from one set
of units to another.
Note that there are many different methods for converting between units.
Provided that you can obtain the answers given to the questions in this
chapter, and understand why, it is acceptable to use methods other than those
given here.
73
Maths for Science
Question 4.1
1 cm 1
Thus 1 cm2 = 100 mm2 and 1 mm2 = cm2.
100
If you want to convert from cm2 to mm2 you need to multiply by 100; if you
Figure 4.1 A square centimetre want to convert from mm2 to cm2 you need to divide by 100.
(not to scale).
Figure 4.2 illustrates another example which is a little harder to visualise.
Each side of the square measures either 1 km or 1000 m (103 m).
Working in kilometres gives
103 m
area = 1 km × 1 km = (1 km)2 = 12 km2 = 1 km2
Working in metres gives
103 m
1 km
74
4 Unit conversions
48 cm 2 = 48 × 10−4 m 2
= 4.8 × 10−3 m 2 in scientific notation
This is the area of the top of the granite block and reassuringly you can
now see that the same answer was obtained in Section 3.1 and in
Section 4.1.
106 mm
1 km
i.e. 1012. Similarly, to convert from mm2 to km2 you need to divide by
(106)2, i.e. 1012.
Question 4.2
1 km
A desk has an area of 1.04 m2. Express this area in:
(a) cm2 (b) μm2 (c) km2
Figure 4.3 A square kilometre
The method for converting between different units of volume is a direct (not to scale).
extension of the method for converting between different units of area.
Suppose you want to know how many mm3 there are in a cm3. There are
10 mm in 1 cm, so each side of the cubic centimetre in Figure 4.4 measures
either 1 cm or 10 mm. The volume can be written as either 1 cm3 or
103 mm3. Thus 1 cm
1 cm
1
1 cm3 = 103 mm3 and 1 mm3 = cm3
10 mm
103
To convert from cm3 to mm3 you need to multiply by 103; to convert from
1 cm
10 mm
In general, to convert between units of volume you need to cube the
10 mm
conversion factor that you would use to convert corresponding lengths.
75
Maths for Science
One of the mistakes that people make when doing unit conversions is to do
the conversion the wrong way round, i.e. to divide by the conversion factor
instead of multiplying or vice versa. The best way to avoid errors of this type
is to apply common sense. If you were asked to convert 3.7 m3 into cm3, you
would expect the numerical answer to be a lot bigger than 3.7 (there are a lot
of cubic centimetres in a cubic metre!). So an answer of 3.7 × 106 cm3 seems
reasonable, whilst an answer of 3.7 × 10−6 cm3 is clearly wrong.
Another common mistake when converting units of area and volume is to
forget to square or cube the conversion factor used for corresponding lengths.
Recall from Section 2.2 that a decimetre (dm) is a tenth of a metre,
i.e. 1 dm = 10−1 m. The dm3 is used by chemists as a unit of volume. The
litre (l) (also mentioned in Chapter 2) was defined in 1901 as the volume of a
kilogram of water at 4 °C, under standard atmospheric pressure. This volume
turns out to be 1.000 28 dm3, and since 1969 a litre has been defined to be
1 dm3. Worked example 4.1 illustrates the magnitude of the litre relative to
other volumes. (Note: it is very easy to confuse the letter ‘l’, used as the
symbol for litres, with the number 1. Take care!)
Figure 4.5 is a summary of unit conversions for length, area and volume, but
you should try to remember the general principles involved rather than
memorising individual conversion factors.
Question 4.3
76
4 Unit conversions
to co 3 to co 3
nvert multiply by 10 nvert multiply by 10
to c 2 to c 2
onvert multiply by (103 ) onvert multiply by (103 )
to c 3 to c 3
onvert multiply by (103 ) onvert multiply by (103 )
77
Maths for Science
Many things move and/or grow in the world around us, and it is useful to
compare different values for speed or rate of growth. Different speeds are
frequently measured in different units, so in order to be able to compare like
with like it is necessary to convert between different units for distance, time
and speed. Box 4.1 considers various examples of speed and growth, and the
text immediately following the box looks at ways of converting from one unit
to another.
78
4 Unit conversions
(a)
(b)
Figure 4.6 (a) Stalactites and stalagmites growing (very slowly!) in the
Treak Cliff Cavern, Derbyshire, England. (b) The Saskatchewan Glacier,
Banff National Park, Canada.
79
Maths for Science
5 minutes
The speeds given so far in this box have related to processes on the
Earth, but remember that the Earth itself is moving too. The rotation of
the Earth on its axis leads to a movement of up to 0.5 km s−1 at the
surface. In addition, the Earth is orbiting the Sun at about 30 km s−1 and
the entire Solar System is moving around the centre of the Galaxy at
about 250 km s−1.
30 km s −1 = 30 ×103 m s −1
= 3.0 ×104 m s −1 in scientific notation
80
4 Unit conversions
The Earth orbits the Sun with a speed of about 3.0 × 104 m s−1. Again the
answer makes sense: it is reasonable to expect that the numerical value of a
speed in m s−1 will be larger than the same speed when given in km s−1.
Next consider what happens when you need to convert only the time part of
units of speed, for instance in converting from km hour−1 to km s−1.
You know that there are 60 minutes in an hour and 60 seconds in a minute, so
1 hour = 60 × 60 s = 3600 s
However, in this case you don’t want to convert from hours to seconds, but
rather from kilometres per hour to kilometres per second. The way forward
comes in recognising that the word ‘per’ and the use of negative exponents in
hour−1 and s−1 indicate division. So to convert from hour−1 to s−1 (or from
km hour−1 to km s−1) you need to find the conversion factor from hours to
seconds and then divide by it.
Since 1 hour = 3600 s,
1
1 km hour −1 = km s −1
3600
In deciding whether to divide or multiply by a particular conversion factor,
common sense can again come to your aid. It is reasonable to expect that a
speed quoted in km s−1 will be smaller than the same speed when quoted in
km hour−1, so it is reasonable to divide by the 3600 on this occasion.
35
35 km Ma −1 = km year −1
106
= 35 ×10−6 km year −1
= 3.5 ×10−5 km yeear −1 in scientific notation
The plates are moving apart at an average rate of 3.5 × 10−5 km year−1. This
answer is reasonable: you would expect the rate of separation quoted in
km year−1 to be smaller than the same rate quoted in km Ma−1.
81
Maths for Science
Question 4.4
Finally, consider conversions for speed in which both the units of distance and
the units of time have to be converted. This is simply a combination of the
techniques illustrated in Worked examples 4.2 and 4.3. Suppose you want to
convert from km hour−1 to m s−1.
1 km = 103 m
1 hour = 3600 s
To convert from km hour−1 to m s−1, you need to multiply by 103 (to convert
the km to m) and divide by 3600 (to convert the hour−1 to s−1):
103
1 km hour −1 = m s −1 = 0.278 m s −1 to three significant figuress
3600
106
1 km Ma −1 = mm year −1 = 1 mm year −1
106
Thus a speed given in km Ma−1 is numerically equal to one given in
mm year−1. The plates are moving apart at a rate of 35 mm year−1. This is
similar to the rate at which human fingernails grow and is easier to imagine
than is 35 km Ma−1.
Question 4.5
Convert each of the following to values in m s−1 and then compare them.
(a) A stalactite growth rate of 0.1 mm year−1.
(b) The average speed of the Saskatchewan Glacier (12 cm day−1).
(c) The speed of separation of the tectonic plates discussed in Worked
examples 4.3 and 4.4 (35 km Ma−1).
(Note: for the purposes of this question, consider 1 year to be 365 days
long.)
82
4 Unit conversions
83
Maths for Science
1 mg = 10−3 g
Therefore 1 mg = 10−3 × 106 μg = 103 μg
1 l = 103 ml
To convert from mg l−1 to μg ml−1, you need to multiply by 103 (to convert
the mg to μg) and as discussed above you need to divide by 103 (to convert
the l−1 to ml−1).
103
103
Thus a concentration given in mg l−1 is numerically equal to one given in
μg ml−1, in particular 50 mg l−1 = 50 μg ml−1.
10−3
1 g cm −3 = kg m −3 = 10−3−(−6) kg m −3 = 103 kg m −3
10−6
Thus
84
4 Unit conversions
Note that the density in units of kg m−3 has worked out to be 1000 times the
value in units of g cm−3. You can always convert from g cm−3 to kg m−3 by
multiplying by 1000 and similarly, you can always convert from kg m−3 to
g cm−3 by dividing by 1000. However, as was the case with the unit
conversions for area and volume, it is better to consider general principles
rather than trying to memorise conversion factors.
Question 4.6
1 g cm 2 s −2 = (10−3 ) × (10−4 ) kg m 2 s −2
= 10(−3) + (−4) kg m 2 s −2
1 −7 kg m 2 s −2
= 10
1 erg = 10−7 J
Box 4.3 discusses another unit of energy, the kilowatt-hour (kW h). You have
probably encountered this unit before, because it is the unit used in payment
for electricity in many countries (Figure 4.8), but you may not have realised
that the kilowatt-hour is a unit of energy, especially since the watt is the SI
unit of power and the hour is a unit of time.
85
Maths for Science
E = pt (4.1)
The SI units of energy, power and time are related by
1J=1W×1s
Equation 4.1 also provides a means of calculating the amount of energy
(in joules) in a kilowatt-hour. A kilowatt-hour is defined to be the energy
delivered when 1 kilowatt of power is expended for 1 hour, i.e.
p = 1 kW = 103 W
t = 1 hour = 60 × 60 s = 3600 s = 3.6 × 103 s
Substituting these values in Equation 4.1 gives
E = pt
= 103 W × 3.6 × 103 s
= 3.6 × 106 J s −1 s (since 1 W = 1 J s −1)
1
= 3.6 × 106 J (since s −1 = and the ‘s’s then cancel)
s
So one kilowatt-hour is equal to 3.6 × 106 J.
86
4 Unit conversions
This section has introduced two commonly used but non-SI units of energy.
Scientists also sometimes measure other quantities in non-SI units, and some
of these, like the kilowatt-hour, have misleading names. For example, the
electronvolt (eV) is also a unit of energy whilst the light-year (ly) is a unit of
distance.
Whatever units values have been quoted in, the safest option is to convert to
SI units at the earliest opportunity. This is an extension of the practice
introduced in Section 4.1, but will sometimes involve converting compound
units, for example, of speed, concentration or density, before carrying out a
calculation.
Question 4.7
Use P = ρgh (Box 3.1 Equation 3.11) to find the pressure at a depth of
1.2 km in seawater of density 1.025 g cm−3. (The magnitude of the
acceleration due to gravity is 9.81 m s−2.)
Question 4.8
87
5 Algebra
Algebra 5
At the end of Chapter 3 you used the following equation
μ
vs = (3.10)
ρ
to calculate the S wave speed, vs , of seismic waves passing through a rock of
density ρ and rigidity modulus μ. But suppose that, instead of knowing ρ and
μ and wanting to find vs , you know vs and ρ and want to find μ. The best
way to proceed is to rearrange Equation 3.10 to make μ the subject of the
equation, where the word ‘subject’ is used to mean the term written by itself,
usually to the left of the equals sign. Rearranging equations is the first topic
considered in Chapter 5. The rest of the chapter introduces methods for
simplifying equations and ways of combining two or more equations together.
c=a+b
c=a+b c + 50 = a + b + 50 2 2
Figure 5.1 (a) The analogy between an equation and a set of kitchen scales. The scales remain balanced if (b) 50 g
is added to both sides or if (c) the weight on both sides is halved.
The scales will remain balanced if you add a 50 g mass to one side of the
scales, or halve the mass on one side, provided you do exactly the same thing
to the other side. In a similar way, you can do (almost) anything you like to
89
Maths for Science
one side of an equation and, provided you do exactly the same thing to the
other side, the equation will still be valid.
The following rule summarises the discussion above:
This rule is fundamental when rearranging equations, but it doesn’t tell you
what operation to perform to both sides of an equation in order to rearrange it
in the way you want. The highlighted points below should help with this, as
will plenty of practice. However, two things are worth noting at the outset.
1 Equations are conventionally written with the subject on the left-hand side
of the equals sign. However, when rearranging an equation it can very
often be helpful simply to reverse the order.
So if you derive or are given the equation c = a + b you can rewrite it as
a + b = c; if you derive or are given the equation ab = c you can rewrite it
as c = ab.
2 Even if you choose the ‘wrong’ operation, provided you correctly perform
that operation to both sides of the equation, the equation will still be valid.
Suppose you want to rearrange the equation c = a + b to obtain an
expression for a. You could divide by two, as illustrated by Figure 5.1c,
giving
c a+b
=
2 2
This is a perfectly valid equation; it just doesn’t help much in the quest for
a. The numbered points below give some hints for more helpful ways
forward, and each guideline is followed by an illustration of its use.
In the numbered hints the words expression and term are used to describe the
parts of an equation. An equation must always include an equals sign, but an
expression or a term won’t. A term may be a single variable (such as vx or ux
in the equation vx = u x + axt ), or a combination of several variables (such as
axt); an expression is usually a combination of variables (such as axt or
ux + axt), but the words are often used interchangeably.
90
5 Algebra
This gives
a+b−b=c−b
or
a−b+b=c+b
or
To rearrange ab = c to make a the subject, note that you need to remove the b
from the left-hand side of the equation. The a is currently multiplied by b, so
you need to divide both sides of the equation by b. This gives
ab c
=
b b
The b in the numerator of the fraction on the left-hand side cancels with the b
in the denominator to give
c
a=
b
a
To rearrange = c to make a the subject, note that you need to remove the b
b
from the left-hand side of the equation. The a is currently divided by b, so
you need to multiply both sides of the equation by b.
91
Maths for Science
This gives
a×b
= c×b
b
The b in the numerator of the fraction on the left-hand side cancels with the b
in the denominator to give
a = cb
Hint 5 If you are trying to make a term the subject of an equation and
you currently have an equation for the square of that term, take the
square root of both sides of the equation.
a=± b
Note the presence of the ± sign, indicating that the answer could be either
positive or negative, as discussed in Section 1.1.3. In practice, the reality of
the problem you are solving sometimes allows you to rule out one of the two
values.
Hint 6 If you are trying to make a term the subject of an equation and
you currently have an equation for the square root of that term, square
both sides of the equation.
a = b2
Hints 1 to 6 all follow from a general principle:
The following worked examples use the principles introduced in the numbered
hints above, in the context of equations which are frequently encountered in
science. Worked example 5.1 also involves substituting numerical values and
units into the equation once it has been rearranged.
92
5 Algebra
mg = W
To isolate m you need to get rid of g, and m is currently multiplied by g so,
from Hint 3 you need to divide by g. Remember that you must do this to both
sides of the equation, giving
mg W
=
g g
The g in the numerator of the fraction on the left-hand side cancels with the g
in the denominator to give
W
m=
g
Substituting values for W and g gives
649 N
m=
9.81 m s −2
Since 1 N = 1 kg m s−2 (from Box 3.3), the units are
N kg m s −2
= = kg
m s −2 m s −2
You then have
649 kg m s −2
m=
9.
81 m s −2
= 66.157 kg
cant figures
So the teenager’s mass is 66.2 kg.
93
Maths for Science
4π2 L
T2 =
g
where g is the magnitude of the acceleration due to gravity. Write down an
equation for T.
Answer
T is currently squared, so from Hint 5, you need to take the square root of
both sides of the equation. This gives
4π2 L
T=
g
Since T is a period of time, its value must be positive, so you only need to
write down the positive square root.
Question 5.1
m
(c) Rearrange ρ = to obtain an equation for m.
V
When rearranging more complicated equations, it is often necessary to proceed
in several steps. Each step will use the rules already discussed, but many
people are perplexed when trying to decide which step to take first. Expertise
in this area comes largely with practice, and there are no hard and fast rules
(it is often possible to rearrange an equation by several, equally correct,
routes). However, the following guidelines may help:
94
5 Algebra
nRT = PV
You now need to remove the nR by which the T is multiplied. Dividing both
sides by nR (Hint 3) gives
nRT PV
=
nR nR
The nR in the numerator of the fraction on the left-hand side cancels with the
nR in the denominator to give
PV
T =
nR
95
Maths for Science
i.e.
ρV = m
Then dividing both sides by ρ gives
ρV m
=
ρ ρ
i.e.
m
V =
ρ
gt2 = 2h
96
5 Algebra
gt 2 2h
=
g g
i.e.
2h
t2 =
g
Now you can take the square root of both sides (Hint 5) to give
2h
t=±
g
(t in this case is time, so you only need to consider the positive value.)
μ
vs =
ρ
Answer
⎛μ⎞
You can consider there to be brackets around ⎜ ⎟ and start by finding an
⎝ρ⎠
⎛μ⎞
expression for ⎜ ⎟ (Hint 9).
⎝ρ⎠
The equation can be written as
μ
= vs
ρ
μ = vs 2 ρ
97
Maths for Science
is useful (arguably the most useful skill developed in this book), but it is
not something that you should do just for the sake of doing so, but rather
because you want to work something out, and rearranging an equation is
the means to this end. Suppose you have been told that S waves pass
through rocks of density ρ = 3.9 × 103 kg m−3 with a speed
vs = 3.0 × 103 m s−1, and you want to find the rigidity modulus μ. The
equation in the form
μ
vs =
ρ
is not much use, but the rearranged form immediately tells you that
μ = vs 2 ρ
= (3.0 × 103 m s −1) 2 × (3.9 × 103 kg m −3 )
= (9.0 × 106 m 2 s −2 ) × (3.9 × 103 kg m −3 )
= 3.51 × 1010 m 2 s −2 kg m −3
= 3.5 × 1010 kg m −1 s −2 to two significant figures
So the rigidity modulus is 3.5 × 1010 kg m−1 s−2.
In simple examples you may be able to work out values without first
rearranging the equation. However, it is always best to rearrange the
equation so that the quantity you are trying to find is on the left-hand
side, before substituting numerical values. This approach avoids
arithmetic slips and rounding errors and is discussed further in
Chapter 6.
Question 5.2
98
5 Algebra
Question 5.3
The mass, m, speed, v , and kinetic energy, Ek, of an object are linked by
the equation Ek = 12 mv 2 .
(a) Rearrange this equation so that v is the subject.
(b) Use your answer to part (a) to estimate (in m s−1 to one significant
2 × 103 J
(ii) an athlete of mass 70 kg to have the same kinetic energy as the tectonic
plate in part (i).
The final group of worked examples in this section involve equations which
may appear rather more complex than the previous ones, but they can all be
rearranged using the rules and guidelines already introduced. Some, like
Worked example 5.8, appear more complex partly because they use symbols
that are rather unwieldy. However, these final worked examples are genuinely
more complicated too, and are best solved by taking a logical stepwise
approach (as the early Arab mathematicians did; see Box 5.2). Rearranging
complicated equations is rather like peeling away layers of an onion,
systematically removing layer by layer in order to get to the part you want.
But that doesn’t mean it should end in tears!
99
Maths for Science
The book, whose title Hisab al-jabr w’al muqabela, can be translated as
‘Transposition and reduction’, explained how it was possible to reduce
any problem to one of six standard forms using the two processes, al-jabr
(transferring terms to eliminate negative quantities) and muqabela
(balancing the remaining positive quantities).
Arab mathematicians like al-Khwarizmi did not use symbols in their
work, but rather explained everything in words. Nevertheless, their
stepwise approach was very similar to the one advocated in this book.
Al-Khwarizmi is also remembered for his work on the solution of
quadratic equations, discussed later in this chapter.
A little less working is shown in Worked examples 5.8, 5.9 and 5.10 than
previously, and hints are not explicitly referred to. This has been done so as to
make the working more akin to what you might reasonably write when
working through the questions in this book. You are encouraged to show as
many steps as necessary in your working, and to use words of explanation
wherever they help you.
ΔGm˜ + T ΔS m˜ = ΔH m˜
Subtracting ΔGm˜ from both sides gives
T ΔS m˜ = ΔH m˜ − ΔGm˜
Dividing both sides by T gives
ΔH m˜ − ΔGm˜
ΔS m˜ =
T
100
5 Algebra
t 2
Fgr2 = Gm1m2
Dividing both sides by Fg gives
Gm1m2
r2 =
Fg
Gm1m2
r=±
Fg
L = 4πR 2σ T 4
(5.1)
where σ (the lower case Greek letter sigma) represents a constant known
as Stefan’s constant, with the value σ = 5.67 × 10−8 W m−2 K−4.
It is impossible to take direct readings for the luminosity, radius or
temperature of distant stars, but indirect measurements can lead to values
for photospheric temperature and luminosity. Figure 5.3 is a so-called
Hertzsprung–Russell diagram, comparing the photospheric temperatures
and luminosity of different stars. Note that different types of stars appear
in distinct groupings on the Hertzsprung–Russell diagram.
101
Maths for Science
4πR 2σ T 4 = L
Dividing both sides by 4πσT4 gives
L
R2 =
4πσ T 4
(Note that the same results would have been achieved by dividing by 4,
π, σ and T4 separately.)
Taking the square root of both sides gives
L
R=±
4πσ T 4
(Since R is the radius of a star, you only need to consider the positive
value.)
high luminosity
supergiants
luminosity in W
Alcyone red
giants
Sirius A
m
ai
n
se
qu Sun
en
ce α Centauri B
Sirius B
low luminosity
white dwarfs
102
5 Algebra
3.2 × 1029 W
R=
4π × 5.67 × 10−8 W m −2 K −4 × (1.2 × 104 K) 4
3.2 × 1029 W m 2 K 4
=
4π × 5.67 × 10−8 W × (1.2 × 104 ) 4 K 4
3.2 × 1029 m 2 K 4
=
4π × 5.67 × 10−8 × (1.2
2 × 104 ) 4 K 4
= 2.17 × 1019 m 2
= 4.7 × 109 m
The radius of Alcyone is 4.7 × 109 m.
Notice that in this example, the units of watts cancelled without having
to be expressed in SI base units.
Question 5.4
103
Maths for Science
2 5 2 7 5
÷ = × (turning the upside down and multiplying)
3 7 3 5 7
2×7
=
3× 5
14
=
15
you can write
a c a d c
÷ = × (turning the upside down and multiplying)
b d b c d
a×d
=
b×c
ad
=
bc
Worked example 5.11 illustrates a division in which several of the terms
cancel.
2ab 2
Simplify ÷
c c
Answer
2
Turning the upside down and multiplying gives
c
2ab 2 2ab c
÷ = ×
c c c 2
Then the ‘2’s and the ‘c’s cancel to give
2ab 2 2 ab c
÷ = × = ab
c c c 2
104
5 Algebra
The method described in Section 1.2.2 for adding and subtracting numerical
fractions can also be extended to algebraic fractions. It is necessary to start by
finding a common denominator (which may be the product of the
denominators of the separate fractions). So just as
2 4 2 × 5 4 × 3 10 12 10 + 12 22
+ = + = + = =
3 5 3 × 5 5 × 3 15 15 15 15
you can write
a c ad cb ad + cb
+ = + =
b d bd db bd
since multiplication is commutative so db = bd.
105
Maths for Science
2c a (b + 2)
×
(a + 2) 2c b
2c a (b + 2) a (b + 2) a (b + 2)
× = =
(a + 2) 2c b (a + 2) b b (a + 2)
It can be tempting to ‘cancel’ the square roots and the ‘+2’s too, but this
would be incorrect:
a a (b + 2) b
≠ and ≠
b b (a + 2) a
a (b + 2) 2c a (b + 2)
So is as far as it is possible to simplify × .
b (a + 2) (a + 2) 2c b
a a a (b + 2)
Note however that is equivalent to , so could also
b b b (a + 2)
a (b + 2)
be written as .
b (a + 2)
106
5 Algebra
Question 5.5
μ0 i1i2 3a 2b 3c
(a) × (b) 2 (c) +
2π d 2b c b
2ab 2ac 1 1 2b 2 2c 2
(d) ÷ (e) − (f) ÷
c b f f +1 (b + c) (a + c)
Question 5.6
The distance, u, of an object from a lens (such as the lens in the simple
camera illustrated in Figure 5.5) is related to the distance, v , from the lens
to the image of the object and the lens’s focal length, f, by the equation
1 1 1
+ =
u v f
1 1
Add and and hence rearrange the equation to make f the subject.
u v
lens
image
u v of object
object
Figure 5.5 The object and image in a simple camera (for Question 5.6).
(a + b)(c + d)
107
Maths for Science
(x − 3)(x + 5) = x 2 + 5x − 3x − 15
= x 2 + 2x − 15
(b) Multiplying the terms gives
(x + y)(x − y) = x 2 − xy + yx − y 2
= x 2 − y2 since xy = yx , so −xy + yx = 0
(c) Squaring (x + y) and then multiplying the terms gives
(x + y) 2 = (x + y)( x + y)
= x 2 + xy + yx + y
2
= x 2 + 2xy + y 2
(x − y) 2 = (x − y)( x − y)
= x 2 − xy − yx + y
2
= x 2 − 2xy + y 2
(x + y)2 ≠ x2 + y2
(x − y)2 ≠ x2 − y2
108
5 Algebra
Question 5.7
1 (a − b) − (a − c)
(a) (vx + u x ) t (b) (c) (k − 2)(k − 3) (d) (t − 2)2
2 2
So far, this section has discussed removing brackets from expressions, but it
can very often be useful to do the reverse.
The numbers 6 and 4 are described as factors of 24 and in general, when
speaking mathematically, ‘factors’ are terms which when multiplied together
give the original expression.
Since, for example,
y(y + 3) = y2 + 3y
you can say that y and (y + 3) are factors of y2 + 3y
Similarly, since
(x + 3)( x − 1) = x 2 − x + 3x − 3
= x 2 + 2x − 3
you can say that (x + 3) and (x − 1) are factors of x2 + 2x − 3.
The verb ‘to factorise’ means to find the factors of an expression. If you are
asked to factorise y2 + 3y then you should write
y2 + 3y = y(y + 3)
and if you are asked to factorise x2 + 2x − 3 you should write
x2 + 2x − 3 = (x + 3)(x − 1)
Note, from Worked example 5.13b, that the factors of x2 – y2 are (x + y) and
(x − y), i.e.
x2 – y2 = (x + y)(x − y) (5.3)
The difference of two squared numbers can always be written as the
product of their sum and their difference.
Question 5.8
109
Maths for Science
F = m(rω2 + g)
This can be reversed to give
m(rω2 + g) = F
Then dividing both sides by (rω2 + g) gives
F
m=
rω 2+g
Question 5.9
y2 + 3y = y(y + 3) (5.4)
110
5 Algebra
x2 + 2x − 3 = (x + 3)(x − 1) (5.5)
x2 + 2x − 3 = 12 + (2 × 1) − 3 = 1 + 2 − 3 = 0, as expected.
So the solutions of the equation x2 + 2x − 3 = 0 are x = −3 and x = 1.
ax2 + bx + c = 0
where a, b and c are numbers, will have solutions given by the quadratic
equation formula
−b ± b 2 − 4ac
x=
2a
If b2 > 4ac (i.e. b2 is greater than 4ac) then b2 − 4ac will be positive,
111
Maths for Science
−b
If b2 = 4ac then b2 − 4ac = 0 and there is only one solution, x = .
2a
If b2 < 4ac (i.e. b2 is less than 4ac) then b2 − 4ac will be negative. This
means that the solutions will involve the square root of a negative
number and hence will involve imaginary numbers. Such numbers were
mentioned in Chapter 1, but will not be considered further in Maths for
Science.
Worked example 5.16 demonstrates the use of the quadratic equation formula
in solving the equation that was solved by factorisation in Worked
example 5.15.
−b ± b 2 − 4ac
x=
2a
−2 ± 22 − [4 × 1 × (−3)]
=
2 ×1
−2 ± 4 − [−1
2]
=
2
−2 ± 16
=
2
−2 ± 4
=
2
So
−2 + 4 2 −2 − 4 −6
x= = = 1 or x = = = −3
2 2 2 2
The solutions can be checked in exactly the same way as in Worked
example 5.15.
The solutions of the equation x2 + 2x − 3 = 0 have again been found to be
x = −3 and x = 1.
112
5 Algebra
Question 5.10
(a) Use your answers to Question 5.7 parts (c) and (d) to solve the
(i) k2 − 5k + 6 = 0 (ii) t2 − 4t + 4 = 0
(b) Use the quadratic equation formula to check your answers to part (a).
E = hf (5.6)
c = fλ (5.7)
Answer
Rearranging Equation 5.7 to make f (the variable you are trying to eliminate)
the subject gives
c
f =
λ
Substituting this expression for f into Equation 5.6 gives
c hc
E = h× =
λ λ
113
Maths for Science
Answer
Since both equations are already given with Fg (the variable you are trying to
eliminate) as the subject, you can simply set the two equations for Fg equal to
each other:
Mm
mg = G
r2
You now need to rearrange to give an equation for r. First note that there is an
m on both sides of the equation, and dividing both sides of the equation by m
gives
M
g =G
r2
Multiplying both sides by r2 gives
gr2 = GM
Dividing both sides by g gives
GM
r2 =
g
Taking the square root of both sides gives
GM
r=±
g
Note that in Worked example 5.18, as in many of the other examples and
questions in this section, there are several equally valid ways of proceeding
(you may, for example, have started by rearranging
Mm
Fg = G
r2
to give an equation for r and then have substituted for Fg from Fg = mg).
However the solution should have been the same whichever method you used.
Question 5.11
Two (or more) different equations containing the same two (or more)
unknown quantities are called simultaneous equations if the equations must
be satisfied (hold true) simultaneously. It is usually possible to solve two
simultaneous equations by using one equation to eliminate one of the
114
5 Algebra
x+y=7 (5.8)
2x − y = 2 (5.9)
Answer
The approach used here is:
2 Substitute this into Equation 5.9 to give an equation for x alone, and so to
y = 7 − x (5.10)
2x − (7 − x) = 2
i.e.
2x − 7 + x = 2
or
3x − 7 = 2
Adding 7 to both sides gives
3x = 9, i.e. x = 3
Substitution of x = 3 into Equation 5.10 shows that
y=7−x=7−3=4
So the solution (i.e. the values for x and y for which both of the equations
hold true) is x = 3 and y = 4. You can check this by substituting the values
for x and y into the left-hand side of Equations 5.8 and 5.9.
115
Maths for Science
You could have arrived at the same result by using several other methods.
However there is only one correct pair of values for x and y.
Worked example 5.19 shows that in order to find two unknown quantities, two
different equations relating them are required. This is always true and by
extension:
Worked example 5.20 shows how four equations can be combined together in
a case where there are four unknown quantities (the total surface area, S, is
required, but the mass, m, and volume, V, of a single particle and the number
of particles, n, are unknown too and so must be eliminated). This worked
example concerns the use of metal particles as catalysts in the chemical
industry (see Box 5.5) and is included for interest only.
S = 4πnr2 (5.11)
The number of particles n is linked to the mass of one particle, m and the
total mass of metal, M by the equation
M
n= (5.12)
m
The mass m of one particle is linked to its volume V and the density of the
metal ρ by the equation
116
5 Algebra
m
ρ= (5.13)
V
The volume V of a particle is given by
4 3
V = πr (5.14)
3
where r is the radius.
Answer
Reversing Equation 5.13 gives
m
=ρ
V
Multiplying both sides by V gives
m = Vρ
Substituting for V from Equation 5.14 gives
4 3
m= πr ρ
3
Substituting this expression for m into Equation 5.12 gives
M
n=
m
⎛4 ⎞
= M ÷ ⎜ πr 3ρ ⎟
⎝ 3 ⎠
⎛ 3 ⎞
= M ×⎜ 3 ⎟
⎝ 4πr ρ ⎠
3M
=
4π r 3ρ
Substituting this expression for n into Equation 5.11 gives
S = 4πnr
2
3M
= 4π × × r2
4π r 3ρ
3M
= 4π × × r2
4π r 3ρ
3M
= × r2
r × r2 ρ
3Μ
=
rρ
117
Maths for Science
Question 5.12
118
6 Putting algebra to work
. Think of a number.
. Double it.
. Add 4.
. Subtract 1.
If you have arrived at an answer of 4, I can tell you that the number you first
thought of was 3; if your answer is 6, the number you first thought of was 5,
if your answer is 11, the number you first thought of was 10, and so on.
exactly the same operations for any number; if you represent the number by
• Think of a number N
• Double it 2N
• Add 4 2N + 4
• Halve your answer 1
2
(2N + 4) = N + 2
• Subtract 1 (N + 2) − 1 = N + 1
So the final answer will always be one more than the number you first thought
of.
119
Maths for Science
. Think of a number.
. Add 5.
. Subtract 2.
. Divide by 2.
Whichever number you first thought of, the answer will always be 4.
Question 6.1
You may wonder why a book entitled Maths for Science has suddenly started
discussing number tricks. There is a serious point to this, namely to illustrate
how you can get from an initial problem to a solution by using algebra.
Worked example 6.1 illustrates another use of algebra.
C = 3J (6.1)
Last year Chris’s age was (C − 1) and Jo’s age was (J − 1), so you can say
(C − 1) + (J − 1) = 78
i.e.
C + J − 2 = 78
C + J = 80 (6.2)
3J + J = 80
120
6 Putting algebra to work
i.e.
4J = 80
J = 20
Thus, from Equation 6.1, C = 3 × 20 = 60.
Thus Chris will be 60 this year and Jo will be 20. But this wasn’t the question
that was asked! When Jo was born, Chris was 60 − 20, i.e. 40 years old.
You may remember questions like Worked example 6.1 from your school
days. Problems like this can seem intimidating, but they are relatively easy to
solve once you have found the equations that describe the problem. Many
people struggle with this first step – they can’t find the equations to use. Look
at Worked example 6.1 carefully; all that has been done in order to derive
Equation 6.1 and Equation 6.2 has been to study carefully the information
given in the question, and to write it down in terms of symbols. So ‘On their
birthday this year Chris will be three times older than Jo’ has become C = 3J.
In solving problems, it is almost always helpful to start by writing down what
you already know. Drawing a diagram to illustrate the situation can help too;
you may find this helpful in Question 6.2.
Question 6.2
121
Maths for Science
3m = 4πr3ρ
Dividing both sides by 3 gives
4πr 3ρ
m=
3
Substituting numerical values
Note that symbols have been used for as long as possible, so as to avoid
numerical slips and rounding errors. However, it is now almost time to
122
6 Putting algebra to work
substitute the values given for r and ρ. First you need to convert the values
given into consistent (preferably SI) units:
ρ = 10.49 g cm −3
= 10.49 ×103 kg m −3
= 1.049 × 104 kg m −3 in scientiific notation
Then
4πr 3ρ
m=
3
4π × (2.5 ×10−3 m)3 ×1.049 ×104 kg m −3
=
3
4π × (2.5 ×10 ) m3 ×1.049 ×104 kg m −3
−3 3
=
3
= 6.8657 ×10−4 m3 kg m −3
= 6.9 ×10−4 kg to two significant figures
Is the answer reasonable?
It is always worth spending a few minutes checking whether the answer you
have arrived at is reasonable. There are three simple ways of doing this (it is
not normally necessary to use all three methods to check one answer):
Method 1
Check the units of the answer. Units have been given next to all the numerical
values in the calculation, and the units on the right-hand side of the equation
have worked out to be kilograms, as you would expect for mass.
If you had made a mistake in transposing the formula for mass, and had
written it as
4πr 2 ρ
m=
3
by mistake, then the units on the right-hand side would have been
m2 × kg m−3 = kg m−1
These are not units expected for mass by itself, so you would have been
alerted to the fact that something was wrong.
Checking units in this way provides a good way of checking that you have
written down or derived an equation correctly; the units on the left-hand side
of an equation should always be equal to the units on the right-hand side. You
can use this method for checking an equation even if you are not substituting
numerical values into it.
123
Maths for Science
Method 2
Estimate the value (in the way described in Section 3.3), and compare it with
the answer found on a calculator. In this case
124
6 Putting algebra to work
Worked example 6.3 shows the use of these tips in solving a different
problem, concerning the conservation of energy. This worked example uses
formulae given in Box 6.1; you may also find these formulae useful when
answering Question 6.3.
(a) (b)
Figure 6.1 Some energy conversions: (a) a child on a slide; (b) water
boiling in an electric kettle.
In both cases some energy is ‘lost’ to other forms (such as heat to the
surroundings and sound) but very often you can assume that all of the
energy initially in one form is converted into just one other form, and so
equate formulae such as those given below for different forms of energy.
Note that some of these are repeated from Box 3.1, for your
convenience. All forms of energy should be quoted using the SI unit of
energy which is the joule (J), where 1 J = 1 kg m2 s−2.
125
Maths for Science
Ek = 12 mv 2
(6.3)
The gravitational potential energy, Eg, of an object of mass m at a height
Δh above a reference level is given by
Eg = mgΔh (6.4)
where g is the acceleration due to gravity.
The heat energy, q, needed to raise the temperature by ΔT for a
substance of mass m and specific heat capacity c is given by
q = mcΔT (6.5)
mcΔT = mgΔh
Dividing both sides by m gives
cΔT = gΔh
126
6 Putting algebra to work
9.81 × 4.8 m s −2 m kg K
ΔT =
5.0 ×102 kg m 2 s −2
9.81 × 4.8 m s −2 m kg K
=
5.0 ×102 kg m 2 s
−2
= 0.094 K to two significant figures
.
The answer could be stated in scientific notation as 9.4 × 10−2 K.
Is the answer reasonable?
In a real question you probably wouldn’t use all the tips for checking
described in Worked example 6.2 and the blue-toned box, but the answer
seems about the size you might expect (you wouldn’t expect a big temperature
rise) and the units have worked out to be kelvin, as expected for a change in
temperature.
Alternatively you could estimate the answer to be
10 m s −2 × 5 m
ΔT ≈
5 ×102 J kg −1 K −1
≈ 10−
1 K
This is the same order of magnitude as the calculated value, so the calculated
value seems reasonable.
127
Maths for Science
Question 6.3
A child climbs to the top of a 1.8 m slide and then slides to the ground.
Assuming that all of her gravitational potential energy is converted into
kinetic energy, find her speed as she reaches the ground. Take
g = 9.81 m s−2 and use appropriate formulae from Box 6.1.
Worked example 6.4 (which is based on the information in Box 6.2) returns to
a discussion of seismic waves travelling through the Earth’s crust. In this
example there are three unknown quantities (the distance, d, from the
earthquake, the time, tp, taken for P waves to reach the seismometer and the
time, ts, taken for S waves to reach the seismometer) so three equations must
be combined in order to find any of the unknown quantities. It is, therefore,
quite a complicated example but it has been included because it summarises
much of what has been discussed in Chapters 5 and 6, and also because it
illustrates the usefulness of algebra in science.
20 seconds time
128
6 Putting algebra to work
so
d
vp = (6.6)
tp
and
d
vs = (6.7)
ts
where d is the distance from the earthquake, tp is the time taken for P
waves to travel to the seismometer and ts is the time taken for S waves
to travel to the seismometer.
and
d
vs = (6.7)
ts
where vp = 5.6 km s−1 and vs = 3.4 km s−1, but d, tp and ts are all unknown,
so another equation is needed.
Although you don’t know the travel time of the two types of wave, you do
know that the difference in the arrival time of the two waves is 20 seconds, so
you can write
t = t s − tp (6.8)
where t = 20 s.
Equations 6.6, 6.7 and 6.8 are three equations containing the three unknowns
d, tp and ts and you need to combine and rearrange them to give an expression
for d.
Combining and rearranging equations
Multiplying both sides of Equation 6.6 by tp gives
tpvp = d
129
Maths for Science
vp
⎛ vp v ⎞
t = d⎜ − s
⎜ vsvp vpvs
⎟⎟
⎝ ⎠
d(vp − vs )
=
vsvp
Reversing the equation so that d is on the left-hand side gives
d(vp − vs )
=t
vsvp
20 s × 3.4 km s −1 × 5.6 km s −1
d=
(5.6 km s −1 − 3.4 km s −1)
20 s × 3..4 km s −1 × 5.6 km s −1
=
2.2 km s −1
= 1.7 ×102 km to two significant figures
130
6 Putting algebra to work
The units work out to be kilometres since the s in the numerator cancels with
one of the s−1 also in the numerator, and one of the km s−1 in the numerator
cancels with the km s−1 in the denominator, i.e.
s × km s −1 × km s −1
= km
km s −1
Is the answer reasonable?
The units have worked out to be kilometres as expected for a distance. If you
had converted the speeds into values in m s−1, you would have obtained a
value for d in metres (d = 1.7 × 105 m).
In this case it is easy to check that the answer is reasonable; many members
of the public reported a small earthquake on that day in Ambleside in the
English Lake District. Ambleside is about 170 km from Edinburgh
(Figure 6.3).
Edinburg
Edinburgh
Am
mb
mblesid
d
Ambleside
st
Belfast
Dublin Mancheste
Manchester
Birmingham
Cardiff London
Figure 6.3 The location of Edinburgh and Ambleside on a map of the British
Isles.
131
Maths for Science
Question 6.4
Use the equation E = mc2, and the fact that mass has SI units of kg whilst
speed has SI units of m s−1, to work out the base unit equivalent of the
joule (J).
132
6 Putting algebra to work
Provided you remember that ‘per’ or ‘per unit’ indicate division, this
definition follows straight from the equation
mass m
density = or ρ =
volume V
The reverse is also true: for quantities which are simply defined, the equation
can be written down straight from the definition. This is essentially the same
as you were doing in Section 6.1. In the same way as ‘on their birthday this
year Chris will be three times older than Jo’ became C = 3J, ‘density is mass
per unit volume’ becomes ρ = m/V. Units can be used to check that the
equation is reasonable.
Question 6.5
133
Maths for Science
the advice given in Section 6.2, ‘Tips for using algebra to solve scientific
problems’, provide a set of guidelines which will be useful in answering all
mathematical questions in your study of science.
Question 6.6
A student has given the answer shown in Figure 6.4 in response to the
following question:
Use the equation V = 43 πr 3 (Equation 3.5) to find the radius of a
planet with volume 6.09 × 1019 m3.
The answer shown is numerically correct, but it includes at least nine errors
(a) Identify the points in the answer that could be improved upon.
(b) Write out a better answer to this question.
134
6 Putting algebra to work
Question 6.7
135
7 Graphs and gradient
137
Maths for Science
500
ash
beech
hawthorn
willows
(5 species)
holly
oaks
(2 species)
sycamore
holm oak
Figure 7.1 Bar chart showing the number of herbivorous insect species
supported by some native and introduced tree species in the UK.
The willows, which are among the commonest tree species in the UK,
can support about 450 insect species. Sycamore, which is just as widely
distributed but came to this country more recently, supports only around
50 species, and the evergreen holm oak, which was introduced a mere
400 years ago, supports fewer than 10 insect species. However, you
should not generalise too much from these examples; there are other
native trees, such as holly, which support very few insect species.
138
7 Graphs and gradient
30
25
percentage of Earth’s surface
20
15
10
0
10 9 8 7 6 5 4 3 2 1 0 1 2 3 4 5 6 7 8 9 10
depth in km altitude in km
sea level
Figure 7.2 Histogram showing the percentage of the Earth’s surface lying within
specified intervals of height and depth, relative to sea level. The percentages of the
Earth’s surface above 5 km altitude and below 7 km depth are too small to show
at the scale of this diagram.
7.1.2 Graphs
Graphs can be used to show the way in which two quantities (one plotted on
the horizontal axis, sometimes referred to as the ‘x-axis’, and the other on the
vertical axis, sometimes referred to as the ‘y-axis’) vary in relationship with
each other. For example, in Figure 7.3, time is plotted along the horizontal
axis, with the years being evenly spaced. This graph shows the large variation
in caterpillar numbers that can occur from year to year, though no overall
trend can be discerned. (In a good year, the caterpillar hatch will be
synchronised with the emergence of new leaves on which the caterpillars can
feed, but variations in weather make the timing of bud burst uncertain. Hence
there are large variations from year to year in the number of caterpillars that
survive.)
It is not necessary to join the data points on a graph of this type; if this is
done, as here, the lines have no significance beyond simply emphasising the
downturn or upturn in the numbers between one year and the next.
The graphs in Figure 7.4 show average monthly temperatures in two locations.
Here the graphs indicate trends in the data; in both locations it is warmer in
July and August than in January and February, and there is a markedly
smaller variation in temperature in Paris (Figure 7.4a) than in Irgiz
(Figure 7.4b). However these graphs could not be used to predict the
temperature on any particular day. Figure 7.4b also illustrates that negative, as
well as positive numbers can be plotted on a graph; in this case the vertical
axis covers temperatures from −20 °C to +25 °C.
Note the labelling of the vertical axes in Figure 7.4. It is important that the
units of temperature are included, but if the units were included next to each
139
Maths for Science
600
550
25
20 20
temperature/°C
average
average temperature/°C
15 15
10 10
5 5
0
J F M A M J J A S O N D J F M A M J J A S O N D
month −5 month
−10
−15
−20
(a) (b)
Figure 7.4 Average monthly temperatures for (a) Paris and (b) Irgiz, Kazakhstan.
individual value the labelling would become messy. The vertical axis could be
labelled as ‘average temperature in °C’ or ‘average temperature (°C)’ but the
normal scientific convention is to write ‘average temperature/°C’. The ‘/’
indicates division; the temperature values are divided by their unit (in this
140
7 Graphs and gradient
case °C) to give pure numbers without units that can be plotted on the graph,
for example
2.3 °C
= 2.3
°C
The same convention that is used to label the quantities and units on the axes
of graphs is also used in the column headings of tables, and the convention
can be extended to include powers of ten, as shown in Table 7.1 and in
Figure 7.5. This can be a useful strategy in graph plotting because it allows
manageable numbers to be used in labelling the divisions on the axes.
Diameter/10−2 m Mass/10−3 kg
0.4 0.1
0.5 0.2
0.7 0.5
1.0 1.4
1.3 3.1
1.5 4.8
1.8 8.2
2.0 11.3
The first row of data in Table 7.1 (plotted as the lowest point on Figure 7.5)
describes an aluminium sphere with diameter 0.4 × 10−2 m (i.e. 4 × 10−3 m)
and mass 0.1 × 10−3 kg (i.e. 1 × 10−4 kg), where the ‘× 10−2 m’ and
‘× 10−3 kg’ are taken from column headings and axis labels. The quantities in
the table and graph have been divided not only by the units but also by a
power of ten. To obtain the actual value of a quantity you therefore need to
multiply by both the unit and the power of ten. Don’t forget to do this!
■ Use Table 7.1 to give the mass of an aluminium sphere whose diameter is
1.0 × 10−2 m.
□ Reading across from ‘1.0’ in the ‘Diameter/10−2 m’ column gives a value
of 1.4. This must be multiplied by 10−3 kg to give a mass of
1.4 × 10−3 kg.
141
Maths for Science
142
7 Graphs and gradient
a mass of 13.1 × 10−3 kg, but most people find it extremely difficult to
draw smooth curves freehand, so if you obtain a value between
12.8 × 10−3 kg and 13.4 × 10−3 kg you have done well.
This process of extending a graph beyond the highest or lowest data points, in
order to find corresponding values of the plotted quantities outside the original
range, is called extrapolation. Extrapolation is always particularly difficult in
regions where graphs curve, or have very steep or very shallow slopes. The
latter situation applies to Figure 7.5 in the region where the mass becomes
very small. It would be practically impossible to determine by extrapolation
the diameter corresponding to a mass of less than 0.1 × 10−3 kg. However,
when the diameter is zero, the mass will also be zero, so the curve must go
through the point at which the axes meet. On any graph the point at which
both plotted quantities are equal to zero is called the origin.
The fact that the graph in Figure 7.5 is curved makes both interpolation and
extrapolation less reliable than they would be if the graph was a straight line.
In Question 7.1 you can practise these processes using a graph that is easier to
deal with.
Question 7.1
Five measurements have been made to investigate the way in which the
voltage across the terminals of a power supply varies according to the
current flowing in the circuit. The data are plotted on Figure 7.6. (The
SI unit of voltage is the volt, symbol V; the SI unit of electric current is the
ampere, symbol A.)
(a) What is the value of the voltage when the current is (i) 1.5 A, (ii) zero?
(b) What is the value of the current when the voltage is (i) 3.4 V, (ii) zero?
5.0
4.0
voltage/V
3.0
2.0
1.0
143
Maths for Science
144
7 Graphs and gradient
ocean
Figure 7.8 shows a graphical representation of the data in Table 7.2. The line
has been drawn to go through the origin, the point at which age = 0 Ma and
distance = 0 km. This has been done because it is clear that newly formed
crust will not have moved any distance.
80
70
60
separation distance/km
50
40
30
20
10
0
1 2 3 4
age/Ma
Figure 7.8 Graph of data in Table 7.2. The black line represents the ‘best-fit’ to
the data.
Although it is obvious just from the table that the separation distance
increases with age, the graph immediately gives more information. First, the
145
Maths for Science
graph illustrates the relationship between the quantities plotted. In this case,
the points lie pretty much on a straight line and the relationship between the
age and the distance is thus said to be linear. Secondly, the graph provides a
good test of the reliability of the data. It is clear that there are no ‘rogue
points’ lying well off the straight line. However, the points do not all lie
exactly on a single line. The black line that has been drawn through them is
the best-fit line, i.e. the line that is most representative of the data as a whole.
The best-fit need not go through any of the data points, though it should be as
close as possible to as many of the points as possible, with approximately the
same number of points above and below the line.
Since the best-fit line is the best representation of the data as a whole, care
should be taken to read values from the line rather than from the data points.
■ Use the red lines on Figure 7.8 to give the separation distance of ocean
crust of age 3.4 Ma.
□ The crust of age 3.4 Ma is separated by 65 km.
146
7 Graphs and gradient
of the eleven data pairs in the table (i.e. v1 to v11 ) and then average all these
speeds. Plotting a graph therefore saves a tedious amount of calculation: using
the best-fit line allows vav to be calculated in a single step. In other words, a
graph provides a reliable way of averaging results.
■ What can you deduce about the spreading rate from the fact that all the
data points are close to the best-fit straight line, with some points lying
above and others below the line?
□ The rate of spreading has remained roughly constant over time. Again the
graph provides this information at a glance, whereas it would require a
lot of calculation to deduce it from the raw data in Table 7.2.
Another way of describing this process of calculating the spreading rate from
the distance–time graph is to say that you have determined the ‘slope’ or
gradient of the best-fit line. Figure 7.9 shows the analogy with the gradients
used to characterise steep hills, which you may have seen on road signs. The
gradient is defined in this context as the ‘rise’ (the total change in vertical
distance) divided by the ‘run’ (the total change in horizontal distance), so
rise
gradient =
run
100 m
=
300 m
/ / m
100
=
// m
300
1
=
3
= 0.33 or 33% to
two significant figures
‘rise’ =
100 m
‘run’ = 300 m
Note that in this particular case the gradient has no units, because it is
calculated by dividing a length by another length. In general, however,
gradients must, as with the example of Figure 7.8, be given their correct units.
In the case of a road, it is common to quote the gradient in the form of a
percentage (33% in the case of Figure 7.9). With a graph it is more usual to
quote the gradient as a number in decimal notation.
The gradient of a straight line is the same all the way along it, so any two
points on the graph can be used to define the rise and the corresponding run.
If, as is the case in Figure 7.10a where a quantity y is plotted against a
quantity x, the graph goes through the origin, it may be convenient to use that
147
Maths for Science
fact in calculating the gradient. Here the rise is (y2 − 0) and the run is
(x2 − 0), so
rise
gradient =
run
y − 0
= 2
x2 − 0
y
= 2
x2
This was effectively the technique used in calculating the sea-floor spreading
rate from Figure 7.8, when just one point on the best-fit line was used to
calculate the speed. However, not all graphs go through the origin, so the
method illustrated by Figure 7.10a is not always applicable. Figure 7.10b
shows the most general method of determining the gradient of a straight-line
graph, which can be used whether or not the line goes through the origin.
y y
y2
y2
rise
rise = (y2 − y1)
= (y2 − 0)
y2
run = (x2 − x1)
0 run = (x2 − 0) x 0 x1 x2 x
x2
y2 − 0 y2 y2 − y1
(a) gradient = = (b) gradient =
x2 − 0 x2 x2 − x1
Figure 7.10 Finding the gradient of a straight-line graph: (a) for a straight line
going through the origin; (b) the general case, where the straight line does not
necessarily go through the origin.
Whatever points are chosen for determining the rise and run, it is always a
good idea to choose ones that are easy to read on at least one axis and
preferably on both axes! It is also good practice to choose points as widely
148
7 Graphs and gradient
emitted
electrons
metal
Figure 7.12 shows a graph arising from a photoelectric effect experiment. The
energy of the emitted electrons (measured in joules) is plotted on the vertical
axis and the frequency of the light (measured in s−1, or hertz) is plotted on the
horizontal axis. What is the gradient of this graph?
Answer
It is clear that even if the line were to be extrapolated to smaller values of
energy and frequency it would not go through the origin, so the method 10
shown in Figure 7.10b is the appropriate one to use in calculating the
energy/10−19 J
8
gradient.
6
From the lines drawn on Figure 7.12,
4
rise
gradient = 2
run
(9.2 × 10−19 J) − (2.6 × 10−19 J)
= 0 1 2
(2.0 × 1015 s −1) − (1.0 × 1015 s −1) frequency/1015 s−1
(9.2 − 2.6) × 10−19 J
=
(2.0 − 1.0) × 1015 s −1 Figure 7.12 A graph of data
from a photoelectric effect
6.6 × 10−19 J
= experiment on a particular
1..0 × 1015 s −1 metal, relating the energy of the
J ejected electrons to the
= 6.6 × 10−19−15 frequency of the light falling on
s −1 the metal. The red lines indicate
1 rise and run.
= 6.6 × 10−34 J s (remembering that = s)
s −1
149
Maths for Science
Question 7.2
The speed of seismic waves (see Box 4.1) can be calculated by measuring
the time for the waves to reach measuring instruments at different distances
from the epicentre of the earthquake. (Note that the focus is the point
within the Earth at which the seismic event takes place, and the epicentre is
the point on the Earth’s surface vertically above the focus.) Some typical
data from such a series of measurements on P waves are plotted in
Figure 7.13. Use the graph to calculate the average speed of the P waves in
km s−1 and to two significant figures.
distance from epicentre/km
200
100
0 10 20 30 40
travel time after earthquake occurred/s
Figure 7.13 Graph showing how long it takes for P waves from a shallow-focus
earthquake to reach three detectors at different distances from the epicentre. A
best-fit line has been drawn.
40 distance travelled on the vertical axis, so that the gradient would be equivalent
to the seismic wave speed. However, for this particular example, plotting the
graph this way round is not standard practice. The convention that scientists
30 follow is to plot on the horizontal axis the variable that is under their control.
Because they can choose the values of this quantity, it is called the
independent variable. In the case of the measurements described in
20 Question 7.2, there is a choice (within reason) of where the seismic wave
detectors are located; therefore distance from the epicentre is the independent
variable. The time taken for the P waves to arrive depends on where the
detectors have been positioned, so this is called the dependent variable.
10
According to the convention, the dependent variable is plotted on the vertical
axis. Figure 7.14 shows the same data as Figure 7.13, but replotted so that the
convention is followed.
0 100 200
distance from epicentre/km
The convention in science is to plot the independent variable on the
horizontal axis and the dependent variable on the vertical axis.
Figure 7.14 Graph showing the
same data as Figure 7.13 plotted
with distance from the epicentre The seismic wave speed can be calculated equally well from Figure 7.14 as
as the independent variable.
from Figure 7.13.
150
7 Graphs and gradient
■ In Question 7.2, you calculated the speed of the seismic wave in units of
km/s (or km s−1). How are the units s km−1 related to these units of
speed?
□ The units s km−1 and km s−1 are reciprocals, i.e.
1
= km s −1
s km −1
Question 7.3
Use Figure 7.14 to determine the average speed of the seismic waves.
Remember to use the correct units at each stage of your calculation. Does
your final answer agree with the value you obtained in Question 7.2?
25
20
depth/cm
15
10
5
0 1 2 3 4 5
time/hours
151
Maths for Science
For the line drawn in Figure 7.15, gradient is given, as before, by:
rise y2 − y1
gradient = =
run x2 − x1
If x1 is 1 hour and x2 is 4 hours, the corresponding values for y1 and y2 are
y1 = 20 cm and y2 = 5 cm, i.e. x2 is greater than x1 but y1 is greater than y2.
This means that:
(5 − 20) cm −15 cm
gradient = = = −5 cm hour −1
(4 − 1) hours 3 hours
In other words, the gradient is negative.
■ What physical meaning do you attach to the gradient in this context?
□ The graph shows that depth is decreasing with time – in other words the
snow is melting. The negative value of the gradient conveys this same
information. The gradient is constant over the time during which the
measurements have been made, so the snow is melting at a steady rate.
Now look at Figure 7.16, which shows the variation of distance, d, from a
given point with time, t, for objects moving in a variety of situations. A
scientific way to say this is that the graphs all show distance ‘as a function of’
time, or d as a function of t.
152
7 Graphs and gradient
Graphs like Figure 7.16, that by their shape show the nature of the
relationship between quantities but do not have scales marked on the axes, are
called ‘sketch graphs’. They can be very useful for illustrating ideas, without
the need for accurate plotting or drawing.
Question 7.4
Figure 7.17 shows the variation in temperature of the troposphere from sea
20
level to an altitude of about 2.5 km. Calculate the gradient of this graph,
temperature/˚C
giving your answer to two significant figures and taking care with signs and
10
Question 7.5
153
Maths for Science
This assumes that the cost per litre is constant however big the delivery. This
constant factor, which is required to turn the proportionality into an equation,
is called the constant of proportionality.
Now consider how this relationship between cost and volume appears on a
graph, such as that plotted in Figure 7.19. If you don’t buy any oil, the cost is
zero (but your heating won’t work!), so the graph must go through the origin.
If you buy 500 litres it costs £300, and if you buy 1000 litres it costs £600.
600
500
total cost/£
400
300
200
100
Figure 7.19 A graph of the total cost of heating oil against the volume delivered.
■ What is the gradient of this graph? What does that value represent?
□ The gradient is
The gradient represents the cost per litre. In this case this is £0.60/litre or
60 pence/litre. The gradient of the graph is the constant of proportionality
between total cost and volume of oil.
Generalising from this example:
y
If y = kx, where y and x are variables and k is a constant,
then y is said to be directly proportional to x, i.e.
gradient = k
y∝x
0 x
154
7 Graphs and gradient
Question 7.6
0 g
0 b
(a) (b) (c)
155
Maths for Science
600
500
total cost/£
400
300
200
100
Figure 7.23 A graph of the total cost of heating oil against the volume delivered,
for a company charging a delivery charge.
The point at which a line on a graph crosses an axis is called the intercept of
the line with that axis.
Generalising from this example, if two quantities y and x are related by an
equation of the form
y = mx + c (7.1)
where m and c are constants, then a graph of y against x will be a straight line
that does not go through the origin. The graph will have gradient m. And
when x = 0, then y = c, so the graph will have intercept c on the vertical axis.
This is illustrated in Figure 7.24.
y
The equation of a straight line is commonly written in the form
y = mx + c gradient of line
intercept of line
gradient = m y = mx + c with vertical axis
c
plotted on plotted on
vertical axis horizontal axis
0 x
Figure 7.24 A straight-line Although the general equation of a straight line is most usually written in the
graph with gradient m and form y = mx + c, it is important to remember that the letters used and their
intercept c on the vertical axis. order are quite arbitrary. For example, if an object has an initial speed ux and
156
7 Graphs and gradient
y = m x + c
vx = ax t + ux
12
10
gradient = 9.81 m s−2
vx
8
vx/m s−1
gradient = ax 4
ux
2
Note that, although the equation y = mx + c does not contain any minus signs,
both the gradient m and the constant c might have a negative value.
Examples of how the gradient and intercept of a straight line may be used to
derive quantities of real interest to scientists are given in Box 7.3 and Worked
example 7.2.
157
Maths for Science
y = m x + c
E = h f + (−φ)
gradient intercept
The gradient calculated in Worked example 7.1 (i.e. 6.6 × 10−34 J s) is
therefore the value of the Planck constant h, and extrapolation of the line
in Figure 7.12 to its intersection with the vertical axis can be used to
determine the work function of the metal. Figure 7.26 shows this
extrapolation. The intercept with the vertical axis is close to
−4 × 10−19 J, thus the work function of the metal is about 4 × 10−19 J in
this case.
10
6
energy/10−19 J
0
1 2
−2
−4
frequency/1015 s−1
Figure 7.26 A graph of E against f for the photoelectric effect (Figure 7.12
with the line extrapolated to the vertical axis).
158
7 Graphs and gradient
14
13
12
11
10
8
P/105 Pa
0 1 2 3 4 5 6 7 8 9 10
h/m
Answer
Comparison of P = ρgh for a graph of P against h with y = mx + c for a
graph of y against x shows that
y = m x + c
P = ρg h (+ 0)
So the gradient of the graph gives ρg and the intercept on the vertical axis is
at 0, as expected since the line passes through the origin.
You could also have found this by comparing P = ρgh with y = kx.
159
Maths for Science
Question 7.8
80
Figure 7.28 shows the variation of weight, WM, against mass, m, for various
70 objects on the Moon, a relationship also described by the equation,
WM = mgM, where gM is the magnitude of the acceleration due to gravity
60 on the Moon. Find the gradient of the graph and hence find the magnitude
50 of the acceleration due to gravity on the Moon, to two significant figures.
WM/N
40
30
7.4 Graphs of different shapes
The previous section showed that it is a relatively straightforward matter to
20
deduce the equation linking two variables when their relationship can be
10 represented by a straight-line graph. But of course not all the quantities of
interest in science are linearly related to one another. Suppose that plotting
0 10 20 30 40 50 one variable against another results not in a straight line but in a curve. How
m/kg could you then determine the relationship between the variables?
Imagine for example that you had taken a set of circular objects with radii
Figure 7.28 A graph of 1, 2, 3, … 6 cm and measured their respective areas. Had you plotted the area
weight, WM, against mass, m,
A as a function of radius r you would have obtained a graph like that in
for objects on the Moon.
Figure 7.29.
160
7 Graphs and gradient
A/cm2
80
This equation shows that A is not directly proportional to r, so you should not 40
have been surprised that plotting A against r did not give a straight line. In
fact, the curved shape of Figure 7.29 is characteristic of a relationship 0 2 4 6 8 10
involving the square of one of the quantities plotted. This particular shape is r/cm
called a parabola.
Figure 7.29 Areas A of
7.4.1 Straight-line graphs for non-linear equations circles plotted as a function of
their radii r.
In the case of the equation A = πr2, it is quite easy to see how the curve of
Figure 7.29 can be transformed into a straight-line graph. A is equal to r2
multiplied by a constant π. So although A is not directly proportional to r, it is
directly proportional to r2:
A ∝ r2
Therefore the result of plotting A against r2 is a straight line, as illustrated in
Figure 7.30. 120
A/cm2
■ Without measuring anything on the graph itself, can you state the value 80
of the gradient of the line in Figure 7.30? 40
□ Comparison with the standard equation for a straight line shows that
0 10 20 30 40 50
r 2 /cm2
y = m x + c
Figure 7.30 Areas A of
A = � r2 (+ 0) circles plotted as a function of
the squares of their radii r2.
so the gradient of the line is π.
Worked example 7.3 considers the equation linking the period of a pendulum
(the time for one complete swing) and its length. If you have ever regulated a
long-case (grandfather) clock, you will know that the length of the pendulum
determines the period and hence affects the accuracy with which the clock
keeps time.
L
T = 2π
g
161
Maths for Science
Answer
There are at least two equally valid ways to plot the data here. Since
L
T = 2π
g
4π2 L
T2 =
g
which can also be written as
4π2
T2 = L
g
L is the independent variable, which according to convention should be
plotted on the horizontal axis.
Comparison with y = mx + c or y = kx shows that a graph of T2 against L has
4π2
gradient =
g
so
4π2
g=
gradient
Alternatively, you could have chosen to plot T against L .
L
T = 2π
g
4π2
g=
(gradient) 2
162
7 Graphs and gradient
Question 7.9
A sketch graph showing the shape of a plot of V against P resulting from such
an experiment is shown in Figure 7.32. A plot of this shape is called a
hyperbola. A characteristic feature of the hyperbola is that as the variable on Figure 7.31 An apparatus for
one axis approaches zero, the curve approaches more and more closely to the measuring how the volume of a
other axis but never actually touches it. sample of gas varies with the
pressure at constant temperature.
A hyperbola arises from plotting two quantities that are linked by one being
directly proportional to the reciprocal of the other. In this case,
1
V ∝
P
163
Maths for Science
k 1
P= , i.e. P ∝
V V
So a graph of P against 1/V would also be a straight line, as illustrated in
Figure 7.34.
0 Note that graphs of 1/V against P and 1/P against V would also be
1/P
straight lines.
Figure 7.33 At constant Figures 7.32 to 7.34 correspond to a situation in which the temperature has
temperature, the volume of a been held constant. However, it would be equally possible to use the
fixed amount of gas is inversely apparatus illustrated in Figure 7.31 to measure the volume of the gas sample
proportional to the pressure. as a function of temperature. Such measurements are the basis of the SI
(kelvin) scale of temperature, which is discussed in Box 7.4.
P
Box 7.4 The absolute zero of temperature
Figure 7.34 shows that the pressure and volume of a fixed amount of gas
at constant temperature are related by an equation of the form
k
P=
V
0 1/V where k is a constant, i.e.
temperature, the pressure of a This equation is a particular case of a more general equation:
164
7 Graphs and gradient
P1
volume
P1
P2 P2
P3 P3
Figure 7.35 (a) At constant pressure, the volume of a fixed amount of gas
is clearly related to the temperature. Here P1 < P2 < P3. (b) Extrapolation
shows that when the volume is zero then the temperature is −273.15 °C.
⎛ temperature ⎞ ⎛ temperature ⎞
⎜ in ⎟ = ⎜ in degrees ⎟ + 273.15
⎜ kelvin ⎟ ⎜ Celsius ⎟
⎝ ⎠ ⎝ ⎠
165
Maths for Science
P1
P2
P3
0
T/K
166
7 Graphs and gradient
80
70
60
50
activity/kBq
40
30
20
10
A little further analysis shows that the time taken for the activity to drop:
. from 80 kBq to 40 kBq = 140 days
. from 40 kBq to 20 kBq = (280 − 140) days = 140 days
. from 20 kBq to 10 kBq = (420 − 280) days = 140 days.
This result demonstrates a very important property of the curve plotted in
Figure 7.37; whatever value of the quantity plotted on the vertical axis is
chosen, the time taken for the quantity to fall to exactly one-half that value is
a constant. This constant interval of time is known as the half-life, and curves
that display this property are called exponential decay curves. To the
precision to which it is possible to read Figure 7.37, the half-life of the
polonium sample is 140 days.
In radioactive decay, the activity is dependent on the number of radioactive
nuclei present, which is usually denoted by the letter N. During each half-life
(denoted by the symbol t1/2 on Figure 7.38) the number of radioactive nuclei
halves, thus if N0 radioactive nuclei are present when timing starts (i.e. at time
t = 0), then
. after one half-life N = N 0 × 12
. after two half-lives N = N 0 × 12 × 12 = N 0 ( 12 ) 2
. after three half-lives N = N 0 ( 12 ) 2 × 12 = N 0 ( 12 )3
so
. after n half-lives N = N 0 ( 12 ) n
After many half-lives, N will approach, though never reach, zero.
167
Maths for Science
N0
t1/2 = half-life
1
N1 = N0 × 2
1
N2 = N1 × 2
1
N3 = N2 × 2
The equation describing the exponential decay shown in Figure 7.38 involves
a special number, e. Like π and 2 , e is an irrational number and to four
significant figures its value is 2.718. The equation describing Figure 7.38 is
N = N0e−λt (7.3)
Question 7.10
Radium has a half-life of 1600 years. How long will it be before the
number of radioactive atoms in a sample is reduced to 161 of the number
there are today?
168
7 Graphs and gradient
n = n0e at (7.4)
where n0 is the starting value of the quantity, n is its value after time t and a
is a positive constant.
Exponential growth is sometimes used as a model by biologists interested in
the populations of organisms. Figure 7.39 illustrates the theoretical increase of
yeast cells according to such a model, in which the population consists of just
two cells at t = 0 and then once in every four-hour period each cell divides
into two. In practice, the death of organisms, as well as the influence of
factors relating to overcrowding, will also affect the population, so that the
increase in the number of organisms will not lie on a true exponential growth
curve.
169
Maths for Science
Question 7.11
14
12
10
8
L/10−2 m
0 20 40 60 80 100
m/10−3 kg
Figure 7.40 A graph of the overall length of a spring against the added mass.
(b) Find the gradient of the line shown in Figure 7.40, giving your answer
to two significant figures.
(c) The graph plotted in Figure 7.40 can be represented by the equation
L = L0 + Sm
Give the values of the constants L0 and S. What property of the spring does
L0 represent?
170
7 Graphs and gradient
171
8 Rate of change and differentiation
173
Maths for Science
tangent to
line 1
height of plant
point P
P line 2
time
(a) (b)
Figure 8.1 (a) A curve, representing the growth of a hypothetical plant. (b) The
tangent to a curve at a point P.
A tangent is a straight line, so its gradient can be found using the method
introduced in Section 7.2.1. Figure 8.2 is a graph of y = x2 and tangents have
been drawn at x = 1 and at x = 3.
y
18
16
14
12
10 tangent
at x = 3
8
4
tangent
2 at x = 1
1 2 3 4 5 x
Using the triangle drawn on the graph, the gradient of the tangent at x = 3 is
rise (15.0 − 9.0) 6.0
gradient = = = = 6.0
run (4.0 − 3.0) 1.0
Note that, because on this occasion x and y are variables without units, the
gradient also has no units.
174
8 Rate of change and differentiation
The gradient of the curve at a point is the same as that of the tangent touching
the curve at that point, so you can say that the gradient of the curve at x = 3
is 6.0 to two significant figures.
Question 8.1
2.0 × 10−3
1.0 × 10−3
175
Maths for Science
176
8 Rate of change and differentiation
Q1
Δy
Q2
tangent to
point P
Q3
P
Δx
2 The gradient of the chord gets closer and closer in value to the gradient of
the tangent at P.
Δy
As Δx approaches zero, the approximation approaches ever closer to the
Δx
exact gradient of the curve at the specified point. This situation is described as
Δy dy dy
a ‘limit’. In this limit, is written as where (said as ‘dee y by
Δx dx dx
dee x’) is called the derivative (or, strictly, the first derivative) of y with
respect to x.
dy
Note that should be regarded as a single symbol. It does not mean a
dx
quantity dy divided by another quantity dx, and the ‘d’s are not separate
quantities so they cannot be cancelled.
Differentiation is simply the process of finding a derivative. Box 8.2 shows
how this can be done from first principles for the example of y = x2. This box
is included for interest only; you do not need to be able to differentiate from
first principles. All you need to be able to do is to apply some very simple
general rules (the first of which is discussed in Section 8.2.2) that enable you
to find the derivative directly from the original equation. It turns out that, if
y = x2,
dy
= 2x
dx
177
Maths for Science
Δy
P
Δx
P could be any point on the curve and since it lies on the curve its x and
(y + Δy) = (x + Δx)2
Multiplying out the bracket on the right-hand side, in the way discussed
in Section 5.2.2, gives
y + Δy = (x + Δx)(x + Δx)
= x 2 + xΔx + xΔx + (Δx) 2
= x 2 + 2xΔx + (Δx) 2
Since y = x2, you can subtract y from the left-hand side and x2 from the
right-hand side to give
Δy = 2xΔx + (Δx)2
Dividing both sides by Δx gives
178
8 Rate of change and differentiation
Δy 2 xΔx + (Δx) 2
=
Δx Δx
= 2 x + Δx
In the limit as Δx approaches zero, the second term on the right-hand
Δy dy
side will disappear, and will become equal to so
Δx dx
dy
= 2x
dx
179
Maths for Science
Worked examples 8.1 to 8.6 show the application of this rule to find
derivatives and thus the gradients of graphs (at specific points or more
generally).
dy
If y = x5, what is and what is the gradient of a graph of y = x5 at x = 2?
dx
Answer
In this case C = 1 and n = 5, so
dy
= 1 × 5x 4 = 5x 4
dx
When x = 2,
dy
= 5 × 24 = 5 × 16 = 80
dx
So at x = 2 the gradient of the graph is 80.
dy
If y = 4x3, what is and what is the gradient of a graph of y = 4x3 at
x = 3? dx
Answer
In this case C = 4 and n = 3, so
dy
= 4 × 3x 2 = 12 x 2
dx
When x = 3,
dy
= 12 × 32 = 12 × 9 = 108
dx
So at x = 3 the gradient of the graph is 108.
180
8 Rate of change and differentiation
Worked example 8.3 considers the application of the rule for differentiation in
the special case when n = 1, and Worked example 8.4 considers what happens
when n = 0; you may like to think about what you expect the results to be.
dy
If y = 4x, what is ?
dx
Answer
In this case C = 4 and n = 1, so
dy
= 4 × 1x1−1 = 4x 0 = 4
dx
(since x0 = 1 for all values of x, as discussed in Section 1.3.1).
Note that y = 4x is the equation of a straight line through the origin, so the
result of Worked example 8.3 should not have surprised you; differentiating
an equation of the form y = kx will always result in a derivative which is a
constant. This constant is equal to the gradient, k, of a graph of y against x (as
discussed in Section 7.3.1).
dy
If y = 3, what is ?
dx
Answer
y = 3 can be written as y = 3x0 (since x0 = 1), so C = 3 and n = 0. Thus
dy
= 3 × 0 × x −1 = 0
dx
(since multiplying anything by 0 gives 0).
Differentiating a constant always gives zero. This should not surprise you
either, since the graph of y = 3 is a horizontal line and the gradient of a
horizontal line is always zero.
Note that the instruction ‘to differentiate’ in the following question and
worked example simply means that you should find the derivative.
Question 8.2
Differentiate the following with respect to x and in each case find the
gradient of the graph of y against x at x = 4.
(a) y = x4 (b) y = 5x (c) y = 3x2 (d) y = 5
The rule for differentiation used in Worked examples 8.1 to 8.4 applies for
negative and fractional values of n too, as illustrated in Worked examples 8.5
and 8.6.
181
Maths for Science
3
Differentiate y = with respect to x.
x
Answer
3
y= can be written as y = 3x−1 (see Section 1.3.1 for a reminder of the use
x
of negative exponents), so C = 3 and n = −1. Thus
dy 3
= 3 × (−1) x −1−1 = −3x −2 = − 2
dx x
Question 8.3
Differentiate the following with respect to x and in each case find the
1 2
(a) y = (b) y =
x x2
dy dy
Note that y and are both functions of x, i.e. the values of y and
dx dx
depend on the value of x. A derivative is sometimes called a derived function
because it is a function that has been derived from another function.
Functions in science are often expressed in terms of variables other than x and
y. For example, you may know that as time, t, changes, the distance, s, of an
object from a certain position varies according to the equation s = 5t2. The
graph of this function is illustrated in Figure 8.6.
182
8 Rate of change and differentiation
The speed at which the object is moving is given by the rate of change s
of distance with time, so to find the object’s speed you need to find the
gradient of the graph shown in Figure 8.6, i.e. to differentiate s with 120
respect to t.
ds 100
= 5 × 2t 2−1 = 10t1 = 10t
dt
80
Similarly, you know from Chapter 7 that the volume, V, of a gas at
constant temperature is inversely proportional to its pressure, P, i.e.
60
1 k
V ∝ or V = = kP −1
P P 40
where k is a constant.
20
Differentiating V with respect to P gives
dV k
= k × (−1)P −1−1 = −kP −2 = − 2 0 1 2 3 4 5 t
dP P
This expression gives the gradient of the graph shown in Figure 7.32.
Figure 8.6 A graph of s = 5t2.
Question 8.4
C dE
(b) If E = where C is constant, what is ?
r dr
An entirely different notation, called function or prime notation, is sometimes
used for derivatives. This notation makes it very clear that both the expression
being differentiated and its derivative are functions, and it identifies the
variable on which the functions depend. In this notation, the function shown
in Figure 8.6 would be written as f(t) = 5t2 and its first derivative would be
written as f ′(t) = 10t. The term f(t), usually said as ‘f of t’, does not mean f
times t, but simply implies that f is a function of t. The term f ′(t) (said as
‘f prime of t’) is the first derivative of f with respect to t.
dy
Unfortunately both f ′(x) and notation are in common use, as is a variation
of the latter which writes dx
d
(5t 2 ) = 10t
dt
for the derivative of 5t2 with respect to time.
dy
This book uses only notation as discussed in the preceding sections, but
dx
you should be aware that other notations are also widely used.
dy
The notation was invented by Gottfried Leibniz, one of the founders of
dx
calculus, and is known as Leibniz notation. Yet another notation, less
commonly used in modern times, writes sj = 10t for the first derivative of
183
Maths for Science
s = 5t2 with respect to t. This notation was first used by Newton, and the fact
that there is such a plethora of notations for differentiation is a lasting
reminder of the bitter dispute between Newton and Leibniz over which of
them invented calculus (see Box 8.3).
184
8 Rate of change and differentiation
0
1 2 3 4 x
−1
y = x2 − 4x + 3 (8.1)
185
Maths for Science
To show how this follows from the sum of the derivatives of the individual
terms, write y = x2 − 4x + 3 as y = u + v + w
where u = x2, v = −4
x and w = 3.
d
x
Since v = −4 x ,
d
v
= −4 (8.4)
d
x
and since w = 3,
d
w
= 0 (8.5)
d
x
Comparing Equation 8.2 with Equations 8.3, 8.4 and 8.5 shows that
dy
d
u
d
v
d
w
=
+
+
dx
d
x
d
x
d
x
dy
= 2x −
4
dx
When x = 2,
dy
=
(
2 ×
2) − 4 =
0
d
x
So the gradient of y = x2 − 4x + 3 at x =
2 is 0. This is consistent with
what you can see from Figure 8.8. At x = 2, a tangent drawn to the curve
would be a horizontal line.
186
8 Rate of change and differentiation
= 5t 4 +18t 2
Question 8.5
y = x2 − 4x + 3
and
dy
= 2x − 4
dx
dy
is itself a function of x and differentiating again gives
dx
d2 y
=2
dx 2
dy d2 y
The graphs of y against x, against x and against x for this example
dx dx 2
are shown in Figure 8.9.
dy
The graph of against x shows how the gradient of the graph of y against x
dx
varies with x.
187
Maths for Science
d2 y dy
The graph of against x shows how the gradient of the graph of
dx 2 dx
against x varies with x.
For y = x2 − 4x + 3, the graph of y against x (Figure 8.9a) is a parabola (as
discussed in Section 7.4).
dy
The graph of against x (Figure 8.9b) is a straight line of gradient 2.
dx d2 y
It should not surprise you then that has a constant value of 2, as shown
in Figure 8.9c. dx 2
dy
y dx
dy
y = x2 − 4x + 3 4 = 2x − 4
3 dx
3
2
2
1
x 1
0
1 2 3 4
x
0
−1 1 2 3 4
−1
(a)
−2
d 2y
dx2 d2 y
=2 −3
dx2
2
−4
1
(b)
x
0
1 2 3 4
(c)
Figure 8.9 (a) A graph of y = x2 − 4x + 3 and (b) its first derivative and (c) its
second derivative.
188
8 Rate of change and differentiation
d2 y
= 2a
dx 2
Question 8.6
sx = u xt + 12 axt 2
(Box 3.1 Equation 3.17)
where sx, ux and ax are respectively distance, initial speed and
acceleration, all in the direction of the x-axis (which in this case is taken
as pointing upwards from the point at which the ball was thrown) and t
is time.
On this occasion, ax = −9.81 m s−2. The acceleration is due to gravity,
and the negative sign indicates that the acceleration due to gravity acts
downwards whilst the x-axis points upwards.
Suppose the ball is thrown upwards with an initial speed of
ux = 6.0 m s−1. Box 3.1 Equation 3.17 then becomes
189
Maths for Science
sx /m
1
0
0.5 1.0 1.5
t/s
(a) (b)
6.0
ax /m s−2
vx /m s−1
0 0
0.5 1.0 1.5 t/s
t/s
−9.81
(c) (d)
Figure 8.10 (a) A ball being thrown into the air. Graphs to show the
variation of (b) distance, (c) speed and (d) acceleration with time for a ball
thrown into the air.
dsx
vx =
dt
= (6.0 m s −1 × 1t1−1) − ( 12
× 9.81 m s −2 × 2t 2−1)
= (6.0 m s −11 × t 0 ) − (9.81 m s −2
× t)
= 6.0 m s −1 − (9.81 m s −2 )t
As shown in Figure 8.10c, the graph of vx after time t is a straight line
with an intercept on the vertical axis at 6.0 m s−1 and a gradient of
−9.81 m s−2. Again as predicted, the speed decreases from its initial
value of 6.0 m s−1, becoming negative as the ball falls back to Earth.
Differentiating Equation 8.6 for a second time gives the rate at which the
object’s speed is changing. This is the object’s acceleration, ax
190
8 Rate of change and differentiation
dvx
ax =
dt
d 2sx
= 2
dt
= 0 − 9.81 m s −2
= −9.81 m s −2
The acceleration has a constant value of −9.81 m s−2, again as expected.
The fact that the final answer is what you would expect provides a useful
check of Box 3.1 Equation 3.17 (which was assumed to be the correct
equation from which to start).
Figure 8.10 shows the variation of the ball’s distance sx, speed vx and
acceleration ax, in the upwards direction and with increasing time. Note
that the gradient of the first graph (sx against t) leads to the second graph
( vx against t) and that the gradient of the second graph ( vx against t)
leads to the final graph (ax against t).
191
Maths for Science
y
20
18
16 tangent
14
12
10
4
tangent
2
tangent
−3 −2 −1 0 1 2 3 x
In each case the gradient of the tangent (and thus of the graph itself) is
equal to the value of y at the point where the tangent was drawn.
192
8 Rate of change and differentiation
The rule that has emerged from this sequence is generally true; the gradient of
a graph of y = ex at a particular point is equal to the value of y at that point,
i.e. for y = ex, the derivative of y with respect to x is equal to y itself:
If y = ex, then
dy
= ex = y
dx
dy
is only equal to y for certain specific exponential functions. However, in
dx
similar fashion to that followed in Section 8.2.2, it is possible to draw out a
more general rule. It can be shown that
. if y = 2ex, then
dy
= 2e x = y
dx
. if y = 3ex, then
dy
= 3e x = y
dx
and so on. So, in general, the derivative of y = Cex with respect to x, where C
is a constant, is
dy
= Ce x = y
dx
Similarly, it can be shown that:
. if y = Ce2x, then
dy
= C × 2e 2 x = 2y
dx
. if y = Ce3x, then
dy
= C × 3e3x = 3y
dx
and so on. So the general rule for all exponential functions is:
193
Maths for Science
This rule is the final rule for differentiation given in this chapter, and its use is
illustrated in Worked examples 8.8 and 8.9.
dy
The fact that is proportional to y for all functions of the form y = Cekx
dx
explains the shape of graphs of exponential growth and decay.
dy
If y = e4x what is ? Express your answer (a) in terms of x, and (b) in terms
of y. dx
Answer
(a) In this case C = 1 and k = 4, so
dy
= 1 × 4e 4x = 4e 4x
dx
(b) Alternatively, since y = e4x, you could write
dy
= 4e 4x = 4y
dx
d2z
= −6 × (−2)e −2t = 12e −2t
dt 2
Since z = 3e−2t, this could also be written as
d2z
= 4z
dt 2
194
8 Rate of change and differentiation
Question 8.7
dy
Since is proportional to y for all exponential functions, exponential
dx
functions can be used to describe situations in which the rate of change
of some quantity at an instant is proportional to the actual value of that
quantity at the same instant.
N
N0
195
Maths for Science
Question 8.8
196
9 Angles and trigonometry
(a)
Earth Moon
(b)
Figure 9.1 Chapter 9 will show how to use angles to find (a) the height of a tree,
and (b) the diameter of the Moon (not to scale).
Section 9.1 describes two different systems used for measuring angles and,
after a brief look at some of the properties of triangles, the rest of the chapter
shows how angles can be used in scientific calculation to determine things
such as the height of a tree and the diameter of the Moon.
197
Maths for Science
90
80
100
70
110
60
0
12
50
0
13
0 40
14
0 30
15
20
160
10
170
0
direction
135° of rotation
90°
180°
45° starting
position
225°
360°
198
9 Angles and trigonometry
Box 9.1 describes the use of angles to define lines of longitude and latitude
on the Earth’s surface, and hence to specify positions on the surface of the
Earth.
North Pole
North 90° N
lines of Pole lines of
latitude longitude 60° N
latitude
90° 30° N north
60°
Equator
30°
Equator
30°
60°
30° S
90°
South latitude
Pole south
60° S
90° S
(a) (b) South Pole
Figure 9.4 (a) A model of the Earth viewed from above the Equator.
(b) Using angles to label lines of latitude.
180°
150° W 150° E
120° W 120° E
90° W 90° E
60° W 60° E
North
30° W 30° E Pole
Greenwich 0°
Meridian
longitude west longitude east
Figure 9.5 A model of the Earth viewed from above the North Pole,
showing lines of latitude and longitude.
199
Maths for Science
In Figure 9.5, which is the view from above the North Pole, the circles
are the lines of latitude and the lines radiating out from the pole are lines
of longitude. It is easy to see, from Figure 9.5, how angles of longitude
can be labelled using degrees. A line running through Greenwich in
London, and known as the Greenwich Meridian, is defined to be
0° longitude, and other lines are labelled by measuring the angles to the
east or west of the Greenwich Meridian.
Figures 9.4 and 9.5 show lines of longitude and latitude at 15° intervals only,
but in reality the lines can be drawn as close together as required, and so can
be used to specify a location very precisely. In order to specify a precise
location, degrees of longitude and latitude need to be subdivided in some way.
Historically this was done by dividing each degree into 60 minutes (or
‘minutes of arc’) in the same way as each hour is divided into 60 minutes (of
time). The symbol ‘′’ is used to represent minutes of arc. The longitude of
Heathrow Airport (approximately 30 km west of Greenwich) is 0°27′ W and
both Greenwich and Heathrow have a latitude of about 51°28′ N.
Minutes of arc are rarely used in modern science; small angles are usually
expressed in decimal notation. Since 28′ is 28 sixtieths of a degree and
28
60
= 0.47 to two significant figures, 51°28′ can be written as 51.47°.
However, astronomers continue to use a further extension of the ‘degrees and
minutes’ notation, simply because the angles they are measuring are
frequently very small (since the objects they are measuring are such a long
way from Earth). In order to measure such small angles, minutes of arc are
further divided into 60 seconds of arc, or arcsecs (in the same way as minutes
of time are subdivided into 60 seconds). So
1 1 1 1
1 arcsec = minute of arc = × degree = degree
60 60 60 3600
As the Earth orbits the Sun, the next nearest star, Proxima Centauri, appears
to move through an angle of 0.772 arcsecs across the sky. This is
0.772
degree = 2.14 ×10−4 degree
3600
Angles in science are frequently measured in radians rather than in degrees
and subdivisions of degrees. Consider the circle shown in Figure 9.6. A part
X
of the circumference, such as that between point X and point Y, is known as
r an arc, and in this case the arc subtends an angle θ. The length of the arc
between X and Y is s and the radius of the circle is r. The radian is defined
θ s
with reference to arc length and radius.
r
200
9 Angles and trigonometry
s
θ (in radians) =
r
3.0 cm
=
2.0 cm
= 1.
5
So the size of the angle is 1.5 radians.
Note that since a length in centimetres has been divided by another length in
centimetres, it could be argued that the answer should have no units.
However, this book will adopt the common practice of writing the word
‘radians’ next to angles given in this measuring system, to distinguish them
from angles measured in degrees or in any other system of angular measure.
An angle subtended by a longer arc in a circle of the same radius will be
larger, as expected. In the above example, an arc of length 5.0 cm would
subtend an angle of
5.0 cm
= 2.5 radians
2.0 cm
However, it is the ratio of arc length to radius that is important in the
definition of a radian. This is illustrated in Figure 9.7, which shows two
concentric circles (i.e. two circles with their centres at the same point). The
smaller circle has radius r, and an arc of length s is shown, subtending an
r
angle θ. The larger circle has radius r′, and the same angle θ is subtended by
an arc of length s′. The superscript ‘′’ is used to indicate that the lengths r′ θ s s′
and s′ (said as ‘r-prime and s-prime’, or ‘r-dash and s-dash’) both relate to the
same circle. (Note: this convention of using primes has nothing to do with the r′
use of primes to indicate minutes of arc and neither of these uses has anything
to do with the f ′(x) notation used for the first derivative of a function.)
The lengths s′ and r′ are bigger than the values of s and r, as you would
expect, but the ratios s/r and s′/r′ are equal, and the angle subtended in
Figure 9.7 Two concentric
radians is
circles.
s s′
θ = =
r r′
Now consider two special cases. In the first, the arc length is exactly equal to
the radius, as shown in Figure 9.8a, i.e. s = r. This means that
s r
θ (in radians) = = =1
r r
i.e. the angle subtended is one radian.
201
Maths for Science
r
θ r r θ s
r
(a) (b)
Figure 9.8 The angle subtended when (a) arc length is equal to radius, and (b)
arc length is equal to circumference.
In the second special case, illustrated in Figure 9.8b, the arc length is a
complete circumference. For all circles, the circumference, C, is linked to the
radius, r, by the formula C = 2πr. This formula, given in Box 3.1, follows
directly from the definition of π, given in Section 1.1.1 as circumference
divided by diameter. So when the arc length, s, is equal to the whole
circumference 2πr
2πr
θ (in radians) = = 2π
r
360°
1 radian =
2π
180°
=
π
≈ 57.3°
where the symbol ‘≈’ means ‘is approximately equal to’, as in Chapter 3.
Similarly, since 360° = 2π radians, then
2π
1° =
360
π
=
180
≈ 0.0175 radians
Note that the numerical conversion factors between radians and degrees are
only approximate (they have been given to three significant figures), so when
converting from radians to degrees or vice versa it is best to go back to first
202
9 Angles and trigonometry
Question 9.1
2π 3π
(a) 0.123 radians (b) radians (c) radians
3 2
Question 9.2
203
Maths for Science
α
α
α
β γ β γ β γ
(a) (b) (c)
β γ β α β
(d) (e) (f)
204
9 Angles and trigonometry
b h
h = a 2 + b2
= (3 cm) 2 + (4 cm)
2
= 9 cm 2 + 16 cm 2
= 25 cm
2
= 5 cm
205
Maths for Science
b = h2 − a 2
= (9.1 m) 2 − (5.1 m) 2
= 82.81 m 2 − 26.01 m 2
= 56.80 m 2
= 7.5 m to
t two significant figures
Question 9.3
Hint: you may find it helpful to start by drawing a diagram of the situation.
206
9 Angles and trigonometry
h′′
b′′
h′
b′
h
b
θ θ θ
a a′ a′′
b b′ b′′
= = (9.4)
h h′ h′′
a a′ a′′
= = (9.5)
h h′ h′′
Question 9.4
Verify the results given in Equations 9.3, 9.4 and 9.5 by measuring the
lengths of the sides of the triangles shown in Figure 9.13 and calculating
the following ratios:
b b′ b′′ b b′ b′′ a a′ a′′
, , , , , , , ,
a a′ a′′ h h′ h′′ h h′ h′′
If the angle θ and hence the shape of the triangle had been different, the ratios
would have had different values. Thus each angle θ gives rise to unique
values for b/a, b/h and a/h, and conversely each value for b/a, b/h or a/h in a
triangle leads to a particular value for θ. This result is so important that the
ratios are given the special names tangent, sine and cosine, usually
abbreviated to tan, sin and cos. Tan, sin and cos are known collectively as
trigonometric ratios (or trig. ratios).
Note that the meaning of the word ‘tangent’ here is different from the one
used in Chapter 8 (the tangent to a curve). Here the tangent of angle θ is
defined by
opposite
tan θ = (9.6)
adjacent
This is the ratio previously described as b/a, where b is the side opposite
angle θ and a is the side (other than the hypotenuse) that is adjacent (next to)
angle θ.
207
Maths for Science
opposite
sin θ = (9.7)
hypotenuse
adjacent
cos θ = (9.8)
hypotenuse
opp
tan θ =
adj
hyp
opp opp
sin θ =
hyp
θ
adj
adj cos θ =
hyp
Note that the trigonometric ratios are defined with respect to a particular angle
in a right-angled triangle. If the other non right-angled angle in the triangle in
Figure 9.14 had been considered, the ‘opposite’ and ‘adjacent’ sides would
have been different, and so the sine, cosine and tangent would have been
different too.
The trigonometric ratios were tabulated many years ago and generations of
scientists used tables and slide rules similar to those shown in Figure 9.15 to
calculate lengths from angles and angles from lengths. Nowadays,
trigonometric ratios are available at the press of a calculator button.
208
9 Angles and trigonometry
209
Maths for Science
Note that when using trigonometric ratios you should always work to at least
four significant figures (although you should round your answer to an
appropriate number of significant figures at the end of a calculation).
Question 9.5
π π
(a) sin 49° (b) cos (c) tan
8 4
You will also need to be able to use your calculator to find the angle which
has a particular sine, cosine or tangent. For example, if you know that
tan θ = 0.75, then what is θ in degrees? What you are looking for is known as
the inverse tangent or arctangent and you need to use a button on your
calculator labelled as ‘tan−1’ or ‘arctan’. Check that you can use your
calculator to give the correct answer, which is that
Question 9.6
(a) Use your calculator to find the angle α (in degrees) for which
cos α = 0.5253.
(b) Use your calculator to find the angle β (in radians) for which
tan β = 1.5574.
Note that although trigonometric ratios have only been defined for angles in a
right-angled triangle, and most of the angles for which trigonometric ratios are
used in this book are acute (i.e. less than 90°), values of sin, cos and tan can
be found for larger angles too. Use your calculator to check that sin π = 0,
cos π = −1 and tan π = 0 (where π is an angle in radians, equal to 180°). Sine
and cosine functions, of the form y = a sin θ and y = a cos θ (where a is a
constant and the angle θ can take any angle) are extensively used in
describing the motion of waves. The detail is beyond the scope of this book,
but it is another application of maths in science!
210
9 Angles and trigonometry
θ
D
Figure 9.16 Using trigonometry to find the height of a tree. θ = 36.5° and
D = 28.6 m.
H = D tan θ
Substituting values for D and θ gives
H = 28.6 m × tan 36.5
°
= 28.6 m × 0.
7400
= 21.2 m to three significant fiigures
It was clearly stated at the beginning of the chapter that θ was the angle
between the ground and a straight line drawn to the top of the tree. But in
reality you are more likely to have taken readings at eye level, perhaps using
an instrument such as a ‘gun clinometer’, whose use is illustrated in
Figure 9.17. The gun clinometer measures the angle shown as α in
Figure 9.17b, and Worked example 9.3 shows how this can be used to find the
height of a tree.
211
Maths for Science
(a)
(b)
Figure 9.17 (a) Using a gun clinometer to find the height of a tree; (b) the gun
clinometer gives the angle α.
212
9 Angles and trigonometry
H = D tan α
= 18.0 m × tan 39.0°
= 18.0 m × 0.8098
= 14.6 m
On this occasion, however, the reading was taken at eye level, so H is not the
height of the tree. Assuming that it is 1.7 m from the ground to the man’s
eyes and that the ground is horizontal, the height of the tree is 1.7 m more
than H, i.e. the height of the tree is
Question 9.8 asks you to use trigonometry in solving another simulated ‘real
world’ problem, but Question 9.7 is given first to enable you to practise the
underlying trigonometric and algebraic skills.
Question 9.7
θ
1.0 m
5.0 m
(a)
10 m
a
h
32°
4.3 m
(b)
π/3
(c)
Figure 9.18 Right-angled triangles for use in Question 9.7 (not drawn to scale).
Question 9.8
213
Maths for Science
66 m
H
theodolite
1.5 m
θθ 15 m
ground ground
W surface W surface
θ θ
angle
of dip T V
(a) (b)
Figure 9.20 (a) The relationship between the angle of dip, θ, width of
outcrop, W, and true thickness, T, for a tilting stratum of rock (shown in
darker brown). (b) The relationship between θ, W, and the vertical thickness,
V, of the stratum.
214
9 Angles and trigonometry
The vertical thickness of the stratum (in Figure 9.20b) may also be of
interest, especially when exploring for underground resources (such as
oil) by drilling.
■ Express sin θ in Figure 9.20a in terms of T and W. Hence find an
equation for the true thickness, T, of a stratum in terms of the width,
W, of the outcrop at the Earth’s surface, and the angle of dip, θ.
□ sin θ can be expressed as
T
sin θ =
W
so
T = W sin θ (9.9)
V
tan θ =
W
so
V = W tan θ (9.10)
T = W sin θ
where W = 71 m and θ = 28°, the thickness is
215
Maths for Science
Question 9.9
r
45°
h
Figure 9.21 Using trigonometry to find the radius of iodide ions (shown in
purple). The small green sphere represents a lithium ion.
If the distance between the centre of a lithium ion and the centre of an
iodide ion is known (this is the so-called internuclear distance, and is
labelled h on Figure 9.21) then
adj r
cos 45° = =
hyp h
where r is the radius of an iodide ion.
Multiplying both sides by h gives
216
9 Angles and trigonometry
Question 9.10
v1
boundary v2
r
217
Maths for Science
Question 9.11
218
9 Angles and trigonometry
90°, as shown in Figure 9.23, the grating acts in such a way as to split
up the incoming beam, forming what is called a diffraction pattern. Some
light passes straight through the grating; this is called the zero-order
beam. Other beams are produced at angles θ1, θ2, etc. from the straight
through position and are known as the 1st, 2nd, etc. order diffracted
beams.
diffraction screen
grating second order
first order
θ2
light beam θ1
θ1 zero order
θ2
first order
second order
d sin θn = nλ
219
Maths for Science
Question 9.12
Another example of the difference between speed and velocity comes when
considering an object orbiting another object at constant speed. Consider, for
example, the Earth orbiting the Sun at about 30 km s−1 (as discussed in
Box 4.1). The Earth’s speed relative to the Sun is approximately constant, but
its direction of movement is constantly changing, so its velocity is constantly
changing too.
Many of the quantities considered elsewhere in this book (e.g. mass,
v
temperature and energy) are scalars, but velocity is not the only scientific
quantity to be a vector. There are many other vector quantities, including force
(a) (b) and acceleration.
A vector may be represented diagrammatically by an arrow, the length of
Figure 9.24 Representing a which specifies the vector’s magnitude, and the direction of which is the same
vector: (a) in printed text; (b) by as the vector’s. By convention, vectors are printed as bold symbols, for
hand. example v , while the magnitude of the vector is written normally, for
220
9 Angles and trigonometry
F = | F | = 10 N
The vertical lines drawn either side of the F provide an alternative way of
indicating the modulus (magnitude) of the vector.
Adding vector quantities together is not as straightforward as adding scalar
b
quantities, since both magnitude and direction need to be taken into account.
Fortunately trigonometry and Pythagoras’ Theorem can be used to help.
a
Imagine an object being acted on by the two forces shown in Figure 9.25.
You want to know the overall effect; what is the total force acting on the
object as a result of a and b? It is not immediately obvious how to proceed
since the two forces have different sizes and are acting in different directions.
One way forward is to resolve each vector into components; any two
dimensional vector (such as one drawn on the page of a book, as here) can be
characterised by its components along two perpendicular axes. Figure 9.26 Figure 9.25 Two
forces a and b acting
shows the components of the vector a along two axes x and y. The
on an object.
components ax and ay are scalar quantities and trigonometry can be used to
find them.
From the definition of cosine
y
adj
cos θ =
hyp
where on this occasion ‘adj’ = ax and ‘hyp’ = a (the magnitude of the
vector a) so ay a
a
cos θ = x
a
and rearranging gives θ
ax
x
ax = a cos θ
Similarly Figure 9.26 The x- and y
components of a.
opp
sin θ =
hyp
ay
=
a
and rearranging gives
ay = a sin θ
221
Maths for Science
= 3.
0 N
= 5.2 N
= 1.2 N
c 2 = cx2 + c 2y
c = cx2 + c 2y
cy c
= (5.5 N) 2 + (6.4 N)
2
= 8.4 N
φ
cx x The direction in which c acts is defined by the angle between c and the x-axis
(labelled φ in Figure 9.27).
= 1.1636
222
9 Angles and trigonometry
Question 9.13
Find the x- and y-components of the vector v shown in Figure 9.28. The y
vector has a magnitude of 8.6 m s−1 and acts at an angle, α, of 42° to the
x-axis.
Question 9.14
Find the magnitude and direction of the vector F shown in Figure 9.29.
vy v
Fx = 4.0 N and Fy = 3.0 N.
y α
vx x
Fy F
Fx x
■ Convert 0.5° into radians, again giving your answer to five significant
figures.
□ 360° = 2π, so
2π
1° =
360
2π
0.5° = 0.5 ×
360
= 8.7266 ×10−3 radians
223
Maths for Science
Comparing the answers to the above questions shows that sin θ ≈ θ and
tan θ ≈ θ, when θ is measured in radians, and also that cos θ ≈ 1. These
results are true for all small angles, in other words
cos θ ≈ 1
For small angles stated in radians,
These small angle approximations hold within 0.5% accuracy for angles less
than about 0.1 radians (6°). Remember though that the final two
approximations are only valid for angles measured in radians.
Question 9.15
For θ = 2.5°, use small angle approximations to estimate sin θ, cos θ and
tan θ.
Earth Moon
r
φ L D
s
r
Figure 9.30 Using small angle approximations to find the Moon’s diameter (not
to scale).
Equation 9.14, which can be used to find the distance to an object of known
diameter or the diameter of an object at a known distance, can be derived in
two ways.
224
9 Angles and trigonometry
In Figure 9.30, the tan of half the subtended angle is given exactly by
⎛ φ ⎞ opp
tan ⎜ ⎟ =
⎝ 2 ⎠ adj
1D
= 2
L
D
=
2L
φ φ ⎛φ ⎞ φ
is a small angle so when is in radians, tan ⎜ ⎟ ≈ . Therefore
2 2 ⎝2⎠ 2
φ D D
≈ i.e. φ ≈
2 2L L
Alternatively, from the definition of a radian (Equation 9.1)
s
φ (in radians) =
r
When the object is distant, φ is small, the arc length s is approximately equal
to the diameter D and the radius of arc r is approximately equal to the
distance to the object L, i.e. s ≈ D and r ≈ L so
D
φ≈
L
225
Maths for Science
Question 9.16
Figure 9.31 A car ferry observed from a beach (for Question 9.16) (not to scale).
226
9 Angles and trigonometry
Question 9.17
h
12 cm
Fy
F
θ θ
Fx
5.0 cm x
(a) (b)
Figure 9.32 (a) A right-angled triangle. For use in Question 9.17a. (b) The
vector F. For use in Question 9.17b.
(b) Find the x- and y-components of the vector F shown in Figure 9.32b.
The vector has a magnitude of 82 N and acts at an angle, θ, of 27° to the
x-axis.
(c) Viewed from Earth, the Sun subtends an angle of approximately 0.50°.
Given that the distance to the Sun is 1.50 × 1011 m, estimate its diameter.
227
10 Logarithms
Logarithms
10
‘Seeing there is nothing (right well-beloved Students of the Mathematics)
that is so troublesome to mathematical practice, nor that doth more molest
and hinder calculators, than the multiplications, divisions, square and cubical
extractions of great numbers, which besides the tedious expense of time are
for the most part subject to many slippery errors, I began therefore to
consider in my mind by what certain and ready art I might remove those
hindrances.’
Thus wrote John Napier in the preface to his book Mirifici logarithmorum
canonis descripio in 1614 (the quote is from the English translation of 1616).
Napier was a wealthy Scottish landowner and theologian, who claimed to
study mathematics only as a hobby. Despite this, he invented both logarithms
(or ‘logs’ for short) and ‘Napier’s bones’ (see Figure 10.1) with the express
purpose of making it easier to do multiplications and divisions. Logarithms
were in regular use for this purpose well into the second half of the 20th
century.
Figure 10.1 Napier’s bones are four-sided rods inscribed with the multiplication
tables. Their memorable name arose from the fact that some early versions were
made from ivory and so looked like bones.
229
Maths for Science
10−1 = = 0.1
10
10−5 = 5 = 0.000 01
10
where 10 is known as the base number.
The process of obtaining a logarithm to base 10 (usually described as ‘taking
the log to base 10’) is the inverse of raising the base 10 to a power. In each of
the above examples the logarithm to base 10 of the number on the right-hand
side of the equation is simply the power to which the 10 on the left-hand side
is raised. The logarithm to base 10 is abbreviated log10 in this book (you may
also see the abbreviation log, without a subscript, used to describe a logarithm
to base 10) so, for example,
100 = 1 so log10 1 = 0
More generally:
230
10 Logarithms
In fact, n could be any number; you may like to start by using your calculator
to check the following to four significant figures (use either the ‘xy’ or ‘^’
button or, if your calculator has one, a button marked ‘10x’):
100.1235 = 1.329
103.456 = 2858
10−1.234 = 0.058 34
For the last of these:
Question 10.1
Without further use of your calculator, write down the values of:
(a) log10 100, (b) log10 0.001, (c) log10 10 , (d) log10 1.329.
Question 10.2
Question 10.3
231
Maths for Science
⎛
I
⎞
where I0 is the intensity of the threshold sound and I is the intensity of the
sound in question. So the sound of a pneumatic drill with an intensity 1012
times that of the threshold has:
232
10 Logarithms
⎛ 1 × 10−7 mol dm −3 ⎞ −7
−log10 ⎜ −3 ⎟ = −log10 (10 ) = −(−7) = 7
⎝ mol dm ⎠
the pH of lemon juice is:
⎛ 8 × 10−3 mol dm −3 ⎞ −3
−log10 ⎜ −3 ⎟ = −log10 (8 × 10 ) = −(−2.1) = 2.1
⎝ mol dm ⎠
and the pH of household bleach is:
Question 10.4
233
Maths for Science
⎛ p⎞ ⎛ 2 ⎞
log10 ⎜ ⎟ = log10 ⎜ ⎟
q
⎝ ⎠ ⎝ 1000 ⎠
= log10 0.
002
= −2.6990 to four decimal places
234
10 Logarithms
Answer
log10 3000 = log10 (3 × 1000)
= log10 3 + log10 1000 (from Equation 10.2)
= 0.4771 + log10 103
= 0.4771 + 3 (from Equation 10.1)
= 3.4771 to four decimal places
Question 10.5
Use the fact that log10 3 = 0.4771, and Equations 10.1 to 10.4, to find the
following without using a calculator. Give your answers to four decimal
places.
(a) log10 300
(b) log10 0.03
(c) log10 9 (Hint: remember that 9 = 32)
These rules for the manipulation of logarithms explain how logarithms were
used to simplify the processes of multiplication and division. Equation 10.2
gives a way of turning multiplication into addition; Equation 10.3 gives a way
of turning division into subtraction; and Equation 10.4 gives a way of
calculating powers and roots.
The rules of logarithms have other uses too, as illustrated in Box 10.2.
⎛N ⎞
k = log10 ⎜ B ⎟
⎝ NA ⎠
where NB is the number of individuals alive before this stage and NA is
the number of individuals alive afterwards.
235
Maths for Science
For example, 43 eggs were laid (N2 in Table 10.1) but only 16 eggs
hatched (N3 in Table 10.1) so the k-value for this stage is:
⎛
N
⎞
⎛
43
⎞
k
3 =
log10 ⎜ 2 ⎟ =
log10 ⎜ ⎟ =
log10 (2.6875) =
0.4293
N
⎝
3 ⎠
⎝
16
⎠
Table 10.1 k-values for various stages in the breeding of 24 pairs of owls in
Wytham Wood in 1952–1953.
maximum number N1 = 51 ⎛N ⎞ ⎛ 51 ⎞
k2 = log10 ⎜ 1 ⎟ = log10 ⎜ ⎟ = 0.0741
of eggs from the N
⎝ 2⎠ ⎝ 43 ⎠
17 pairs that did breed
actual number N2 = 43 ⎛N ⎞ ⎛ 43 ⎞
k3 = log10 ⎜ 2 ⎟ = log10 ⎜ ⎟ = 0.4293
of eggs laid N
⎝ 3⎠ ⎝ 16 ⎠
number of eggs N3 = 16 ⎛N ⎞ ⎛ 16 ⎞
k4 = log10 ⎜ 3 ⎟ = log10 ⎜ ⎟ = 0.0280
that hatched ⎝ N4 ⎠ ⎝ 15 ⎠
number of chicks N4 = 15 ⎛N ⎞ ⎛ ⎞
k5 = log10 ⎜ 4 ⎟ = log10 ⎜ ⎟ = ..........
that fledged ⎝ N5 ⎠ ⎝ ⎠
number of N5 = 9 ⎛N ⎞ ⎛ 72 ⎞
k tot = log10 ⎜ 0 ⎟ = log10 ⎜ ⎟ = 0.9031
owlets that survived N
⎝ 5⎠ ⎝ 9 ⎠
to form pairs
■ Use the data in Table 10.1 to find k5. Enter your result in
Table 10.1.
□ Substituting N4 = 15 and N5 = 9 gives
⎛
N ⎞
k
5 =
log10 ⎜ 4 ⎟
⎝
N5 ⎠
⎛
15
⎞
=
log10 ⎜ ⎟
⎝
9
⎠
= log10 (1.6667)
=
0.2218
236
10 Logarithms
⎛N ⎞ ⎛ 72 ⎞
k tot = log10 ⎜ 0 ⎟ = log10 ⎜ ⎟ = 0.9031
N
⎝ 5⎠ ⎝ 9 ⎠
■ Use the data given in Table 10.1 to find k1 + k2 + k3 + k4 + k5
□ k1 + k2 + k3 + k4 + k5 = 0.1498 + 0.0741 + 0.4293 + 0.0280 + 0.2218
= 0.9030
Note that, within rounding errors, this is the same as the value
calculated for ktot from
⎛N ⎞
k tot = log10 ⎜ 0 ⎟
⎝ N5 ⎠
⎛N ⎞
k1 = log10 ⎜ 0 ⎟ = log10 N 0 − log10 N1 (from Equation 10.3)
⎝ N1 ⎠
Similarly
⎛N ⎞
k2 = log10 ⎜ 1 ⎟ = log10 N1 − log10 N 2
⎝ N2 ⎠
and so on, until
⎛N ⎞
k5 = log10 ⎜ 4 ⎟ = log10 N 4 − log10 N 5
⎝ N5 ⎠
So
k1 + k2 + k3 + k4 + k5 = (log10 N 0 − log10 N1) + (log10 N1 − log10 N 2 ) + …
+ (log10 N 4 − log10 N5 )
Apart from log10 N0 and log10 N5, all of the logarithms on the right-hand
side are both added and subtracted, so
k1 + k2 + k3 + k4 + k5 = log10 N 0 − log10 N5
⎛N ⎞
= log10 ⎜ 0 ⎟
⎝ N5 ⎠
= k tot
237
Maths for Science
The theory behind log-log and log-linear graphs (described in the next section)
relies on the application of the rules of logarithms to algebraic expressions,
illustrated in Worked example 10.2.
1 1
(iv) log10 3 + log10 x−2 (v) log10 3 + (vi)
log10 x 2 log10 3x 2
Answer
log10 (3x −2 ) = log10 3 + log10 x −2 (from Equation 10.2)
= log10 3 − 2loog10 x (from Equation 10.4)
Thus the expressions that are equivalent to log10(3x−2) are (iii) and (iv).
Question 10.6
238
10 Logarithms
A log10 A
0 r 0 log10 r
(a) (b)
Figure 10.2 Graphs of (a) A against r, and (b) log10 A against log10 r for the
equation A = πr2.
Reversing the order of the two terms on the right-hand side gives
log10 A = 2 log10 r +
log10 π
y =
m x + c
gradient intercept on
Table 10.2 The radius and area of various circles, and corresponding logarithms to
base 10.
Question 10.7
Find the gradient and intercept on the vertical axis of the straight line
shown in Figure 10.3.
239
Maths for Science
2.6
2.4
2.2
2.0
1.8
log10(A/cm2)
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
log10(r/cm)
Reversing the order of the two terms on the right-hand side gives
240
10 Logarithms
log10 y
y
0.6
0.4
0.2
0
0.1 0.2 0.3 0.4 0.5 0.6
−0.2 log10 x
−0.4
−0.6
0 x
(a) (b)
Figure 10.4 Graphs of (a) y against x and (b) log10 y against log10 x for the equation y = 3x−2.
Question 10.8
241
Maths for Science
T
TE
30 Saturn
28
26
24
22
20
18
16
14
12
Jupiter
10
4 Mars
Earth
2 Venus
Mercury
0 2 4 6 8 10 a
aE
Reversing the order of the two terms on the right-hand side gives
242
10 Logarithms
log10 T
TE
1.5
Saturn
1.0 Jupiter
0.5
Mars
− 0.5
Mercury
243
Maths for Science
N log10 N
0
0 t t
(a) (b)
Figure 10.7 Graphs of (a) N against t and (b) log10 N against t for the equation
N = N0e−λt.
Reversing the order of the two terms on the right-hand side gives
log10 N = −λ t log10 e + log10 N
0
= (−λ log10 e)t + log
10 N 0
This can be compared with the general equation of a straight-line graph,
y = mx + c
gradient intercept on
the vertical axis
So a graph of log10 N against t will be a straight line with a gradient of
−λ log10 e and an intercept on the vertical axis of log10 N0. Note that the
gradient of Figure 10.7b is negative, as expected.
Question 10.9
244
10 Logarithms
1000
disintegrations per minute (on a log scale)
100
10
1
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
time/minutes
Figure 10.8 A graph of disintegrations per minute (on a log scale) against time
for the radioactive decay of the excited state of barium-137.
245
Maths for Science
if p = eq then ln p = q
e3 = 20.09
e0.6931 = 2.000
e −1 = 0.3679
From these results it follows that:
e3 = 20.09 so ln 20.09 = 3
e0.6931 = 2 so ln 2 = 0.6931
Question 10.10
246
10 Logarithms
ln en = n (10.7)
ln(p × q) = ln p + ln q (10.8)
⎛ p⎞
ln ⎜ ⎟ = ln p − ln q (10.9)
⎝q⎠
ln(pn) = n ln p (10.10)
You may be wondering why logs to base e are useful; why are they
sometimes used in preference to logs to base 10? One reason stems from the
fact that taking a logarithm to base e is the inverse of raising e to a power.
This means that equations such as N = N0e−λt can be turned into simpler
straight-line equations by taking logarithms to base e than is possible by
taking logarithms to base 10.
Taking the log to base e of both sides of the equation N = N0e−λt gives
ln N = ln(N 0e −λt )
= ln N 0 + ln e −λt (from Equation 10.8)
= ln N 0 − λ t (from
m Equation 10.7)
Reversing the order of the two terms on the right-hand side gives
ln N = −λt + ln N0
This can be compared with the general equation of a straight-line graph,
y = mx + c
ln N = –λ t + ln N 0
y = m x + c
gradient intercept on
247
Maths for Science
ln N
intercept on the
vertical axis at ln N0
gradient of −λ
0
t
Box 10.4 investigates the relationship between decay constant λ and half-life,
t1/2. The half-life of a radioactive decay process (first introduced in Chapter 7)
is the time taken for the number of radioactive nuclei, and hence the activity,
to fall by half.
eλt1/ 2 = 2
248
10 Logarithms
ln(eλt1 / 2 ) = ln 2
i.e. λt1/2 = ln 2 (from Equation 10.7)
ln 2 ln 2
λ= or t1/ 2 = (10.11)
t1/ 2 λ
Thus a decay constant of 4.4 × 10−3 s−1 (for barium-137) corresponds to
a half-life of
ln 2
= 1.6 × 102 s to two significant figures
4.4 × 10−3 s −1
Question 10.11
kR = Ae(− Ea /RT )
(10.12)
where kR is the ‘rate constant’ at a particular temperature T, A is the
Arrhenius A-factor (or A-factor), Ea is the Arrhenius activation energy (or
activation energy) and R is the ‘gas constant’.
Taking the log to base e of both sides of Equation 10.12 gives
ln kR
= ln(Ae(−Ea /RT ) )
= ln A + ln e(−Ea /RT )
(from Equation 10.8)
E
= ln
n A − a (from Equation 10.7)
RT
Reversing the order of the two terms on the right-hand side gives
Ea
ln kR = − + ln A
RT
E 1
= − a + ln A
R T
249
Maths for Science
−Ea 1
ln kR = + ln A
R T
y = m x + c
gradient intercept on
the vertical axis
So if both A and Ea are constants, independent of temperature (a
reasonable assumption for most reactions when studied over a limited
range of temperature), a graph of ln kR against 1/T will be a straight line
of gradient −Ea/R and intercept on the vertical axis ln A. A graph of
ln kR against 1/T (as shown in Figure 10.10) is referred to as an
Arrhenius plot.
ln kR
intercept on the
vertical axis at ln A
gradient of −Ea
R
0
1
T
250
10 Logarithms
5.0
4.0
ln (heart rate)
3.0
2.0
1.0
0
3.3 × 10−3 3.4 × 10−3 3.5 × 10−3
T − 1/K−1
251
Maths for Science
Question 10.12
252
11 Probability and descriptive statistics
253
Maths for Science
to some genuine scientific problems: for example, what is the probability that
two people planning to have a child will both turn out to be carriers of the
cystic fibrosis gene?
The nature of the fraction in Equation 11.1 shows that the probability of any
given outcome cannot be smaller than 0 or larger than 1. A probability of 0
represents impossibility, while a probability of 1 represents inevitability. The
closer the probability of a given outcome is to 1, the more likely that outcome
is to occur. This is illustrated diagrammatically in Figure 11.1.
your guess is as
impossible good as mine inevitable
outcome outcome
0 1 1
2
increasingly unlikely increasingly likely
254
11 Probability and descriptive statistics
Dice games involve rolling six-sided dice (singular ‘die’). Each face of a die
is marked with a different score: one, two, three, four, five or six
(Figure 11.2). If the die is not loaded and the rolling is done fairly, then all
outcomes are equally likely, so the probability of any one of the six possible
outcomes (for example, scoring a three) is 16 . Again, the sum of the
probabilities of all the possible outcomes is
1 1 1 1 1 1
+ + + + + =1
6 6 6 6 6 6
So on one roll of the die the probability of scoring a three is 16 and the
probability of not scoring a three is 56 . Another way of expressing this is to
say that on a single roll of the die there is only one way of scoring a three,
but there are five ways of not scoring a three. Clearly, it is more likely than
not that a number other than three will be scored. This is just one illustration
of a more general finding: the most likely outcome is the one that can occur
in the greatest number of ways.
Provided nothing biases the result to make one outcome inherently more likely
than others, the definition given by Equation 11.1 can be rewritten to
encompass the number of ways in which a particular outcome may come Figure 11.2 Two dice.
about:
Question 11.1
What is the probability of one card drawn at random from a shuffled pack
of playing cards being: (a) a heart, (b) red, (c) an ace, (d) a picture card?
Note: if you are unfamiliar with playing cards, you need the following
information. There are 52 cards in a pack, divided into four suits: hearts
(red), diamonds (red), spades (black) and clubs (black). Each suit contains
13 cards, made up of one ace, nine ‘number’ cards (from 2 to 10 inclusive)
255
Maths for Science
and three picture cards (Jack, Queen, King). Figure 11.3 shows the 13 cards
in the suit of clubs.
1
will approach its theoretical value of 2
.
A failure to appreciate the fact that the number of attempts needs to be
extremely large before the probability of a particular outcome will reliably
approach the theoretical value is at the root of many popular misconceptions
about probabilities. One commonly held fallacy about coin tossing is that if
the first ten tosses of a coin have produced several more heads than tails, then
the eleventh toss is more likely than not to come up tails. This is not true.
Although in the extremely long run the imbalance between heads and tails is
expected to be negligible, on any one toss heads or tails are equally likely,
irrespective of previous history. Coins have no memory!
Question 11.2
(a) You toss a single coin three times. It comes down heads on the first two
occasions. What is the probability that you will get heads on the third
throw?
256
11 Probability and descriptive statistics
(b) If you toss two coins simultaneously and they both come down heads,
what is the probability that when you then toss a third coin it will also
come down heads?
For the rest of this chapter, probabilities will usually be expressed as fractions,
1 H H H
2 H H T
3 H T H
4 H T T
5 T H H
6 T H T
7 T T H
8 T T T
Of the eight combinations, only one – shown in red – represents the desired
outcome of three heads. On the basis of Equation 11.2, the probability of all
three coins coming up heads is therefore 18 .
The same result can be obtained using the multiplication rule for
probabilities: the probability that the first coin will show heads is 1
2
, and the
257
Maths for Science
same is true for both the second and the third coins. The probability that all
three will show heads is
1 1 1 1
× × =
2 2 2 8
Notice carefully how this situation differs from the one featured in Question
11.2: both the scenarios described in Question 11.2 correspond to having a
choice only between outcomes 1 and 2 in the list above (because the outcome
of the first two tosses is already known as being two heads).
The multiplication rule is expressed in its most general form by saying that:
1 1 1
× =
25 25 625
For a child whose parents are both carriers, the probability of inheriting a
copy of the CF gene from both parents is 14 . This is therefore the
258
11 Probability and descriptive statistics
probability that the child of such parents will have symptoms of the
disease.
■ What is the probability of a child born to white European parents
having cystic fibrosis?
□ 1
The probability that both parents are carriers is approximately 625 ,
and the probability that a child whose parents are both carriers will
have the disease is 14 . So the probability of a child born to white
European parents having cystic fibrosis is approximately
1 1 1
× =
625 4 2500
Question 11.3
(a) If you toss two coins at the same time, what is the probability of getting
two tails?
(b) If you throw a pair of dice, what is the probability of getting a pair of
sixes?
Question 11.4
259
Maths for Science
Question 11.5
If you were to draw one playing card from a pack of 52, what would be the
probability of that card being either the Jack, Queen or King of diamonds?
There are also cases in which both the addition and multiplication rules
operate as shown in Worked example 11.2.
260
11 Probability and descriptive statistics
But in a family with just one boy and two girls, the boy may be the eldest,
the middle or the youngest child, and these possibilities are mutually
exclusive. So the probability of the family consisting of a boy and two girls
(born in any order) is
1 1 1 3
+ + =
8 8 8 8
(Note that in fact the assumption that a baby is just as likely to be a boy as a
girl is not quite true. UK statistics show that for every 100 girls born, 106
boys are born.)
As with the coin-tossing example earlier, you may find that a table of the
possibilities helps in visualising the situation. Of the eight possible
combinations of three children, only three – shown in red – comprise one boy
and two girls.
Question 11.6
If you toss two coins simultaneously, what is the probability of getting one
head and one tail?
261
Maths for Science
262
11 Probability and descriptive statistics
plant had white flowers had no bearing on whether its seeds were round
or wrinkled or on what height the plant was.
You can think of ratios as simply another way of writing fractions. If, for
instance, you discovered from a paint chart that a green paint had been mixed
from yellow paint and blue paint in the ratio 3 : 2, you would understand that
the green paint was made up of three parts yellow paint and two parts blue
paint. In other words, 53 of the mixture was yellow and 52 was blue. Adding
both sides of the ratio together has given the denominator of the fractions.
Knowing the denominator, it is then easy to express the ratio in terms of
percentages: 53 = 100
60
so 60% of the mixture is yellow and 40% is blue. A
60 : 40 ratio is exactly the same as a 3 : 2 ratio – it is just a matter of
multiplying or dividing both sides of the ratio by 20. Sometimes it is
convenient to simplify even further, in this case by dividing both sides by two
to express the 3 : 2 ratio in the equivalent form of 1.5 : 1. Note that, like
fractions, ratios do not have units attached to them.
Ratios are quoted in many applications. For example, fertilisers are
characterised on their labelling by the ratio of two or three major ingredients,
each indicated by a letter. These letters are N (for nitrogen, which is required
for leaf growth), P (for phosphorus, which in the form of phosphates is
required for root development) and K (for potassium, which in the form of
potash is required for flowers and fruit). Typical ratios for three common types
of fertiliser are shown in Table 11.1.
Fertiliser N P K Others
bone meal 1 5 0 19
lawn tonic 4 1 0 5
tomato food 6 5 9 80
263
Maths for Science
5 5 1
= =
1 + 5 + 19 25 5
4 4 40
= =
4 + 1 + 5 10 100
So lawn tonic contains 40% N.
9 9
=
6 + 5 + 9 + 80 100
So tomato food contains 9% K.
As already noted for the paint example, it is quite common for ratios to be
expressed in a form such that one of the parts is 1, even if this means that the
other part is a decimal number. Question 11.7 gives an illustration of a ratio
expressed in such a way.
Question 11.7
In the atmosphere, the ratio of the volume of oxygen to the volume of other
gases is 0.26 : 1. What percentage of the atmosphere is oxygen?
The ratio of 705 : 224 that Mendel obtained for purple- to white-flowered
plants (see Box 11.2) can be simplified by dividing both sides of the ratio by
224 to obtain the equivalent ratio of 3.15 : 1. Notice that one side of this ratio
is exact: 224
224
is exactly equal to 1. However, the other side is not exact and a
choice has to be made about how many significant figures to quote; two or
three significant figures are usually sufficient in this context. Mendel’s data
relating to the other independent pairs of characteristics involving seeds and
stem lengths can be simplified in a similar way by dividing the larger number
by the smaller, to obtain:
In each case the ratio is close to 3 : 1. In other words, the character from the
P-generation that was present in all members of the F1-generation is present in
only about 34 of the F2-generation. By the same token, the character that
completely vanished in the F1-generation reappears in about 14 of the F2
generation. In fact, modern understanding of genetics leads to the theoretical
264
11 Probability and descriptive statistics
Figure 11.5 Maize cobs. The cob on the right of the photograph has the four
types of grain described in Worked example 11.3.
The types of grain on 20 cobs from the same plant were counted and the
aggregate results were:
Assuming that the theoretical ratios for these characteristics are whole
numbers, what would be the theoretical probability that a single grain chosen
at random from a large number of cobs would be a pale smooth one?
265
Maths for Science
Answer
Dividing through by the smallest number in the sample, which in this case is
531, gives:
If it is assumed that the theoretical ratios are whole numbers, these data
strongly suggest that the ratios would be:
The theoretical fraction of grains that are pale and smooth is therefore
3 3
=
9 + 3 + 3 + 1 16
This is also the probability of one grain selected at random being pale and
smooth. This probability could be expressed as a fraction ( 163 ), a decimal
number (0.1875) or a percentage (18.75%).
266
11 Probability and descriptive statistics
267
Maths for Science
268
11 Probability and descriptive statistics
number of measurements
0
2.446 2.448 2.450 2.452 2.454 2.456 2.458 2.460 2.462 2.464 2.466
(a) cell constant/nm
25
number of measurements
20
15
10
0
2.446 2.450 2.454 2.458 2.462 2.466
(b) cell constant/nm
300
measurements
250
number of
200
150
100
50
0
2.446 2.450 2.454 2.458 2.462 2.466
(c) cell constant/nm
measurements
number of
Figure 11.7 Distributions for repeated measurements of the unit cell constant of a batch of industrial catalyst. The
number of measurements increases in going from (a) to (d).
269
Maths for Science
24.544 nm
or 2.4544 nm to five significant figures
10
(The reason for giving the result to this number of significant figures will
be discussed shortly.)
1 n
x= ∑ xi
n i =1 (11.3)
The i = 1 below the summation sign indicates that the first value for xi in the
sum is x1, and the n above it indicates that the last value in the sum is xn. In
other words, all integer values of i (x1, x2, x3, etc.) are to be included up to xn.
(The summation sign with the information attached to it is usually said as
‘sum of x sub i from one to n’.)
270
11 Probability and descriptive statistics
(11.4)
i =1
Step 4
Divide by the total number of measurements (i.e. n) to obtain the mean of all
the square deviations. This may be written as:
1
n 2
d
i 2 =
∑ di
n
i =1
(11.5)
Step 5
Take the square root of this mean to obtain the ‘root mean square deviation’
σ. It is this quantity σ that is known as the standard deviation (σ is the lower
case version of the Greek letter sigma). Step 5 may be written as:
σ
=
di 2
or, substituting for di 2 from Equation 11.5, as:
1
n 2
σ
=
∑ di
n
i =1
(11.6)
271
Maths for Science
Since di was defined in Equation 11.4 as (xi – x ), one final substitution into
Equation 11.6 gives σ in its most easily used format:
1 n
σ = ∑ ( xi − x )2
n i =1
(11.7)
At the end of this process, all the data in Table 11.2 can be summarised by
saying that the ten measurements had a mean of 2.4544 nm and a standard
deviation of 0.0046 nm. The calculation of standard deviation is given in and
below Table 11.3.
Table 11.3 Calculation of the standard deviation for the set of measurements
originally given in Table 11.2.
Since
di 2 = 2.104 × 10−5 nm 2
it follows that
σ = di 2
= 4.587 ×10−3 nm
= 0.0046 nm
There are several things worth noting about this result and the data in
Table 11.3.
272
11 Probability and descriptive statistics
First, all the quantities have units associated with them. The values of xi were
measured in nanometres, so deviations will also be in nanometres and the
squares of the deviations in nm2, as shown in the column headings in the
table.
A second feature to notice is that the sum of all the deviations is equal to
zero, i.e.
∑ di = 0
This is always true and provides a useful check of your arithmetic. At the end
of Step 1 it is well worth adding up all the values you have calculated for the
deviations to ensure that they do indeed total zero. If they don’t, you have
made an arithmetic slip somewhere which needs to be put right before you
proceed to Step 2.
Looking now at the details of the calculation, the original measurements of
length were made to the nearest picometre (i.e. 0.001 nm), represented by
three decimal places. More digits were carried in the calculations to avoid
rounding errors. However, what is the appropriate number of digits to quote in
the final answer? When all the ten results in Table 11.3 were added, the result
was
273
Maths for Science
8, 6, 9, 12, 10
Step 3
Having input the data, you can then get most calculators to tell you the
number of items of data. If your calculator can do this, it should return the
answer ‘5’ here. It doesn’t matter if your calculator doesn’t have this function,
but if it does, it’s well worth using this checking device. If you have to input
a long string of data values, it’s quite easy to miss one out inadvertently.
Step 4
When you know you have the data correctly stored, find out how to display
the mean; you should get the answer ‘9’ here.
Step 5
Now find out how to display the standard deviation. The problem here is that
there is another type of standard deviation (which you will meet in
Section 11.2.7) and the notation used by calculators and spreadsheets can be
extremely confusing. Many calculators use the symbol σ for the standard
deviation that you are looking for now, but you may also come across sn.
Notation such as ‘STDEVP’ is used in spreadsheets such as Microsoft Excel.
You are looking for a function that may be described simply as ‘standard
deviation’ or as the ‘standard deviation of the entire population’.
It is essential that you check that you can use your calculator or spreadsheet
to obtain the answer ‘2’ for the data listed in Step 2.
Once you are sure you know how to use your calculator or spreadsheet to
perform calculations of mean and standard deviation, apply this skill to
Question 11.8.
Question 11.8
274
11 Probability and descriptive statistics
Measurement Diameter/mm
1 1.09
2 1.00
3 1.25
4 1.24
5 1.29
6 0.89
7 1.09
8 1.14
9 1.22
10 1.01
275
Maths for Science
mean value
σ
σ
Figure 11.8 The shaded area under this normal distribution curve represents the
measurements that lie within one standard deviation of the mean. 68% of all
measurements are expected to fall within this range.
quantity, the narrower the peak of the distribution curve and the smaller the
standard deviation will be. A very broad distribution, on the other hand,
corresponds to measurements with considerable scatter and the standard
deviation will be large. These trends are illustrated in Figure 11.9. The
measurements of quantity w are subject to large random uncertainties, while
those of quantity y are more precise and those of z more precise still.
w w y y z z
276
11 Probability and descriptive statistics
mean
mean
(a) (b)
mean
mean
(c) (d)
In many cases, especially if the distribution is skewed, the mean is not the
best way of representing an average or typical value. Imagine for example a
small company with a single owner who pays himself £1000 000 a year and
10 employees who are each paid £20 000. The statement that the mean annual
income of these 11 workers is more than £100 000 (i.e. £1200 000 divided by
11) – although true – is somewhat misleading. In such cases, two other
quantities, the mode and the median, may represent the data more fairly.
The mode is the most frequently occurring value in the set of data. If the data
are plotted on a histogram or a bar chart, the mode will be the value
corresponding to the tallest bar.
■ What is the mode of the earnings in the company described above?
□ The mode is £20 000. This is certainly more representative of the typical
earnings than the mean would be.
Note that in some cases there may be more than one value for the mode; for
example, the distribution shown in Figure 11.10d is ‘bimodal’, i.e. it has two
modes.
The median is the middle value in a series when the values are arranged in
order of size. This means that half the measurements have values that are
bigger than the median and half have values that are smaller than the median.
If there are an odd number of measurements, the median is the middle
measurement; if there are an even number of measurements it is the mean of
the middle two values.
277
Maths for Science
To see how this works, consider the following example. Ten plants of a
particular species were chosen at random and the number of flowers on each
plant were counted. The results were:
8, 7, 4, 8, 10, 7, 9, 7, 8, 7
■ What is the mode for these data?
□ The best way of answering this is to compile a table showing the number
of plants with particular numbers of flowers:
number of flowers 4 7 8 9 10
number of plants 1 4 3 1 1
The mode is 7 flowers. There are more plants with 7 flowers than with
any other number of flowers
4, 7, 7, 7, 7, 8, 8, 8, 9, 10
7+8
= 7.5
2
Question 11.9
The heights of nine different specimens of the same type of plant were
measured in centimetres, and the results in descending order were:
Box 11.4 illustrates a case in which the median gives a more representative
summary of the data than the mean.
278
11 Probability and descriptive statistics
Table 11.5 The recovery location of Storm Petrels ringed at their nests on one
of the Shetland islands.
Taking all 28 observations into account, the mean distance from their
nest site at which the birds have been recovered is 554.5 km. However,
this is not a very useful way in which to summarise the data, because in
fact 13 out of the 28 birds (i.e. nearly half) moved less than 100 km, and
only two moved further than the mean distance.
The median distance is 114 km and this is a more typical value.
This example shows how the mean can be highly dependent on a small
number of measurements that are a long way from the mode. In this
case, the single recovery from South Africa has an enormous influence
on the mean. But the median is ‘resistant’ to extreme values. Even if the
bird recovered in South Africa had flown all the way to New Zealand,
the median value would have remained 114 km.
279
Maths for Science
280
11 Probability and descriptive statistics
must come from the distribution of values in a sample drawn from the
population. However, this time it is not appropriate to use the formula for the
standard deviation of repeated measurements of one quantity which was:
1 n
σ = ∑ (xi − x)2
n i=1
(11.7)
1 n
s= ∑ ( xi − x )2
n − 1 i =1
(11.8)
8, 6, 9, 12, 10
The first four steps for calculator use are the same as before, only Step 5 will
be different.
Step 1
Put the machine into statistical mode.
Step 2
Input all the data.
Step 3
If your calculator can tell you the number of items of data, check that it gives
the answer ‘5’ here.
281
Maths for Science
Step 4
When you know you have the data correctly stored, display the mean; you
should get the answer ‘9’ here.
Step 5
Now find out how to display the sample standard deviation. The appropriate
button will probably be marked s but σn−1 or sn−1 are other possibilities. On a
spreadsheet such as Microsoft Excel, the function may confusingly be labelled
STDEV (not STDEVP – the ‘P’ here refers to the fact that this is the function
to use if you are calculating a value for standard deviation based on data for
the entire population). Whatever the case, make sure that you can use your
calculator or spreadsheet to give an answer of ‘2.2’ (to one decimal place) for
the sample standard deviation in this case.
While this example is useful to familiarise yourself with the process, it doesn’t
represent a realistic scenario, not least because the hypothetical data set is so
small. Because the aim is to estimate the mean and standard deviation for a
whole population by carrying out measurements just on a sample, it is
important to ensure that the sample is representative of the population as a
whole and that usually requires it not only to be chosen without bias, but also
to be reasonably large. In Question 11.10, the sample consists of 20 plants.
Question 11.10
8, 8, 4, 8, 8, 7, 9, 7, 7, 5, 9, 10, 6, 9, 7, 4, 8, 5, 11, 6
From these data, estimate to one decimal place the mean number of flowers
per plant and the standard deviation for the whole colony.
282
11 Probability and descriptive statistics
Question 11.11
283
12 Statistical hypothesis testing
(a)
(b)
285
Maths for Science
Some of this is medieval and some is much later in age. It has been
known for some time that Bulbous Buttercup (Ranunculus bulbosus)
tends to occupy the drier ridges and Creeping Buttercup (R. repens) the
wetter furrows. Also found in the same area is Green-winged Orchid
(Orchis morio), a rare plant in England, shown in the inset for
Figure 12.2. A study was undertaken to find out whether the distribution
and/or performance of Green-winged Orchid might also be influenced by
ridge-and-furrow topography. Various measurements were made on a
sample of plants growing in a local nature reserve. Figure 12.2 illustrates
some of the measurements taken. These included the horizontal and
vertical distances of each plant from the nearest ridge crest, the height of
the plant, the number of leaves and the number of flowers. Whether a
plant was growing on the north-west or the south-east slope of the ridge
was also recorded, since the two slopes might differ with respect to mean
temperature, moisture availability, etc.
position of
ridge crest
plant
height
vertical
distance
horizontal distance
286
12 Statistical hypothesis testing
to ridge crests than expected by chance?’, ‘Does the amount of water in soil
increase with distance from the nearest ridge crest?’, ‘Do the Green-winged
Orchids growing nearer ridge crests tend to be taller or have more leaves and/
or flowers than those growing further away?’.
There are two major branches of statistical hypothesis-testing: tests of
association (e.g. ‘Are Green-winged Orchids found in association with ridge
crests significantly more frequently than would be expected by chance?’) and
tests of difference (e.g. ‘Is there a significant difference between the mean
height of plants growing on the north-west rather than the south-east slopes of
ridges?’).
■ Would an investigation into whether there is a significant increase in the
water content of soil with increasing distance from the nearest ridge crest
be a test of association or a test of difference?
□ Since you would be looking to see if there is an association between soil
water content and distance from ridge crest, this would be a test of
association.
287
Maths for Science
experimental
number of measurements
treatment
control
s1 s1 s2 s2
x
x1 x2
that there is no difference between the population mean of the treated plants
(μ1) and the population mean of the control plants (μ2). Expressing this
statement mathematically, the null hypothesis would be that
μ1 = μ2
or, equivalently, that μ1 − μ2 = 0.
At the same time, the scientist has to put forward an alternative hypothesis
that is the logical ‘mirror image’ of the null hypothesis. In this case the
alternative hypothesis would be that there is a difference between the means
of the treated and control plants. Expressing this statement mathematically, the
alternative hypothesis would be that
μ1 ≠ μ2
or μ1 − μ2 ≠ 0.
■ Is it possible for both the null and alternative hypotheses to be true?
□ No. If either is true, then the other must be false.
Once statements of the null and alternative hypotheses have been made, a
quantity called the test statistic is calculated. The test statistic is a number, on
the basis of which a decision can be made to accept or reject the null
hypothesis. The value of the test statistic depends on the characteristics of the
samples being compared, and in most cases it is calculated using one or more
equations. Things are often so arranged that the value of the test statistic
comes out to be zero if the null hypothesis is true (for instance, by including
288
12 Statistical hypothesis testing
the term (x1 − x2 ) in the numerator of the equation, where x1 and x2 , the
means of the two samples, are the best available estimates of the unknowable
values of μ1 and μ2). However, because of the vagaries of sampling, it would
be extremely unlikely for the means of two samples drawn from even the
same population to be identical (for instance, two samples of control plants
are very unlikely to have exactly the same mean). So, the question is ‘How
large does the test statistic have to be before one can be reasonably confident
that the samples were drawn from different populations (and therefore
conclude, in this example, that the experimental treatment probably did have a
significant effect)?’ In fact, it is impossible to give a definitive answer to this
question; it can be answered only in terms of probabilities.
A computer program or spreadsheet can be used to find the probability that
the calculated value of the test statistic could have arisen by chance if the null
hypothesis were true. In a particular instance, this might turn out to be
something like 1 in 63, i.e. 63 1
, which is 0.015 87 to four significant figures.
An alternative approach, used in this book, is to compare the value of the test
statistic with lists of critical values calculated for a few predetermined
significance levels expressed in terms of probabilities. In this context, the
probabilities are usually abbreviated to P and expressed in decimal notation,
e.g. 0.1, 0.05 and 0.01 (as percentages, these values would be 10%, 5% and
1%). For any particular significance level, the critical value is the most
extreme (usually largest) value that the test statistic could be expected to have
if the null hypothesis were true. Of course, if the null hypothesis is true then
any deviation from the test statistic’s expected value (which, as noted above,
is usually zero) must have arisen purely by chance. If the significance level
corresponding to the value of the test statistic turns out to be quite low
(usually because the test statistic is rather high), then it must be accepted that
the null hypothesis is unlikely to be true. If the null hypothesis is false, then
the alternative hypothesis is assumed to be true. Only at this stage can the
scientist conclude:
. either that the treatment did have a significant effect (because the null
hypothesis was probably false and therefore the alternative hypothesis
probably true)
. or that the treatment did not have a significant effect (because the null
hypothesis is likely to have been true).
It is extremely important to realise that the particular significance level at
which a null hypothesis is rejected – and hence the alternative hypothesis is
accepted – is a matter of convention. For most situations in science, the usual
convention is to reject a null hypothesis if the probability P is less than the
0.05 significance level, i.e. if P < 0.05. However, in employing this
convention, it is also important to realise that you could be rejecting a true
null hypothesis or accepting a false one. Indeed, you are explicitly accepting
that on average, if you were to carry out 100 statistical tests, you would reach
the wrong conclusion for 5 of these tests (although you would not know
which ones). If the work you are engaged in really matters, for example,
medical research in which human lives might be at stake, then you would
probably employ more exacting criteria, such as rejecting null hypotheses only
if P < 0.01 or even P < 0.001. On the other hand, insisting on the use of such
289
Maths for Science
rigorous criteria for routine scientific work would mean that many null
hypotheses that really are false would have to be accepted, and this would
undoubtedly hinder scientific progress.
The important features of statistical hypothesis testing are summarised below:
When null hypotheses are rejected, the results are described as being
statistically significant or sometimes just as ‘significant’. A consequence of
this is often a feeling that ‘non-significant’ results are of less value than
‘significant’ ones. Indeed, there is probably a reporting bias whereby
significant differences are more likely to be published in scientific papers than
non-significant ones. This undervaluing of non-significant differences is
unfortunate because the whole point of the exercise is to try to find out what
is happening in the real world. It may be just as important to know that an
effect is not produced by one experimental treatment as to know that another
treatment does produce the effect.
Question 12.1
290
12 Statistical hypothesis testing
of leaves they possess without knowing the actual heights or the actual
numbers of leaves. If the actual heights or numbers of leaves are known, then
these data are described as being at the interval level.
Data collected at the interval level can, if necessary, be analysed at the ordinal
level. For instance, you might know that Plant A has 8 leaves and that Plant B
has 5 leaves (interval level data). Nevertheless, you could choose to ignore
some of this information and simply treat Plant A as having more leaves than
Plant B (ordinal level data). Of course, if all you knew was that Plant A has
more leaves than Plant B, then you could not convert this information into
interval level data for analysis.
Categorical level data cannot usually be treated as if they were at interval or
ordinal level (although you might argue that, for instance, red-flowered plants
have more of a particular pigment than pink-flowered plants of the same
species). However, by applying arbitrary criteria, interval or ordinal level data
can sometimes be converted into categorical data for analysis. For instance,
one of the seven pairs of contrasting characters used by Mendel in his
pioneering research on the genetics of garden peas (see Box 11.2) was ‘tall’
versus ‘short’. This categorical distinction made sense only because, in this
particular case, there was no overlap between ‘tall’ and ‘short’ plants.
The reason for distinguishing between the different levels of measurement is
that different statistical tests must be used to analyse categorical, ordinal and
interval level data. Sometimes, when analysis of data at the interval level fails
to reveal statistically significant differences, such differences may be shown
up when the data are re-analysed at the ordinal level. However, because some
information about the samples has effectively been ‘thrown away’ in the
process, any statements eventually made about the populations from which the
samples were drawn are necessarily less complete than they might have been.
Question 12.2
In each of the following cases, explain briefly whether the data should be
treated as being at the categorical, the ordinal or the interval level.
(a) A count is made of the number of parasites on each member of a
sample of sheep.
(b) A sample of sheep are counted as either ‘parasitised’ (i.e. carrying one
or more parasites) or ‘unparasitised’ (i.e. carrying no parasites).
(c) A sample of sheep are counted as ‘unparasitised’ (i.e. carrying no
parasites), ‘lightly parasitised’ (i.e. carrying 1–5 parasites), ‘moderately
parasitised’ (i.e. carrying 6–10 parasites) or ‘heavily parasitised
(i.e. carrying more than 10 parasites).
291
Maths for Science
possible associations between two categorical variables, but for simplicity the
χ2 goodness-of-fit test will be referred to simply as ‘the χ2-test’ for the rest of
this chapter.
In the Green-winged Orchid study, described in Box 12.1, horizontal distance
from the nearest ridge crest (as shown in Figure 12.2) was recorded for 210
plants growing on several ridges. Because the ridge crest-to-furrow distance
varied slightly between ridges, each of these distances was divided into five
equal categories (category 1 being 0.0–19.9% of the distance from the crest,
category 2 being 20.0–39.9% of the distance, category 3 being 40.0–59.9% of
the distance, etc.) so that the data from different ridges could be pooled for
analysis. This procedure enables interval level data (the horizontal distance of
each plant from the nearest ridge crest) to be treated as categorical level data
(the distance category into which each plant falls). If the 210 plants were
distributed uniformly with respect to the ridge crest, then a fifth of them
(i.e. 42) would be expected to occur within each distance category. A
reasonable null hypothesis would be that, if it were possible to collect data on
the entire population of Green-winged Orchids growing in fields with ridge
and-furrow topography, then there would be no difference between the number
of plants observed in each distance category and the number that would be
expected on the assumption that the plants were distributed uniformly. The
alternative hypothesis would be that the number of plants observed in each
distance category was not equal to the number of plants expected. Accepting
this alternative hypothesis implies accepting that the plants were distributed
non-uniformly.
In fact, of the sample of 210 plants, 105 occurred in the first distance
category, 74 in the second, 28 in the third, 3 in the fourth and none in the
fifth. It certainly appears that the plants were not uniformly distributed. The
χ2-test allows a definitive statement to be made on the probability that the
population of plants from which the sample was drawn could have been
distributed uniformly despite the apparently non-uniform distribution observed
in the sample. Only if this probability is sufficiently low (conventionally if
P < 0.05) can the null hypothesis be rejected and the alternative hypothesis
(with its implication that the plants were distributed non-uniformly) accepted.
The first stage in performing a χ2-test is usually to draw up a table to compare
observed and expected numbers in different categories. The table for the
sample of 210 orchid plants is given in Table 12.1, and compares the number
of individuals, Oi, that were observed in each distance category, with the
number Ei expected on the basis of the null hypothesis. As a check, the total
number in the Oi column should equal the total number in the Ei column. The
trickiest part of most χ2-tests is deciding the ‘expected’ numbers. In this case,
if the null hypothesis were true, a fifth of the plants (i.e. 42) would be
expected to fall into each distance category.
292
12 Statistical hypothesis testing
(Oi − Ei ) 2
Ei
Step 4
The results from Step 3 are totalled. This leads to the test statistic χ2, where χ2
is given by
n
(Oi − Ei ) 2
χ2 = ∑ (12.1)
i =1 Ei
The easiest way to calculate χ2 is to extend Table 12.1 to include columns for
(Oi − Ei ) 2
(Oi − Ei), (Oi − Ei)2 and .
Ei
This has been done in Table 12.2, and χ2 is the total of the values in the right
hand column. Notice that, as a further check, the total of the (Oi − Ei) column
must be zero, since the total number of individuals observed is equal to the
total number of individuals expected.
293
Maths for Science
(Oi − Ei ) 2
As an example of the way in which each value of is calculated,
Ei
consider the first distance category. For this distance category, Oi = 105 and
Ei = 42, so
The next stage is to compare the value of the test statistic χ2 (which, in this
case, is 201.762) with the critical values listed in Table 12.3. The sizes of the
critical values in such a table depend on both the significance level (P = 0.1,
P = 0.05 and P = 0.01, given across the top of the table) and the number of
degrees of freedom (given down the left-hand side of the table).
The number of degrees of freedom can be found by counting the number of
‘cells’ in the table that contain observed counts (i.e. ignoring expected counts,
totals, etc.) and subtracting one.
In this case, Table 12.1 has five cells that contain observed counts, so
number of degrees of freedom = 5 − 1
=4
Box 12.2 gives a brief explanation of why it is reasonable for the number of
degrees of freedom to be four in this case.
294
12 Statistical hypothesis testing
Table 12.3 Critical values of χ2 for different degrees of freedom and at three
levels of significance.
Note: The null hypothesis is usually rejected if, for the appropriate number of
degrees of freedom, the calculated value of χ2 is greater than the value tabulated
at the P = 0.05 significance level.
295
Maths for Science
Table 12.3 Critical values of χ2 for different degrees of freedom and at three
levels of significance.
Degrees of freedom Critical value
P = 0.1 P = 0.05 P = 0.01
1 2.706 3.841 6.635
2 4.605 5.991 9.210
3 6.251 7.815 11.341
4 7.779 9.488 13.277
5 9.236 11.070 15.086
Reading across the row for 4 degrees of freedom, it can be seen that the χ2
value of 201.762 is greater than 7.779 (corresponding to a significance level
of 0.1), greater than 9.488 (corresponding to a significance level of 0.05) and
greater than 13.277 (corresponding to a significance level of 0.01). So the
significance level is less than P = 0.01. In fact, the significance level is
considerably less than 0.01 (because 201.762 is much larger than 13.277).
Thus, the probability that the plants in the population from which the sample
was drawn were distributed uniformly is much less than 0.01 (which can be
written as P << 0.01). There can be little doubt that the plants were not
distributed uniformly with respect to distance from the ridge crest. The null
hypothesis can therefore be rejected – and the alternative hypothesis accepted
296
12 Statistical hypothesis testing
297
Maths for Science
Table 12.4 gives values for Oi, Ei and total sample size and has been extended
(Oi − Ei ) 2
to give (Oi − Ei), (Oi − Ei)2 and .
Ei
Table 12.4 An extended table for Worked example 12.1.
Question 12.3
298
12 Statistical hypothesis testing
For example, since the original mass of one sample was 22.85 g and its dry
mass was 11.32 g, its water content was
(22.85 g − 11.32 g)
×100% = 102%
11.32 g
(This percentage is greater than 100% because there was slightly more water
than soil in the sample.)
In fact, several soil samples (known as ‘replicate’ samples) were taken at each
horizontal distance, and their mean water content was calculated and used for
the rest of the investigation. Figure 12.5 shows how the mean water content of
the soil samples taken on the north-west slope of the ridge varied with
horizontal distance from the nearest ridge crest.
140
mean water content / % dry mass
120
100
80
60
40
20
Figure 12.5 Mean water content (as a percentage of dry mass) of soil samples
plotted against horizontal distance from ridge crest.
299
Maths for Science
variable 2
variable 2
variable 2
(a) variable 1 (b) variable 1 (c) variable 1
Figure 12.6 (a) A perfect positive correlation between two variables (i.e. r = +1). (b) A perfect negative correlation
between two variables (i.e. r = −1). (c) Zero correlation between two variables (i.e. r ≈ 0).
Several different sorts of correlation coefficient have been devised. In this case
it is appropriate to calculate the Spearman rank correlation coefficient (rS).
This, as the term ‘rank’ suggests, is based on ordinal level data. The null
hypothesis is that there is no correlation between soil water content and
horizontal distance from ridge crest (i.e. rS = 0) and the alternative hypothesis
is that the two variables are correlated (i.e. rS ≠ 0).
The data on mean soil water content for the north-west slope of the ridge are
summarised in Table 12.5.
Before the test statistic can be calculated, the following steps should be
completed:
Step 1
Work out the rank (order) of each of the eight horizontal distances, (RA)i
(which will range between 1 and 8).
Step 2
Work out the rank of each matching value for mean water content, (RB)i
(which will also range between 1 and 8).
300
12 Statistical hypothesis testing
Step 3
Calculate each difference, Di = (RA)i − (RB)i.
Step 4
Square each difference, to give Di2.
Step 5
n
Total all the values for Di2 from Step 4 to give ∑ Di 2 .
i =1
i =1
n
Notice that
∑ Di (the sum of the differences of the ranks) should always be
i =1
zero, which provides a check that the ranks have been worked out correctly.
Table 12.6 Extension of Table 12.5 to include ranks ((RA)i and (RB)i), differences between ranks (Di) and values of Di2.
Horizontal distance/cm Rank (RA)i Mean water content/% Rank (RB)i Di = (RA)i − (RB)i Di2
0 1 76 1 0 0
50 2 83 3 −1 1
100 3 93 4 −1 1
150 4 80 2 2 4
200 5 102 6 −1 1
250 6 95 5 1 1
300 7 120 7 0 0
350 8 130 8 0 0
n n
∑ Di = 0 ∑ Di 2 = 8
i =1 i =1
In the case of the data in Table 12.6, it was possible to assign a unique rank
to each value for horizontal distance and mean water content, but sometimes
quantities ‘tie’ (i.e. have the same rank). Worked example 12.2, at the end of
this section, illustrates what to do when this is the case.
301
Maths for Science
The test statistic, the Spearman rank correlation coefficient, is rS and this
is calculated using Equation 12.2:
n
6∑ Di 2
rS = 1 − i =1 (12.2)
n(n 2 − 1)
n
where ∑ Di 2 is the sum of the differences of the ranks and n is the
i =1
number of pairs of measurements.
n
Substituting ∑ Di 2 = 8 (from Table 12.6) and n = 8 into Equation 12.2 gives
i=1
6×8
rS = 1 −
8(82 − 1)
= 0.905
The final stage is to compare the value of the test statistic rS (0.905 in this
case) with the critical values listed in Table 12.7. The critical values are again
given to three decimal places and the size of the critical values depends on
both the significance level (P = 0.1, P = 0.05 and P = 0.01, given across the
top of the table) and the number of pairs of measurements (given down the
left-hand side of the table). In this case the number of pairs of measurements
is 8, and looking across the appropriate row it can be seen that the calculated
rS value of 0.905 is greater than 0.881, corresponding to a significance level
of 0.01. Thus the probability, P, that there is no correlation between water
content and horizontal distance from the ridge crest is less than 0.01; the null
hypothesis must be rejected at the 1% significance level, and the alternative
hypothesis accepted. There is a statistically significant (positive) correlation
between mean soil water content and horizontal distance from ridge crest.
It is extremely important to appreciate that even a statistically significant
correlation between two variables does not prove that changes in one variable
cause changes in the other variable.
302
12 Statistical hypothesis testing
Table 12.7 Critical values for the Spearman rank correlation coefficient (rS)
for different numbers of pairs of measurements and at three levels of
significance.
Note: (i) The null hypothesis is usually rejected if, for the appropriate number of
pairs of measurements, the calculated value of rS is greater than or equal to the
value tabulated at the P = 0.05 significance level.
(ii) The lower part of Table 12.7 does not have entries for odd numbers of pairs
of measurements. Should the data you are analysing comprise (say) 17 pairs of
measurements, it is better to err on the side of caution and compare your value
of the test statistic with the critical values for 16 pairs rather than those for 18
pairs. Because each critical value for 16 pairs of measurements is higher than
the corresponding value for 18 pairs, this makes it less likely that you will
mistakenly reject a true null hypothesis.
303
Maths for Science
Worked example 12.2 illustrates how to rank data when two or more
measurements are identical. They must be given the same mean rank, and then
account must be taken of all the identical measurements before the rank of the
next, non-identical, value is decided. So, if two measurements tie for first
place, they are each given a rank of
1+ 2
= 1.5
2
and the next available rank is 3.
Answer
Table 12.9 is an extension of Table 12.8, to include values for (RA)i, (RB)i, Di,
n
Di2 and ∑ Di 2 for the data in this worked example.
i=1
Note, for example, that the water speed was measured to be 0.2 m s−1 at two
sampling stations, so these stations ‘tie’ for second place in the ranking of
water speed (after the station with a water speed of 0.1 m s−1).
304
12 Statistical hypothesis testing
Table 12.9 Extension of Table 12.8 to include ranks ((RA)i and (RB)i), differences between ranks (Di) and values of Di2.
Water speed/m s−1 Rank (RA)i Number of nymphs Rank (RB)i Di = (RA)i − (RB)i Di2
0.8 9 35 12 −3 9
1.1 11 28 11 0 0
0.5 5.5 11 6 −0.5 0.25
0.7 7.5 12 7 0.5 0.25
0.2 2.5 7 4 −1.5 2.25
0.4 4 5 1 3 9
0.5 5.5 6 2.5 3 9
1.3 12 21 9 3 9
0.9 10 23 10 0 0
1.7 13 43 13 0 0
0.2 2.5 10 5 −2.5 6.25
0.1 1 6 2.5 −1.5 2.25
0.7 7.5 19 8 −0.5 0.25
n n
∑ Di = 0 ∑ Di 2 = 47.5
i=1 i =1
6 × 47.5
rS = 1 −
13(132 − 1)
= 0.870
Reading across the row for 12 pairs of measurements (in the absence of a row
for 13 pairs) in Table 12.7, it can be seen that P < 0.01. The null hypothesis
must therefore be rejected at the 1% significance level and the alternative
hypothesis accepted. There is a statistically significant positive correlation
between water speed and number of Stonefly nymphs.
Question 12.4
Returning to the study described in Box 12.1, Figure 12.8 shows how the
mean water content of the soil samples taken from the north-west slope of
the ridge varies with vertical distance from ridge crest. Use the data given
in Table 12.10 (an expanded version of Table 12.5) to determine whether
there is a statistically significant correlation between soil water content and
vertical distance from ridge crest. The horizontal distances from the ridge
are not required for this question so this column has been greyed out.
305
Maths for Science
140
100
80
60
40
20
0 2 4 6 8 10 12 14
vertical distance from ridge crest/cm
Figure 12.8 Mean water content (as a percentage of dry mass) of soil samples
plotted against vertical distance from ridge crest.
Table 12.10 Vertical distances from the nearest ridge crest and mean soil water
content (as a percentage of dry mass) for samples taken at various horizontal
distances from the nearest ridge crest on the north-west slope of a ridge.
306
12 Statistical hypothesis testing
■ State the null and alternative hypotheses that would be appropriate for a
t-test.
□ Since a t-test would be concerned with the difference between the means
of two populations (1 and 2), the appropriate null hypothesis would be
that μ1 = μ2 (or equivalently μ1 − μ2 = 0) and the appropriate alternative
hypothesis that μ1 ≠ μ2 (or μ1 − μ2 ≠ 0).
Question 12.5
In each of the following cases, explain whether the samples are matched or
unmatched.
(a) A comparison is made between the heights of a sample of Green-winged
Orchids growing in one nature reserve and those of a sample growing in
another nature reserve.
(b) The numbers of nymphs of two species of Stonefly are counted in each
of 10 samples taken at different positions along a stream.
307
Maths for Science
x1 − x2
t= (12.3)
SED
The term SED represents the ‘standard error of the differences in the sample
means’. SED is calculated using Equation 12.4, in which n1 and n2 are the two
sample sizes.
SC 2 SC 2
SED = + (12.4)
n1 n2
The term SC2 (which appears twice in Equation 12.4) represents the ‘common
population variance’. SC2 is calculated using Equation 12.5, in which s1 and s2
are the two estimated population standard deviations (also known as sample
standard deviations, as discussed in Section 11.2.7).
308
12 Statistical hypothesis testing
22.193 cm 2 22.193 cm 2
SED = +
14 16
= 1.724 cm
What does a value of t = −1.450 mean? Did the populations of plants growing
on the north-west and south-east slopes of this ridge really differ in mean
height or could the observed difference in mean height between the two
samples (i.e. 2.5 cm) have arisen by chance?
The fact that the test statistic t turns out to have a negative value can be
ignored. If it happened that the mean height of the sample of plants growing
on the north-west slope of the ridge had been subtracted from that of the
sample growing on the south-east slope, rather than the other way around,
then t would have been +1.450. Only the absolute value of t (i.e. the number
without its sign, in this case 1.450) is of any consequence.
The critical values of t are given in Table 12.12.
For the t-test for unmatched samples, the number of degrees of freedom
is given by
(n1 − 1) + (n2 − 1)
(14 − 1) + (16 − 1) = 13 + 15 = 28
Reading across the row corresponding to 28 degrees of freedom, it can been
seen that the test statistic (i.e. 1.450) is less than all the critical values.
Therefore all that can said is that P > 0.1. Since P is not less than 0.05, the
null hypothesis (that μ1 = μ2) cannot be rejected on the basis of the data
collected. There is therefore no evidence that the samples were taken from
different populations of plants. The plants growing on the north-west and
south-east slopes of this ridge do not differ statistically significantly from one
another in mean height.
309
Maths for Science
Table 12.12 Critical values of t for the t-test for unmatched samples for
different degrees of freedom and at three levels of significance.
Note: The null hypothesis is usually rejected if, for the appropriate number of
degrees of freedom, the calculated value of t is greater than the value tabulated
at the P = 0.05 significance level.
310
12 Statistical hypothesis testing
Descriptive statistics for the number of flowers per plant for samples of
plants growing on the north-west and south-east slopes of another ridge are
given in Table 12.13. Is there a statistically significant difference between
the slopes in the mean number of flowers per plant?
Table 12.13 Mean number of flowers per plant ( x ), estimated population standard
deviation of number of flowers per plant (s) and sample size (n) for a sample of
plants growing on the north-west slope of a ridge (Sample 1) and another sample
growing on the south-east slope (Sample 2).
311
Maths for Science
Question 12.7
1 16
2 13
3 15
4 17
5 15
6 24
total 100
A friend comments that the die must be biased. Carry out a χ2-test to
investigate this possibility. You should start by stating your null hypothesis
and you may need to use Table 12.3 in reaching your conclusion.
(b) The lengths and widths of a sample of beans are given in Table 12.15.
Calculate the Spearman rank correlation coefficient (rS) for these data and
use this to determine whether there is a statistically significant correlation
between length and width. You may need to use Table 12.7 in reaching
your conclusion.
312
12 Statistical hypothesis testing
Length/mm Width/mm
21.4 9.0
17.2 8.5
20.3 8.8
21.2 9.3
25.4 9.6
23.5 9.4
19.9 9.2
21.3 10.9
26.8 13.3
18.6 8.8
20.4 9.1
19.8 9.8
313
Questions: answers and comments
Comments on the answers are given in curly brackets { }. You are not
expected to have included them in your answers.
Chapter 1
Question 1.1
(a) (−3) − (−4) = (−3) + 4 = 1
(b) (−10) − (−5) = (−10) + 5 = −5
(c) 6 ÷ (−2) = −3
(d) (−12) ÷ (−6) = 2
Question 1.2
The lowest temperature in the oceans (the freezing point of seawater) is
31.9 °C colder than the highest recorded temperature, which is 30.0 °C.
Therefore, the freezing point of seawater is 30° C − 31.9 °C = −1.9° C.
Question 1.3
(a) 117 − (−38) + (−286) = −131
(b) (−1624) ÷ (−29) = 56
(c) (−123) × (−24) = 2952
Question 1.4
(a) The lowest common denominator is 6, so
2 1 2× 2 1 4 1 3
− = − = − =
3 6 3× 2 6 6 6 6
Dividing top and bottom by 3 gives
3 1
=
6 2
Alternatively,
2 1 2 × 6 1 × 3 12 3 9
− = − = − =
3 6 3 × 6 6 × 3 18 18 18
Dividing top and bottom by 9 gives the same answer as before
9 1
=
18 2
315
Maths for Science
1 1 2 1 × 10 1 × 15 2 × 6
+ − = + −
3 2 5 3 × 10 2 × 15 5 ×
6
10 15 12
= + −
30 30 30
13
=
30
(c) In this case, the lowest common denominator isn’t immediately obvious,
but a common denominator will certainly be given by the product of 3 and
28, so
5 1 5 × 3 1 × 28
− = −
28 3 28 × 3 3 × 28
15 28
= −
84 84
13
=−
84
Question 1.5
The original fraction is
4 1
= = 0.25
16 4
{You may have chosen any fraction. You may have also chosen any integer
for your calculations. In the following answers the number 2 is used, but the
principles hold good whatever choice of (non-zero) number is made.}
(a) Suppose you were to add 2 to the numerator and to the denominator
4+2 6
= = 0.333 to three decimal places
16 + 2 18
This is not the same as the original fraction. {There is just one special case in
which this kind of operation would not change the value of the fraction and
that is adding 0 to top and bottom, which obviously leaves the fraction
unchanged.}
(b) Suppose you were to subtract 2 from the numerator and from the
denominator
4−2 2
= = 0.143 to three decimal places
16 − 2 14
This is not the same as the original fraction. {Again, subtracting 0 from top
and bottom is the only case in which this operation leaves the fraction
unchanged.}
316
Questions: answers and comments
4 2
= = 0.5
16 4
This is not the same as the original fraction.
Question 1.6
(a)
2 2×3 6
×3 = =
7 7 7
(b)
5 5 1 5 ×1 5
÷7 = × = =
9 9 7 9 × 7 63
(c)
1/ 6 1 1 1 3 3 1
= ÷ = × = =
1/ 3 6 3 6 1 6 2
(d)
3 7 2 3× 7 × 2 42
× × = =
4 8 7 4 × 8 × 7 224
Dividing the top and bottom by 2, and then by 7 gives
21 3
42 42 21 21 3
= = = =
224 224 112 112 112 16 16
Alternatively, the original could have been simplified in the same way before
carrying out any multiplication:
3 7 1 2 1 3 × 1 ×1 3
× × = =
42 8 71 2 × 8 ×1 16
Question 1.7
(a) 70% expressed as a fraction is
70 7
70% = =
100 10
{Dividing the numerator and denominator by 10.}
317
Maths for Science
20 2 1
20% = = =
100 10 5
= =
25 25 × 4 100
Thus
11
= 44%
25
11
Alternatively, = 0.44 {found by dividing 11 by 25 on a calculator}.
25
(d) It is not easy to see what number the numerator and denominator of 8
Question 1.8
2
(a) 5
of 20 can be written as
4
2 2
20
× 20 = × = 2× 4 = 8
5
5 1
1
(b) 8
of 24 can be written as
3
7 7
24
× 24 = × = 7 × 3 = 21
8
8 1
1
15 15 12 3
12
×120 = × = × = 3 × 6 = 18
100 10 2
1
2 1
1
(d) 60% of 5 is
1
60
6
5 6
× 5
= × = = 3
100 10 2
1 2
318
Questions: answers and comments
Question 1.9
Increase in length = 69 cm – 60 cm = 9 cm
9
As a fraction of the baby’s original length this is .
60
9
= 0.15 and 0.15 ×100% = 15%
60
The percentage increase in length is 15%.
Question 1.10
(a)
1 1 1
2−2 = = =
2 2 2×2 4
{You might have gone one step further and expressed this in decimal notation
as 0.25.}
(b)
1
= 33 = 3 × 3 × 3 = 27
3−3
(c)
1 1
= =1
40 1
(d)
1 1 1
= = = 0.0001
104 10 × 10 × 10 × 10 10 000
Question 1.11
(a) 29 = 512
(b) 3–3 = 0.037 to three decimal places
{It doesn’t matter if you quoted more digits in your answer than this. There is
more explanation in Chapter 2 about how and when to round off the values
given on your calculator display.}
(c)
1
= 4−2 = 0.0625
42
Question 1.12
(a) 230 × 22 = 230+2 = 232
(b) 325 × 3−9 = 325+(−9) = 316
319
Maths for Science
(c)
102 ⎛ 1 ⎞
= 102 ÷103 = 102−3 = 10−1 ⎜ or ⎟
103 ⎝ 10 ⎠
(d)
102
= 102 ÷10−3 = 102−(−3) = 102+3 = 105
10−3
or alternatively
102 1
−3
= 102 × −3 = 102 ×103 = 105
10 10
(e) 10−4 ÷ 102 = 10−4−2 = 10−6
(f)
105 ×10−2
= 105+ ( −2) −3 = 100 (or 1)
103
Question 1.13
(a) (416)2 = 416×2 = 432
(b) (5−3)2 = 5(−3)×2 = 5−6
1
This could also be written as
56
Question 1.14
(a) 16 = (16)1/ 2 = 4
(b) 3
1000 = (1000)1/ 3 = 10
(c) 251/ 2 = 25 = 5
(d) 491/ 2 = 49 = 7
320
Questions: answers and comments
Question 1.15
(a) Using Equation 1.3
4× 1
(24 )1/ 2 = 2 2 = 22 = 4
(b) Using Equation 1.3
4× 1
104 = (104 )1 / 2 = 10 2 = 102 = 100
(c) Using Equation 1.3
Question 1.16
(a) Multiplication takes precedence over subtraction, so
35 − 5 × 2 = 35 − (5 × 2)
= 35 − 10
= 25
= 60
(c) Again, the brackets take precedence over the (implied) multiplication, so
5(2 − 3) = 5 × (−1) = −5
(d) Here the addition inside the square root sign (which acts as brackets) takes
precedence:
3 × 16 + 9 = 3 × 25 = 3 × 5 = 15
(e) Here the exponent takes precedence:
23 + 3 = 8 + 3 = 11
(f) Here both brackets take precedence over the (implied) multiplication:
(2 + 6)(1 + 2) = 8 × 3 = 24
321
Maths for Science
Question 1.17
If you had difficulty with any part of Question 1.17 you should refer
back to the section of the book indicated, and try the Additional
exercises on the accompanying website.
4 2 4×3 2×5
+ = +
5 3 5× 3 3× 5
12 10
= +
15 15
22
=
15
{See Section 1.2.2 Adding and subtracting fractions.}
(c) Expressed as a single fraction:
4 2 4×2 8
× = =
5 3 5 × 3 15
{See Section 1.2.4 Multiplying fractions.}
(d) Expressed as a single fraction:
4 2 4 3 42 3 2×3 6
÷ = × = × = =
5 3 5 2 5 21 5 × 1 5
{See Section 1.2.5 Dividing fractions.}
(e) Expressed as a percentage:
3 3 × 20 60
= =
5 5 × 20 100
Thus = 60%
5 3
Alternatively, = 0.6 and 0.6 × 100% = 60%
5
{See Section 1.2.6 Percentages.}
(f) 15% of 250 is
15 15 250 375
× 250 = × = = 37.5
100 100 1 10
{See Section 1.2.7 Calculating with fractions and percentages.}
(g) The increase is £56 − £50 = £6
322
Questions: answers and comments
(k) 41/ 2 = 4 = 2
So (41/2)4 = 24 = 2 × 2 × 2 × 2 = 16
Alternatively,
1 ×4
(41/ 2 ) 4 = 4 2 = 42 = 16
{See Section 1.3.4 Roots and fractional exponents.}
(l) The exponent takes precedence, then the multiplication, so
4 + 3 × 23 = 4 + 3 × 8 = 4 + 24 = 28
{See Section 1.4 Doing calculations in the right order.}
Chapter 2
Question 2.1
(a)
(b)
1
2.1×10−2 = 2.1×
100
2.1
=
100
= 0.021
323
Maths for Science
(c)
1
0.6 ×10−1 = 0.6 ×
10
0.6
=
10
= 0.06
Question 2.2
(a)
215 = 2.15 ×100
= 2.15 ×102
(b)
46.7 = 4.67 ×10
= 4.67 ×101
(c)
(d)
8.76
0.000 0876 =
100 000
8.76
= 5
10
= 8.76 ×10−5
Question 2.3
(a) 1 Gs = 109 s, so 3 Gs = 3 × 109 s
(b) 1 pm = 10−12 m, so 150 pm = 150 × 10−12 m
In scientific notation this is
= 1.
5 ×102 ×10−12 m
= 1.5 ×10−10 m
1
(c) 1 ms = 10−3 s, so 1 s = ms = 103 ms
10−3
So 0.8 s = 0.8 × 103 ms
324
Questions: answers and comments
8
0.8 ×103 ms = × 103 ms
10
= 8 × (103 ×10−
1) ms
= 8 × 103−1 ms
= 8 × 102 ms
(d) 1 μm = 10−6 m
1
1 nm = 10−9 m, so 1 m = nm = 109 nm
10−9
Then
8 μm = 8 ×10−6 m
= 8 × 10−6 ×109 nm
= 8 × 10−6+9 nm
= 8 × 103 nm
Question 2.4
(a) One million = 106, so the distance is
diameter = 2 × 6.
97 ×107 m
= 13.
94 ×107 m
= 1.394 ×108 m
~ 108 m
(c) 2π can be written as
2π = 2 × 3.
14 (to two decimal places)
= 6.28
This is greater than 5, so can be rounded up to the next power of ten to give
the order of magnitude, i.e. 2π ~ 10 (or 101).
(d) The mass of a carbon dioxide molecule is
325
Maths for Science
Question 2.5
(a) (i) 10−2 m = 0.01 m and 100 m = 1 m
so the difference between them is (1 − 0.01) m = 0.99 m.
(ii) 100 m = 1 m and 102 m = 100 m, so the difference between them is 99 m.
(iii) 102 m = 100 m and 104 m = 10 000 m, so the difference between them is
9900 m.
{Note that as you go up the scale, the interval between each successive pair of
tick marks increases by 100 times.}
(b) (i) The height of a child is about 100 m, i.e. 1 m. The height of Mount
Everest is about 104 m (actually 8800 m, but it is not possible to read that
accurately from the scale in Figure 2.4). So Mount Everest is ~104 times taller
than a child.
(ii) The length of a typical virus is 10−8 m and the thickness of a piece of
paper is 10−4 m, so the number of viruses that when laid end to end would
stretch across the thickness of the paper is:
10−4
∼ = 10−4−(−8) = 10−4+8 = 104
10−8
Question 2.6
Magnitude 7 on the Richter scale represents four points more than magnitude
3, and each point increase corresponds to a 10 times increase in maximum
ground movement. So a magnitude 7 earthquake corresponds to 104
(i.e. 10 000) times more ground movement than a magnitude 3 earthquake.
Question 2.7
(a) 1.274 99 m is 1.3 m to two significant figures.
{The sixth digit is a 9, so the fifth digit must be rounded up. 9 rounds up to
10 so the fourth digit must be rounded up too. However, don’t forget to
include the final zero in your answer as this is necessary to indicate that the
number is rounded to five significant figures.}
326
Questions: answers and comments
Question 2.8
(a) 0.004 123 m is 0.004 12 m to three significant figures, or 4.12 × 10−3 m in
scientific notation.
{The leading zeros are placeholders only, so the first significant digit is the 4.
The fourth significant digit is 3 so the third significant digit is unchanged.}
(b) 50.19 °C is 50.2 °C to three significant figures.
{The zero here counts as any other digit. The fourth digit is a 9, so the third
digit must be rounded up. The answer could also be correctly given as
5.02 × 101 °C in scientific notation.}
(c) To express 700.1 °C unambiguously to three significant figures, you should
start by expressing it in scientific notation:
Question 2.9
If you had difficulty with any part of Question 2.9 you should refer back
to the section of the book indicated, and try the corresponding
Additional exercises on the accompanying website.
327
Maths for Science
Thus
= 3.46 × 10−5 kg
Chapter 3
Question 3.1
(inch)2, cm2 and square miles all have units of (length)2, so they are all units
of area.
s2 cannot be a unit of area because the unit which has been squared, the
second, is a unit of time not of length.
m−2 cannot be a unit of area because the metre is raised to the power minus 2,
not 2.
km3 cannot be a unit of area because the kilometre is cubed not squared. In
fact, it is a unit of volume.
Question 3.2
(a) (i)
6.732
= 4.458 278146 = 4.46 to three significant figures
1.51
{6.732 is known to four significant figures, and 1.51 is known to three
significant figures. The number of significant figures in the answer is the same
as in the input value with the fewest significant figures, i.e. three.}
328
Questions: answers and comments
{2.0 and 0.025 are both given to two significant figures, so the answer is
given to two significant figures too. Note that this means that the answer
should be written as 0.050 not 0.05. See Section 2.4.}
(iii) 37.6 + 1.23 = 38.83 = 38.8 to one decimal place.
{Since 37.6 is given to one decimal place.}
(iv) The first step of the calculation is to express both numbers to the same
power of ten.
= 3.76 ×106
{Since the 3.71 in 3.71 × 106 is given to two decimal places, the base number
in the answer is given to two decimal places too.}
Alternatively the calculation could be done as follows:
= 376 ×104
= 3.76 × 106
{Since the 371 in 371 × 104 is to the nearest whole number, the 376 in
376 × 104 is given to the nearest whole number too.}
(b) The total mass = 3 × 1.5 kg = 4.5 kg.
{Note that you have exactly three bags of flour, so it would not be correct to
round the answer to one significant figure.}
Question 3.3
(a) Working to three significant figures and rounding to two significant figures
at the end of the calculation gives:
2
⎛ 4.2 ⎞
⎟ = (1.35) = 1.82 = 1.8 to two significant figures
2
⎜
⎝ 3.1 ⎠
{Squaring is repeated multiplication, so it is reasonable to quote the final
answer to two significant figures. However, working to two significant figures
throughout introduces a sizeable rounding error and gives a final answer of
2.0.}
(b)
= 0.
1637
= 0.16 to two significant figu
ures
329
Maths for Science
{The addition is done first and since the 1.2 is given to one decimal place, the
outcome of the addition is justified to one decimal place (1.6). In considering
the division, you then need to think in terms of significant figures. Both 1.6
and 9.5 have two significant figures, so the final result should be given to two
significant figures too. However, if you had rounded the intermediate step
from 1.555 to 1.6, a rounding error would have led to a final incorrect answer
of 0.17.}
Question 3.4
(a)
8 × 104 8 104
= ×
4 × 10−1 4 10−1
= 2 × 104−(−1)
= 2 × 105
(c)
Question 3.5
(a) The answer is likely to be given on your calculator as 4 × 1011. To an
appropriate number of significant figures (two) this should be written
(2.4 ×104 ) 2
= 4.0 ×1011
1.44 ×10−3
{If you obtained an answer of 4 × 105, this is probably because you
multiplied by 10−3 instead of dividing. You should use the special scientific
notation key on your calculator to enter 1.44 × 10−3.}
(b) The answer is likely to be given on your calculator as 1 × 1012. Strictly,
since the 104 implies that this value is only known to the nearest order of
magnitude, the answer should be written
330
Questions: answers and comments
{If you obtained an answer of 1 × 1013, this is probably because you entered
10 × 104 instead of 1 × 104 into your calculator. If you obtained an answer of
2.5 × 1019, you probably multiplied by 5 × 103 instead of dividing. If you
obtained an answer of 2.5 × 109, you probably multiplied by 10−5 × (5 × 103).
If you obtained a different incorrect answer, you may have made more than
one of these mistakes!}
Question 3.6
area = (9.78 ×10−3 m) 2
= (9.78 ×10−3 ) 2 m 2
= 9.56 ×10−5 m 2 to three signnificant figures
Question 3.7
To one significant figure,
distance to Proxima Centauri ≈ 4 × 1016 m
distance to the Sun ≈ 2 × 1011 m
Question 3.8
(i) and (iii) are equivalent. Multiplication is commutative, so
x(y + z) = (y + z)x.
(ii) and (v) are equivalent. Both multiplication and addition are commutative,
so xy + z = z + yx.
Note that (i) and (iii) are not equivalent to (ii) and (v). In (i) and (iii) the
whole of (y + z) is multiplied by x.
Substituting x = 3, y = 4 and z = 5 gives
(i) a = x(y + z) = 3 × (4 + 5) = 27
(ii) a = xy + z = (3 × 4) + 5 = 17
(iii) a = (y + z)x = (4 + 5) × 3 = 27
(iv) a = x + yz = 3 + (4 × 5) = 23
(v) a = z + yx = 5 + (4 × 3) = 17
331
Maths for Science
Question 3.9
The equivalent equations are (i) and (iii), since
bc 2 abc 2 bac 2
a = =
d d d
b 2c 2 b 2 a 2c 2
Note that only the c is squared, so (ii) m = a and (v) m =
d d
are different. Only the numerator of the fraction is multiplied by a, so
abc 2
(iv) m = is different too.
ad
Question 3.10
(a) length = 8.4 × 10−2 m
width = 5.7 × 10−2 m
height = 4.8 × 10−2 m
Substituting gives
volume = length × width × height
= 8.4 ×10−2 m × 5. 7 ×10−2 m × 4.8 ×10−2 m
= 2.298 × 10−4 m3
= 2.3 ×10−4 m3 to two significant figures
i.e. the volume of the block is 2.3 × 10−4 m3 to two significant figures.
(b) m = 0.620 kg
V = 2.298 × 10−4 m3
Substituting these values into Equation 3.9 gives
m
ρ=
V
0.620 kg
=
2.298 ×10−4 m3
= 2. 7 ×10 3 kg m −3 to two significant figures
i
i.e. the density of the block is 2.7 × 103 kg m−3.
{Since the values for length, width and height are given to two significant
figures, both answers should be given to two significant figures too. However,
to avoid rounding errors, the calculated value for V (from part (a)) is used to a
greater number of significant figures in the calculation in part (b).}
332
Questions: answers and comments
Question 3.11
m = 2.1 kg and a = 4.6 m s−2
Substituting into Equation 3.6 gives
F = ma
= 2.1 kg × 4.6 m s −2
= 9.66 kg m s −2
= 9.7 N since 1 N = 1 kg m s −2 (Box 3..3)
The force is 9.7 N and the units have worked out to be newtons, as expected
for a force.
Question 3.12
μ = 3.0 × 1010 N m−2
The answer obtained to Question 3.10 was ρ = 2.7 × 103 kg m−3 to two
significant figures. Good practice implies that you should work with an
intermediate value with at least one more digit than you will quote in the final
answer, but since the answer to Question 3.10 would have been
ρ = 2.70 × 103 kg m−3 to three significant figures, it doesn’t make a great deal
of difference on this occasion!
Substituting into Equation 3.10 gives
μ
vs =
ρ
3.0 ×1010 N m −2
=
2.70 ×103 kg m −3
and since
1 1
1 N = 1 kg m s −2 , m −2 = 2
and −3 = m3
m m
this can be written as
3.0 × 1010 kg m s −2 m3
vs =
2.70 ×103 kg m 2
3.0 ×1010 kg m s −2 m 2 × m
=
2.70 ×1
103 kg m 2
3.0 ×1010 kg m s −2 m 2 m
=
2.70 ×103 kg m 2
= 1.11×107 m 2 s −2
= 3.3 × 103 m s −1 to two significant figures
333
Maths for Science
So the S waves travel with a speed of 3.3 × 103 m s−1 through granite. The
units have worked out to be m s−1, as expected for a speed.
Question 3.13
If you had difficulty with any part of Question 3.13 you should refer
back to the section of the book indicated, and try the corresponding
Additional exercises on the accompanying website.
= 1.31 ×106
{Since the 1.10 in 1.10 × 106 is given to two decimal places, the base number
in the answer is given to two decimal places too.}
(iii)
(2.5 + 0.25) × (2.5 − 0. 25) = 2.75 × 2.25
= 6.
1875
= 6.2 to two significaant figures
{The addition and subtraction are done first and since the 2.5 is given to one
decimal place, the outcomes of the addition and subtraction are justified to
one decimal place (2.8 and 2.3). In considering the multiplication, you then
need to think in terms of significant figures. Both 2.8 and 2.3 have two
significant figures, so the final result should be given to two significant figures
too. However, if you had rounded in the intermediate step and calculated
2.8 × 2.3, a rounding error would have led to a final incorrect answer of 6.4.}
{See Section 3.1.2 Significant figures in calculations and Section 3.1.3
Avoiding rounding errors.}
(b) Using a calculator
1.012 ×108
= 1.0 ×106 to two significant figures
9.9 ×104 ×10−7
{This answer, to two significant figures, is the correct answer if you assume
the 10−7 is exactly known. If the 10−7 is only known to the nearest order of
magnitude, the correct answer is 106.
If you obtained an answer of 1.0 × 105, this is probably because you entered
10 × 10−7 instead of 1 × 10−7 into your calculator. If you obtained an answer
334
Questions: answers and comments
≈ 1 × 104+ 2
≈ 1 × 106
{See Section 3.2.1 Calculating in scientific notation without a calculator and
Section 3.3 Estimating answers.}
(d) m = 1 × 10−3 kg and c = 3.00 × 108 m s−1
Substituting in Equation 3.7 gives
E = mc 2
= 1 ×10−3 kg × (3.00 ×108 m s −1) 2
= 1 ×10−3 kg × (3.00 ×108 ) 2 (m s −1) 2
= 1 ×10−3 kg × 9.00 ×1016 m 2 s −2
= 9 × 1013 kg m 2 s −2
= 9 × 1013 J since 1 J = 1 kg m 2 s −2
So the energy corresponding to a mass deficit of 1 × 10−3 kg (1 g) is
9 × 1013 J, a huge amount.
{See Section 3.4 An introduction to symbols, equations and formulae, in
particular Section 3.4.4 Using equations.}
(e) G = 6.673 × 10−11 N m2 kg−2
m1 = 5.97 × 1024 kg
m2 = 7.35 × 1022 kg
r = 3.84 × 108 m
335
Maths for Science
Chapter 4
Question 4.1
f = 4.83 × 1014 Hz = 4.83 × 1014 s−1, since 1 Hz = 1 s−1 (from Box 3.3)
λ = 621 nm
Converting this value into SI base units (which must be done before the
calculation in order to give a meaningful result) gives
λ = 621 × 10−9 m = 6.21 × 10−7 m, since 1 nm = 10−9 m (Box 2.2)
Substituting into Equation 3.13 gives
v = fλ
= 4.83 ×1014 s −1 × 6.21 ×10−7 m
= 29.99 ×1014−7 m s −1
= 3.00 ×108 m s −1 to three significant figures
{Note that this is the speed of light in a vacuum. Light of this frequency and
wavelength is in the red part of the visible spectrum.}
Question 4.2
(a) 1 m = 100 cm, so 1 m2 = (100 cm)2 = 1002 cm2
Thus 1.04 m2 = 1.04 × 1002 cm2 = 1.04 × 104 cm2
(b) 1 m = 106 μm, so 1 m2 = (106 μm)2 = (106)2 μm2
336
Questions: answers and comments
Question 4.3
(a) 1 km = 103 m, so 1 km3 = (103 m)3 = (103)3 m3 = 109 m3
volume of Mars = 1.
64 ×1011 km3
= 1.
64 ×1011 ×109 m3
= 1.64 ×1020 m3
Question 4.4
(a) 1 cm = 10−2 m
So 1 cm day−1 = 10−2 m day−1
(applying the same conversion factor as in converting from cm to m).
Therefore
minute).
1
1 cm day −1 = cm s −1
8.64 ×104
337
Maths for Science
Therefore
12
12 cm day −1 = cm s −1
8.64 ×104
= 1.4 ×10−4 cm s −1 to two significaant figures
Question 4.5
(a) 1 mm = 10−3 m
1 year = 365 × 24 × 60 × 60 s = 3.154 × 107 s
To convert from mm year−1 to m s−1 you need to multiply by 10−3 (to convert
the mm into m) and divide by 3.154 × 107 s (to convert the year−1 into s−1).
10−3
1 mm year −1 = m s −1
3.154 ×107
so
10−3
0.1 mm year −1 = 0.1 × m s −1
3.154 ×107
= 3 × 10−12 m s −1 to one significant figure
Therefore, the stalactite is growing at about 3 × 10−12 m s−1.
(b) 1 cm = 10−2 m
1 day = 24 × 60 × 60 s = 8.64 × 104 s
To convert from cm day−1 to m s−1 you need to multiply by 10−2 (to convert
the cm into m) and divide by 8.64 × 104 (to convert the day−1 into s−1).
10−2
1 cm day −1 = m s −1
8.64 ×104
10−2
12 cm day −1 = 12 × m s −1
8.
64 ×104
= 1.4 ×10−6 m s −1 to two significant
g
figures
So the glacier is moving at about 1.4 × 10−6 m s−1.
(c) 1 km = 103 m
1 Ma = 106 year = 106 × 365 × 24 × 60 × 60 s = 3.154 × 1013 s
To convert from km Ma−1 to m s−1, you need to multiply by 103 (to convert
the km into m) and divide by 3.154 × 1013 (to convert the Ma−1 into s−1).
103
1 km Ma −1 = m s −1
3.154 ×1013
103
35 km Ma −1 = 35 × m s −1
3.154 ×1013
= 1.1×10−9 m s −1
to two signnificant figures
338
Questions: answers and comments
So the plates are moving apart at an average rate of 1.1 × 10−9 m s−1.
Comparing the answers to parts (a), (b) and (c) shows that the tectonic plates
are moving apart approximately 300 times faster than the stalactite is growing.
The glacier under consideration moves about 1000 times faster still, but
remember that there is considerable variation in the speeds at which all of
these processes take place.
Question 4.6
(a) 1 l = 103 ml
To convert from μg l−1 to μg ml−1 you need to divide by 103 (to convert the
l−1 into ml−1; the μg is unchanged).
1
1 μg l−1 = μg ml−1 = 10−3 μg ml−1
103
10 μg l−1 = 10 ×10−3 μg ml−1
= 1.0 ×10−2 μg ml−1 in scientific notationn
(b) Note that 10 μg l−1 = 10 μg dm−3 (since 1 litre is defined to be equal to
1 dm3) so the problem is one of converting 10 μg dm−3 to mg dm−3.
1 mg = 103 g so
1
1g = mg = 103 mg
10−3
1 μg = 10−6 g
Therefore 1 μg = 10−6 × 103 = 10−3 mg
To convert from μg dm−3 to mg dm−3 you need to multiply by 10−3 (to
convert the μg into mg; the dm−3 is unchanged)
1 μg dm−3 = 10−3 mg dm−3
10 μg dm −3 = 10 ×10−3 mg dm −3
on
So a concentration of 10 μg l−1 is equal to 1.0 × 10−2 mg dm−3.
(c) Note again that 10 μg l−1 = 10 μg dm−3, so the problem is one of
converting 10 μg dm−3 into g m−3.
1 μg = 10−6 g
1 dm = 10−1 m, so 1 dm3 = (10−1 m)3 = (10−1)3 m3 = 10−3 m3
To convert from μg dm−3 to g m−3 you need to multiply by 10−6 (to convert
the μg into g) and divide by 10−3 (to convert the dm−3 into m−3).
10−6
1 μg dm −3 = g m −3
10−3
339
Maths for Science
10−6
10 μg dm −3 = 10 × g m −3
10−3
= 10 ×10−6−(−3) g m −3
= 10 ×10−3 g m −3
1 −2 g m −3 in scientific notation
= 1.0 ×10
So a concentration of 10 μg l−1 is equal to 1.0 × 10−2 g m−3.
Question 4.7
g = 9.81 m s−2
The values given for ρ and h need to be converted into SI units:
1 km = 103 m
So
10−3
1 g cm −3 = kg m −3 = 10−3−(−6) kg m −3 = 103 kg m −3
10−6
So
P = ρ gh
= 1.025 ×103 kg m −3 × 9.81 m s −2 × 1.
2 ×103 m
kg × m × m 1 1
2 × 107 kg m −1 s −2
= 1.2
{Note that 1 kg m−1 s−2 is equivalent to 1 pascal (Pa) (Box 3.3) so a good
answer would say that the pressure is 1.2 × 107 Pa.}
340
Questions: answers and comments
Question 4.8
If you had difficulty with any part of Question 4.8 you should refer back
to the section of the book indicated, and try the corresponding
Additional exercises on the accompanying website.
75 cl = 75 ×10−2 dm3
= 75 ×10−2 ×10−3 m3
= 75 ×10−5 m3
= 7.5 ×10−4 m3 in
n scientific notation
{See Section 4.2 Converting units of area and volume.}
(b) 1 kg = 103 g
1 cm = 10−2 m, so
1
1m = cm = 102 cm
10−2
Therefore 1 m3 = (102 cm)3 = (102)3 cm3 = 106 cm3.
To convert from kg m−3 to g cm−3 you need to multiply by 103 (to convert
from kg to g) and to divide by 106 (to convert from m−3 to cm−3).
103
1 kg m −3 = g cm −3 = 103−6 g cm −3 = 10−3 g cm −3
106
9.1×102 kg m −3 = 9.1 ×102 ×10−3 g cm −3
= 9.1 ×102−3 g cm −3
= 9.1 ×10−1 g cm −3
Thus the density of olive oil is 9.1 × 10−1 g cm−3, which could also be
written as 0.91 g cm−3.
{You may have done this conversion simply by dividing by 1000 to convert
from kg m−3 to g cm−3.}
{See Section 4.4 Converting units of concentration and density.}
341
Maths for Science
Ek = 12
mv 2
= 12
× 7.5 × 10−3 kg × (416.67 m s −1) 2
= 651 kg (m s −1) 2
Chapter 5
Question 5.1
(a) v = f λ can be reversed to give f λ = v .
To isolate f you need to remove λ, and f is currently multiplied by λ so,
according to Hint 3, you need to divide by λ. Remember that you must do this
to both sides of the equation giving
fλ v
=
λ λ
342
Questions: answers and comments
The λ in the numerator of the fraction on the left-hand side cancels with the λ
in the denominator to give
v
f =
λ
(b) Etot = Ek + Ep can be reversed to give Ek + Ep = Etot.
To isolate Ek you need to remove Ep, and Ep is currently added to Ek so,
according to Hint 1, you need to subtract Ep. Remember that you must do this
to both sides of the equation giving
Ek + Ep − Ep = Etot − Ep
So
Ek = Etot − Ep (since Ep − Ep = 0)
m m
(c) ρ = can be reversed to give =ρ
V V
To isolate m you need to remove V, and m is currently divided by V so,
according to Hint 4, you need to multiply by V. Remember that you must do
this to both sides of the equation giving
mV
= ρV
V
The V in the numerator of the fraction on the left-hand side cancels with the V
in the denominator to give
m = ρV
Question 5.2
(a) b = c − d + e can be written as
c − d + e = b (with e on the left-hand side).
Adding d to both sides gives
c−d+e+d=b+d
i.e. c + e = b + d.
c+e−c=b+d−c
i.e. e = b + d − c
(b) P = ρgh can be written as
ρgh = P (with h on the left-hand side).
343
Maths for Science
= 2GM
2 gives
v esc
2
R 2GM
= 2
vesc
2 vesc
i.e.
2GM
R=
vesc
2
(d) E = hf − φ
Adding φ to both sides (to get φ on the left-hand side) gives
E + φ = hf − φ + φ
i.e. E + φ = hf
Subtracting E from both sides gives
E + φ − E = hf − E
i.e. φ = hf − E
344
Questions: answers and comments
bc 2
a=
d
can be written as
bc 2
=a
d
(with c on the left-hand side). Multiplying both sides by d gives
bc 2d
= ad
d
i.e. bc2 = ad
Dividing both sides by b gives
bc 2 ad
=
b b
i.e.
ad
c2 =
b
Taking the square root of both sides gives
ad
c=±
b
(f)
b
a=
c
can be written as
b
=a
c
(with b on the left-hand side). Squaring both sides gives
b
= a2
c
Multiplying both sides by c gives
bc
= a 2c
c
i.e. b = a2c
345
Maths for Science
Question 5.3
(a) Start by finding an equation for v 2 :
Ek = 12
mv 2 can be written as
1
2
mv 2
= Ek (with the v 2 on the left-hand side).
Multiplying both sides by 2 gives
mv 2 = 2Ek
Dividing both sides by m gives
2Ek
v2 =
m
Taking the square root of both sides gives
2Ek
v=±
m
You only need to consider the positive value on this occasion, i.e.
2Ek
v=
m
(b)(i) If Ek = 2 × 103 J and m = 4 × 1021 kg
2 × 2 ×103 J
v=
4 × 1021 kg
4 × 103 kg m 2 s −2
=
4 × 1021 kg
= 1 × 10−18 m 2 s −2
= 1 × 10−9 m s −1
{At this speed, the plate would move about 3 cm in a year.}
(ii) If Ek = 2 × 103 J and m = 70 kg
2 × 2 ×103 J
v=
70 kg
= 57.1 m 2 s −2
= 8 m s −1 to one significant figure
{The athlete, having a smaller mass, has to move a lot faster than the tectonic
plate.}
346
Questions: answers and comments
Question 5.4
(a) vx = u x + axt can be written as
u x + a xt = v x
Subtracting ux from both sides gives
a xt = v x − u x
Dividing both sides by t gives
vx − u x
ax =
μ
(b) Squaring both sides of vs = gives
ρ
vs 2 =
ρ
Multiplying both sides by ρ gives
ρvs
2 = μ
Dividing both sides by vs 2
gives
μ
ρ=
vs 2
L
(c) Multiplying both sides of F = by d2 gives
4πd 2
Fd 2 =
4π
Dividing both sides by F gives
L
d2 =
4πF
Taking the square root of both sides gives
L
d =±
4πF
{Note that if you consider just the positive value, this is the same as
Equation 3.20 from Box 3.1, albeit written rather differently.}
Question 5.5
(a)
μ0 i1i2 μ0 × i1i2 μ0i1i2
× = =
2π d 2π × d 2πd
347
Maths for Science
3a 3a
(b) Note that 2 means divided by 2.
2b 2b
3a 3a 1 3a
2= × =
2b 2b 2 4b
(c) The product c × b will be a common denominator, so you can write
2b 3c 2b × b 3c × c 2b 2 + 3c 2
+ = + =
c b c×b b×c cb
This is the simplest form in which this fraction can be expressed.
(d)
2ab 2ac 2ab b
÷ = ×
c b c 2ac
Cancelling the ‘2a’s gives
2ab 2ac 2a b b b × b b2
÷ = × = =
c b c 2a c c × c c 2
(e) The product f (f + 1) will be a common denominator, so you can write
1 1 ( f + 1) f
− = −
f f + 1 f ( f + 1) ( f + 1) f
f + 1 − f
=
f ( f + 1)
=
f ( f + 1)
2b 2 2c 2 2b 2 (a + c)
÷ = ×
(b + c) (a + c) (b + c) 2 c2
b 2 ( a + c)
=
c 2 (b + c)
This could also be written as follows but cannot be simplified further.
2
⎛ b ⎞ (a +
c)
⎜ c ⎟
(b + c)
⎝ ⎠
{Note that, for all parts of Question 5.5 and for many other questions
involving simplification, it is possible to check that the algebraic expression
you end up with is equivalent to the one that you started with by substituting
numerical values for the variables. For example, setting a = 2, b = 3 and
c = 4 in the expression in part (d) gives
348
Questions: answers and comments
2ab 2ac ⎛ 2 × 2 × 3 ⎞ ⎛ 2 × 2 × 4 ⎞
÷ =⎜ ⎟÷⎜ ⎟
c b ⎝ 4 ⎠ ⎝ 3 ⎠
16
= 3÷
3
= 3×
16
=
16
In the answer to part (d)
b 2 32 9
= =
c 2 42 16
So the values obtained for the original expression and the simplified version
are the same.}
Question 5.6
The equation can be written as
1 1 1
= +
f u v
v u
Question 5.7
(a)
1 1 1
(vx + u x )t = vxt + u xt
2 2 2
or alternatively
1 vt ut v t + u xt
(vx + u x )t = x + x or x
2 2 2 2
(b)
(a − b) − (a − c) a − b − a + c
=
2 2
−b + c
=
2
349
Maths for Science
(k − 2)(k − 3) = k 2 − 3k − 2k + 6
= k 2 − 5k + 6
(d)
(t − 2) 2 = (t − 2)(t − 2)
= t 2 − 2t − 2t + 4
= t 2 − 4t + 4
Question 5.8
(a) y2 − y = y(y − 1)
(b) x2 − 25 = (x + 5)(x − 5), by comparison with Equation 5.3.
You can check that the factorisation is correct by multiplying the brackets out.
This gives
(x + 5)( x − 5) = x 2 − 5x + 5x − 25
= x 2 − 25
Question 5.9
Both the terms on the right-hand side of Etot = 12 mv 2 + mg Δh include m, so
you can rewrite the equation as
Etot = m( 12 v 2 + g Δh)
m( 12 v 2 + g Δh) = Etot
350
Questions: answers and comments
Question 5.10
(a)(i) From the answer to Question 5.7(c)
k2 − 5k + 6 = (k − 2)(k − 3)
Thus, if k2 − 5k + 6 = 0, then (k − 2)(k − 3) = 0 too, so
k − 2 = 0 or k − 3 = 0
i.e. k = 2 or k = 3
Checking for k = 2:
k2 − 5k + 6 = 22 − (5 × 2) + 6 = 4 − 10 + 6 = 0, as expected.
Checking for k = 3:
k2 − 5k + 6 = 32 − (5 × 3) + 6 = 9 − 15 + 6 = 0, as expected.
t2 − 4t + 4 = (t − 2)2
Thus, if t2 − 4t + 4 = 0, then (t − 2)2 = 0 too, so
t − 2 = 0, i.e. t = 2
Checking your answer:
t = 2 gives t2 − 4t + 4 = 22 − (4 × 2) + 4 = 4 − 8 + 4 = 0, as expected.
So the only solution of the equation t2 − 4t + 4 = 0 is t = 2.
(b)(i) Comparison of k2 − 5k + 6 = 0 with ax2 + bx + c = 0 shows that a = 1,
b = −5 and c = 6 on this occasion, so the solutions are
−b ± b 2 − 4ac
k=
2a
−(−5) ± (−5) 2 − ( 4 × 1 × 6 )
=
2 ×1
5 ±
25 − 24
=
2
5 ±1
=
2
so
5 +1 6 5 −1 4
k= = = 3 or k = = =2
2 2 2 2
So the solutions of the equation k2 − 5k + 6 = 0 are k = 2 and k = 3. This is
the same answer as was obtained in part (a) and could be checked in the same
way.
351
Maths for Science
−b ± b 2 − 4ac
t=
2a
−(−4) ± (−4) 2 − (4 × 1 × 4)
=
2 ×1
4 ±
16 − 16
=
2
4±0
=
2
= 2
Question 5.11
(a) Rearranging p = mv to make v the subject, by dividing both sides by m,
gives
p
v=
m
Substituting in Ek = 12 mv 2 gives
2
1 ⎛ p⎞
Ek = m
2 ⎜⎝ m ⎟⎠
1 p2
= m
2 m2
p2
=
2m
(b) Since both equations are already written with E (the variable you are
trying to eliminate) as the subject, you can simply set the two equations for E
equal to each other:
1
2
mv 2 = mg Δh
v
1 2
2
= g Δh
v 2 = 2g Δh
Taking the square root of both sides of the equation gives
v = ± 2g Δh
352
Questions: answers and comments
Question 5.12
If you had difficulty with any part of Question 5.12 you should refer
back to the section of the book indicated, and try the corresponding
Additional exercises on the accompanying website.
u 2 + 2as = v 2
Subtracting u2 from both sides gives
2as = v 2 − u 2
Dividing both sides by 2a gives
v2 − u 2
s=
2a
{See Section 5.1 Rearranging Equations.}
(b)
mv 2 v mv 2 2
÷ = ×
2 2 2 v
m×v× v
=
v
= mv
{See Section 5.2.1 Simplifying algebraic fractions.}
353
Maths for Science
(c)
(x + 1) 2 = (x + 1)(x + 1)
= x 2 + 1x + 1x +1
= x 2 + 2x + 1
{See Section 5.2.2 Using brackets in algebra.}
(d) You may have spotted (from the answer to part (c)) that
x2 + 2x + 1 = (x + 1)2
Therefore, if x2 + 2x + 1 = 0, it follows that (x + 1)2 = 0
i.e. x + 1 = 0, so x = −1 is the only solution.
Alternatively, using the quadratic equation formula, with a = 1, b = 2 and
c = 1 gives
−b ± b 2 − 4ac
x=
2a
−2 ± 22 − (4 × 1 × 1)
=
2 ×1
−2 ± 4 − 4
=
2
−2 ± 0
=
2
= −1
expected.
(e) Since both equations are already written with E (the variable you are
trying to eliminate) as the subject, you can simply set the two equations for E
equal to each other:
1
2
mv 2 = mcΔT
v
1 2
2
= cΔT
v 2 = 2cΔT
Taking the square root of both sides of the equation gives
354
Questions: answers and comments
v = ± 2cΔT
{See Section 5.3 Combining equations.}
Chapter 6
Question 6.1
Let the number selected be represented by x
• Adding 5 gives x+5
• Doubling the result gives 2(x + 5) = 2x + 10
• Subtracting 2 gives (2x + 10) − 2 = 2x + 8
2x + 8
• Dividing by 2 gives = x+ 4
2
• Taking away the number you (x + 4) − x = 4
first thought of gives
Question 6.2
Let H represent Helen’s height in cm and R represent Rupa’s height in cm.
R = H + 15 (i)
The height of the wall is equal to Rupa’s height up to her shoulders (R − 25)
plus Helen’s height up to her eyes (H − 10), thus Helen
R + H − 35 = 300
Adding 35 to both sides gives
R + H = 335 R
R−25
Substituting for R from (i) gives
(H + 15) + H = 335
Rupa
2H + 15 = 335
355
Maths for Science
2H = 320
Dividing both sides by 2 gives
Question 6.3
Following the approach described in Worked example 6.3.
v
1 2
2
= g Δh
v 2 = 2g Δh
Taking the square root of both sides gives
v = ± 2g Δh
On this occasion you only need to consider the positive square root,
i.e. v = 2g Δh .
Substituting values
Substituting Δh = 1.8 m and g = 9.81 m s−2 gives
v = 2 × 9.81 m s −2 ×1.8 m
= 5.9 m s −1 to two significant figures
(noting that m 2 s −2 = m s −1 ).
356
Questions: answers and comments
An estimated value is
v ≈ 2 × 10 m s −2 × 2 m
≈ 40 m 2 s −2
≈ 6 m s −1, since 40 ≈ 36
The speed seems quite high; in reality not all of the child’s gravitational
potential energy would be converted into kinetic energy.
Question 6.4
The units of m are kg and the units of c are m s−1, so the units of mc2 are
kg × (m s−1)2 = kg m2 s−2
Therefore, since E = mc2
Question 6.5
(a) From Equation 6.9, possible definitions are
(b) The units of force are newtons (N) and the units of area are m2, so the
N [units of F]
= N m −2
m 2 [units of A]
Question 6.6
(a) The following aspects of the answer could be improved:
. Insufficient working has been shown.
. The values for r3 and r have been linked by an equals sign; these values
are not equal.
. The abbreviation ‘e19’ has been used to indicate scientific notation (and
also, * has been used instead of ×). You are advised to use more
conventional notation, i.e. ‘× 1019’ and ‘×’.
. 4
3
has been expressed as 1.33 and π has been expressed as 3.14; both of
these values are true but only to three significant figures, thus introducing
the potential for rounding errors.
. It is not clear that the division applies to both 1.33 and 3.14 (as an
alternative to the approach shown in part (b), this could be corrected by
using brackets).
. The answer is given to too many significant figures. Since V was given to
three significant figures, the answer should be given to three significant
357
Maths for Science
4πr3 = 3V
Dividing both sides by 4π gives
3V
r3 =
4π
{It would have been perfectly acceptable to divide by 4 and π as two separate
steps.}
Taking the cube root of both sides gives
3V
r= 3
4π
{You might have written this as
1/ 3
⎛ 3V ⎞
r =⎜ ⎟ }
⎝ 4π ⎠
V = 6.09 × 1019 m3 so
3V
r= 3
4π
3 × 6.09 ×1019 m3
= 3
4π
= 3 1.
454 ×1019 m3
358
Questions: answers and comments
Question 6.7
If you had difficulty with any part of Question 6.7 you should refer back
to the section of the book indicated, and try the corresponding
Additional exercises on the accompanying website.
mg Δh = 12 mv 2
g Δh = 12 v 2
v2
Δh =
2g
Substituting values
Substituting v = 5.6 m s−1 and g = 9.81 m s−2 gives
(5.6 m s −1) 2
Δh =
2 × 9.81 m s −2
(5.6) 2 (m s −1) 2
=
2 × 9.81 m s −2
(5.66) 2 m m s −2
=
2 × 9.81 m s −2
(5.6) 2 m m s −2
=
2 × 9.81 m s −2
= 1.6 m to tw
wo significant figure
s
The starting height is 1.0 m, so the ball rises to a total height of
359
Maths for Science
(5 m s −1) 2
Δh ≈
2 × 10 m
s −2
25
≈ m
20
≈ 1.25 m
which is fairly close to the accurately calculated value (1.6 m) for Δh.
A height of 2.6 m is reasonable for a ball being thrown into the air.
{See Section 6.2 Using algebra to solve scientific problems and Section 6.4
Writing maths.}
strain
strain rate =
time
Since extension and original length are both lengths, with SI units of metres,
these would cancel out so strain has no units. Time has SI units of seconds, so
the SI units of strain rate are
1
= s −1
s
{See Section 6.3.2 Equation to definition and Section 6.3.1 Equation to units.}
Chapter 7
Question 7.1
(a) (i) The red lines on Figure 7.41 show that, by interpolation, when
current = 1.5 A then voltage = 2.0 V.
(ii) The line through the data points can be extended at each end, as shown in
Figure 7.41. This process of extrapolation to the vertical axis shows that when
the current is zero the voltage has a value of 5.0 V.
(b) (i) The green lines of Figure 7.41 show that, by interpolation, when
voltage = 3.4 V then current = 0.8 A.
(ii) Extrapolation to the horizontal axis shows that when the voltage is zero
the current has a value of 2.5 A.
360
Questions: answers and comments
5.0
4.0
voltage/V
3.0
2.0
1.0
Question 7.2
Using the red lines in Figure 7.42,
rise
gradient =
run
(170 − 10) km
=
(32 − 4) s
160 km
=
28 s
= 5.7 km s −1
Therefore speed = 5.7 km s−1 to two significant figures.
{You are likely to have chosen different points from which to calculate your
gradient, but you should have obtained a similar answer.}
200
distance from
epicentre/km
170
100
10
0 10 20 30 40
4 32
travel time after earthquake occurred/s
361
Maths for Science
Question 7.3
40 rise
gradient =
run
(32 − 2) s
32 =
30 (170 − 0) km
30 s
=
20
170 km
= 0.176 s km −1
Therefore
10
1
speed =
2 0.176 s km −1
0 100 200 = 5.7 km s −1 to two significant figures
170
distance from To the precision to which it is possible to read the graph, this is the same
epicentre/km
value as before.
Figure 7.43 For use with
answer to Question 7.3. Question 7.4
Using the red lines in Figure 7.44,
rise
gradient =
20 run
temperature/°C
(7 − 20) °C
=
(2.0 − 0) km
10 −13 °C
7 =
2.0 km
= −6.5 °C km −1
0 1 2
altitude/km You could equally correctly have written this as −6.5 °C/km. The negative
value of the gradient implies that temperature decreases with increasing height
above sea level and your sentence should reflect this. For example you could
Figure 7.44 For use with write: ‘For each successive kilometre of height gained above sea level, the
answer to Question 7.4.
atmospheric temperature falls by 6.5 °C’.
362
Questions: answers and comments
Question 7.5
Using the red lines in Figure 7.45,
rise
gradient =
run
[(−50 − (−20)] °C
=
(10 − 5.5) km
−30 °C
=
4.5 km
= −6.77 °C km −1
10
5 10
0
altitude/km
temperature/°C
−10
−20
−30
−40
−50
−60
This agrees quite well with the value obtained in the answer to Question 7.4.
In fact temperature does decrease with altitude at an almost constant rate
through the troposphere.
Question 7.6
The line corresponding to v = rz has the larger (steeper) gradient. Therefore
r is larger than s.
Question 7.7
If two quantities are directly proportional to each other, a graph in which one
is plotted against the other will be a straight line through the origin.
Therefore, only (c) in Figure 7.22 corresponds to a proportional
relationship: u ∝ z . In this case, the gradient is negative, i.e. the constant of
proportionality is negative.
363
Maths for Science
Question 7.8
The line goes through the origin and when m = 50 kg, WM = 80 N. Therefore
rise
gradient =
run
(80 − 0) N
=
(50 − 0) kg
80 kg m s −2
=
50 kg
= 1.6 m s −22
The equation WM = mgM can be written WM = gMm and comparison with
y = mx + c, or y = kx for a graph which goes through the origin, shows that
gradient = gM on this occasion. Thus the magnitude of the acceleration due to
gravity on the Moon is 1.6 m s−2 to two significant figures.
Question 7.9
Since M is directly proportional to d3, these are the quantities to plot. The
spheres are selected and then their masses are measured, so d is the
independent variable, and so according to convention d3 should therefore be
plotted on the horizontal axis. In other words, the convention would be to plot
M against d3. Comparison with y = mx + c, or y = kx for a graph which goes
through the origin, shows that on this occasion
πρ
gradient = 16 πρ or
6
{If you chose to defy convention and plot d3 against M, the gradient would
6
have the reciprocal value, i.e. .}
πρ
Question 7.10
After n half-lives, the number of radioactive atoms is reduced to ( 12 ) n of the
original number. Since 161 = ( 12 ) 4 , four half-lives must elapse before the
number of radioactive atoms will be 161 of the number there are today. So
4 × 1600 years = 6400 years must elapse for this to happen.
Question 7.11
If you had difficulty with any part of Question 7.11 you should refer
back to the section of the book indicated, and try the corresponding
Additional exercises on the accompanying website.
(a) Figure 7.46 shows that a length of 13.0 cm (13.0 × 10−2 m) corresponds to
a mass of 100 × 10−3 kg.
364
Questions: answers and comments
14
12
10
8
L/10−2 m
0 20 40 60 80 100
m/10−3 kg
{See Section 7.1.2 Graphs. Note that it is important to read the powers of ten
and the units from the axis labels.}
(b) When m = 0 kg, L = 6.8 × 10−2 m and when m = 100 × 10−3 kg,
L = 13.0 × 10−2 m (see Figure 7.46), so
rise
gradient =
run
=
(100 − 0) × 10−3
kg
6.2 ×10−2
m
=
100 ×10−3 kg
= 0. 62 m kg −1
365
Maths for Science
y = m x + c
L = S m + L0
shows that the gradient of the graph gives S and the intercept on the vertical
axis gives L0.
Thus L0 = 6.8 × 10−2 m and S = 0.62 m kg−1.
L0 represents the length of the spring when no masses have been added,
i.e. the unextended length.
(d) Since d is directly proportional to t2, these are the variables to plot. The
description of the experiment shows that d is the independent variable, which
according to convention should be plotted on the horizontal axis.
The equation d = 12 gt 2 can be rearranged to give
2d
t2 =
g
which can also be written
2
t2 = d
g
Comparison with y = mx + c, or y = kx for a graph which goes through the
origin, shows that on this occasion
2
gradient =
g
and so
2
g=
gradient
g
{If you chose to plot d against t2, then the gradient would be , in which
case g = 2 × gradient.} 2
Chapter 8
Question 8.1
(a) The gradient of the tangent drawn to the graph of y = x2 (Figure 8.2) at
x = 1 is
rise (3.0 − 0.0) 3.0
gradient = = = = 2.0
run (2.0 − 0.5) 1.5
366
Questions: answers and comments
y
18
16
14
12
10
6
tangent
at x = 2
4
2 tangent
at x = 1
1 2 3 4 5 x
Question 8.2
(a) y = x4, so C = 1 and n = 4, and
dy
= 1 × 4x 4−1 = 4x3
dx
When x = 4,
dy
= 4 × 43 = 44 = 256
dx
So at x = 4 the gradient of the graph is 256.
367
Maths for Science
dy
= 5x1−1 = 5x 0 = 5
dx
The gradient of the graph is 5 for all values of x.
{You may have been able to give this result without differentiating y = 5x,
from your knowledge of the gradient of straight-line graphs.}
(c) y = 3x2, so C = 3 and n = 2, and
dy
= 3 × 2 x 2−1 = 6x1 = 6x
dx
When x = 4,
dy
= 6 × 4 = 24
dx
So at x = 4 the gradient of the graph is 24.
(d) y = 5, so C = 5 and n = 0, and
dy
= 5 × 0 × x −1 = 0
dx
The gradient of the graph is 0 for all values of x.
{You may have been able to give this result without differentiating, by
remembering that y = 5 is a horizontal line, and the gradient of a horizontal
line is always zero.}
Question 8.3
(a)
1 1
y= = 1 / 2 = x −1/2
x x
so C = 1 and n = − 12
dy 1 1 x −3/ 2 1
= − x − 2 −1 = − = − 3/ 2
dx 2 2 2x
Since x3/2 = x1+1/2 = x1 × x1/2, this could also be written as
dy 1
=−
dx 2x x
When x = 4,
dy 1 1 1
=− =− =−
dx 2×4× 4 2×4×2 16
So at x = 4 the gradient of the graph is − 161 or −0.0625.
368
Questions: answers and comments
(b)
2
y= = 2x −2
x2
so C = 1 and n = −2
dy 4
= 2 × (−2) x −2−1 = −4x −3 = − 3
dx x
When x = 4,
dy 4 1 1
=− 3 =− 2 =−
dx 4 4 16
So at x = 4 the gradient of the graph is − 161 or −0.0625.
Question 8.4
dx
(a) x = t7, so = 7t 7 −1 = 7t 6
dt
(b)
C
E= = Cr −1
r
so
dE C
= C × (−1)r −1−1 = −Cr −2 = − 2
dr r
Question 8.5
z = 4y2 + 2y
Differentiating each of the terms separately gives
dz
= (4 × 2y 2−1) + (2 × 1y1−1)
dy
= 8y1 + 2y 0
= 8y + 2
When y = 0,
dz
= (8 × 0) + 2 = 0 + 2 = 2
dy
So at y = 0 the gradient of the graph is 2.
Question 8.6
(a) x = 2t3 + 4t2 − 2t + 3
369
Maths for Science
d2x
= (6 × 2t) + 8 = 12t + 8
dt 2
(b)
2
z= = 2y −1
y
Differentiating with respect to y gives
dz 2
= 2 × (−1) y −1−1 = −2y −2 = − 2
dy y
Differentiating again gives
d2z 4
= −2 × (−2) y −2−1 = 4y −3 = 3
dy 2 y
Question 8.7
(a) y = 5e2x so C = 5 and k = 2
dy
= 5 × 2e 2x = 10e 2x = 2y (since y = 5e 2x )
dx
(b) z = et/2 so C = 1 and k = 1
2
dz 1 t/2 z
= e = (since z = et/2 )
dt 2 2
Question 8.8
If you had difficulty with any part of Question 8.8 you should refer back
to the section of the book indicated, and try the corresponding
Additional exercises on the accompanying website.
(a) x = 3t2 + 2t + 5
Differentiating the terms separately gives
dx
= (3 × 2t 2−1) + (2 × 1t1−1) + 0
dt
= 6t1 + 2t 0
= 6t +
2
370
Questions: answers and comments
When t = 2,
dx
= ( 6 × 2) + 2 = 12 + 2 = 14
dt
So the gradient of the graph is 14 at t = 2.
{See Section 8.2.1 The principles of differentiation, Section 8.2.2
Differentiation by rule, Section 8.2.3 Using different symbols and different
notation and Section 8.2.4 Differentiating sums.}
(b)
2
y=
x2
can be written y = 2x−2, so
dy
= 2 × (−2) x −2−1
dx
= −4 x
−3
4
=− 3
x
From
dy
= −4x −3
dx
d2 y
= (−4) × (−3)x −3−1
dx 2
= 12x −4 (since multiplying two negativee numbers gives a positive)
12
= 4
x
Note: multiplying two negative numbers is covered in Section 1.1.2.
{See Section 8.2.2 Differentiation by rule and Section 8.2.5 Second
derivatives.}
(c) y = 4e3x so the first derivative is
dy
= 4 × 3e3x = 12e3x
dx
dy
as y = 4e3x, this could also be written = 3y
dy dx
Since = 12e3x , the second derivative is
dx
d2 y
= 12 × 3e3x = 36e3x
dx 2
371
Maths for Science
d2 y
= 9y
dx 2
{See Section 8.3 Differentiating exponential functions.}
Chapter 9
Question 9.1
(a) 2π radians = 360°, so
360°
1 radian =
2π
360°
0.123 radians = 0.123 ×
2π
≈ 7.05° to three significant figures
r
(b) 2π radians = 360°, so
2π 360°
radians = = 120°
3 3
(c) 2π radians = 360°, so π radians = 180°,
3π 3 × 180°
radians = = 270°
2 2
Question 9.2
(a) 360° = 2π radians, so
2π
1° = radians
360
2π
2π
90° = 90 × radians
360
π
= radians
2
372
Questions: answers and comments
{This answer could have been written as 1.6 radians (to two significant
figures), but note that π/2 radians is an exact answer which 1.6 radians is
not.}
(c) 360° = 2π radians, so
2π
1° = radians
360
2π
Question 9.3
(a) You need to find length a in Figure 9.33. wall
From Pythagoras’ Theorem
θ
a2 + (1.15 m)2 = (4.50 m)2
so
4.50 m
a2 = (4.50 m)2 − (1.15 m)2 a
ladder
a = 20.25 m 2 − 1.3225 m 2
= 18.
9275 m 2
= 4.35 m to three significant figures
g
75.2°
(b) Now you need to find angle θ in Figure 9.33.
ground 1.15 m
The interior angles in a triangle add up to 180°, so
Figure 9.33 For use with
θ + 75.2° + 90° = 180°
answer to Question 9.3.
i.e.
θ = 180° − 75.2
° − 90°
= 14.8°
373
Maths for Science
Question 9.4
The lengths of the sides are as follows:
a = 2.0 cm, b = 1.5 cm, h = 2.5 cm
a′ = 4.0 cm, b′ = 3.0 cm, h′ = 5.0 cm
a″ = 6.0 cm, b″ = 4.5 cm, h″ = 7.5 cm
Therefore
b 1.5 cm
= = 0.75
a 2.0 cm
b′ 3.0 cm
= = 0.75
a′ 4.0 cm
b′′ 4.5 cm
= = 0.75
a′′ 6.0 cm
From this you can see that
b b′ b′′
= =
a a′ a′′
b 1.5 cm
= = 0.60
h 2.5 cm
b′ 3.0 cm
= = 0.60
h′ 5.0 cm
b′′ 4.5 cm
= = 0.60
h′′ 7.5 cm
so you can see that
b b′ b′′
= =
h h′ h′′
a 2.0 cm
= = 0.80
h 2.
5 cm
a′ 4.
0 cm
= = 0.80
h′ 5.0 cm
a′′ 6.0 cm
= = 0.80
h′′ 7.5 cm
again you can see that
a a′ a′′
= =
h h′ h′′
374
Questions: answers and comments
Question 9.5
(a) sin 49° = 0.7547
π
(b) cos = 0.9239
8
π
(c) tan =1
4
{Since the angles in parts (b) and (c) are given in radians, your calculator
needs to be in ‘radians mode’ in order to obtain the correct answers to these
parts.}
Question 9.6
(a) cos−1(0.5253) = 58.31°
(b) tan−1(1.5574) = 1.0000 radians
{Your calculator needs to be in ‘degrees mode’ for part (a) and in ‘radians
mode’ for part (b) in order to obtain the correct answers.}
Question 9.7
(a)
opp
tan θ =
adj
5.0 m
=
1.
0 m
= 5.0
4.3 m
h=
cos 32°
4.3 m
=
0.8480
375
Maths for Science
(c)
opp
sin θ =
hyp
so
π a
sin =
3 10 m
Multiplying both sides by 10 m gives
π
a = 10 m × sin
3
= 10 m × 0.
8660
{Note that ‘opp’ and ‘adj’ must be the sides opposite and adjacent to the
angle you are trying to find.}
Question 9.8
The height H can be found from
H = height of West Tower + height of base of Cathedral
− height of theodolite
= 66 m + 15 m − 1.
5 m
= 79.5 m
You know θ = 2.27°, so you can use the following equation to find D.
H
tan θ =
D
Multiplying both sides by D gives
D tan θ = H
Dividing both sides by tan θ gives
H
D=
tan θ
79.5 m
=
tan 2.27°
= 200
6 m
≈ 2000 m
So you can estimate the distance of the theodolite from Ely Cathedral to be
about 2 km.
376
Questions: answers and comments
Question 9.9
Using Equation 9.10,
V = W tan θ
where W = 65 m and θ = 36°, the vertical thickness is
V = 65 m × tan 36°
Question 9.10
From Equation 9.11,
r = h cos 45°
where r is the required radius and h = 302 pm = 302 × 10−12 m or
3.02 × 10−10 m. Thus
Question 9.11
You know i = 45.0°, r = 26.3°, v1 = 3.00 × 108 m s−1, and you are trying to
find v2 , the speed of light in glass.
Snell’s Law (Equation 9.12) states that
sin i v1
=
sin r v2
Multiplying both sides by v2 and by sin r gives
v 2 sin i = v1 sin r
Dividing both sides by sin i gives
sin r
v2 = v1
sin i
sin 26.3°
= 3.00 ×108 m s −1 ×
sin 45.0°
0..4431
= 3.00 ×108 m s −1 ×
0.7071
= 1. 88 ×108 m s −1
So the speed of light in glass is 1.88 × 108 m s−1.
377
Maths for Science
Question 9.12
nλ = d sin θn
Dividing both sides by n
d sin θ n
λ=
n
Substituting d = 1.64 × 10−6 m, θn = 24.1° and n = 1
ures
So the wavelength of the light is 6.70 × 10−7 m.
Question 9.13
You know
adj vx
cos α = =
hyp v
so
vx = v cos α
= 8.6 m s −1 × cos 42°
= 6.4 m s −1 to two significant figures
opp v y
sin α = =
hyp v
so
v y = v sin α
= 8.6 m s −1 × sin 42°
= 5.8 m s −1 to two significant figures
378
Questions: answers and comments
Question 9.14
F = Fx 2 + Fy 2
= (4.0 N) 2 + (3.
0 N) 2
Fy F
= 5.
0 N
β
opp
tan β = Fx x
adj
Fy
=
Fx Figure 9.34 For use with
Question 9.14.
3.0 N
=
4.0 N
= 0.75
Question 9.15
First of all convert 2.5° into radians.
2π
360° = 2π, so 1° =
360
2π
2.5° = 2.5 ×
360
ficant figures
sin θ ≈ θ for all small angles measured in radians, so sin 2.5° ≈ 0.044
cos θ ≈ 1 for all small angles
tan θ ≈ θ for all small angles measured in radians, so tan 2.5° ≈ 0.044
{Checking on a calculator shows that sin 2.5° = 0.043 62 and
tan 2.5° = 0.043 66 to four significant figures, so the small angle
approximations are good estimates.}
379
Maths for Science
Question 9.16
Let the distance to car ferry = L
The length of car ferry = D = 86 m
The angle subtended = φ = 3.5°
Converting φ into radians:
360° = 2π radians, so
2π
1° =
360
2π
3.5° = 3.5 ×
360
= 0.0611 radians
86 m
L≈
0.
0611
≈ 140
8 m
The ferry is approximately 1.4 km away.
Question 9.17
If you had difficulty with any part of Question 9.17 you should refer
back to the section of the book indicated, and try the corresponding
Additional exercises on the accompanying website.
(a) {There are several different ways of tackling this question. The method
given uses Pythagoras’ Theorem to find h and trigonometry to find θ, then
uses the fact that the internal angles in all triangles add up to 180° to find φ .}
380
Questions: answers and comments
θ + φ + 90° = 180°
i.e.
φ = 180° − 90° − 67°
= 23°
{See Section 9.2 A quick look at triangles and Section 9.3 Calculating with
angles: trigonometry.}
(b) From the definition of cosine:
adj Fx
cos θ = =
hyp F
so
F x = F cos θ
= 82 N × cos 27°
= 82 N × 0.
891 01
ures
Similarly
opp Fy
sin θ = =
hyp F
so
F y = F sin θ
= 82 N × sin 27°
= 82 N × 0.
453 99
ures
{See Section 9.4 Using trigonometry and Pythagoras’ Theorem to help with
vectors.}
381
Maths for Science
2π
1° =
360
2π
0.5° = 0.5 ×
360
= 8.7266 ×10−3 radians
Substituting gives
Chapter 10
Question 10.1
(a) Since 100 = 102, log10 100 = 2
(b) Since 0.001 = 10−3, log10 0.001 = −3
1
(c) Since 10 = 101/ 2 , log10 10 =
2
(d) Since 1.329 = 100.1235 (from the section of text just above the question)
log10 1.329 = 0.1235
Question 10.2
(a) log10 2 = 0.3010
(b) log10 2000 = 3.3010
{Note that log10 2000 is exactly 3 greater than log10 2. This result will be
discussed further in Sections 10.2 and 10.3.}
382
Questions: answers and comments
Question 10.3
(a) 101.5 = 31.62
(b) p = 31.62
{Because of the way in which log to base 10 is defined, the answer to (b)
follows straight from the answer to part (a).}
Question 10.4
(a) For human blood the hydrogen ion concentration is 4.0 × 10−8 mol dm−3,
so
−8
= − log10 (4.0 × 10 )
= −((−7.4)
= 7.
4
(b) For the hair shampoo, the hydrogen ion concentration is
3.2 × 10−6 mol dm−3, so
−6
= − log10 (3.2 × 10 )
= −((−
5.5)
= 5.
5
Question 10.5
(a)
log10 300 = log10 (3 ×100)
= log10 3 + log10 100 (from Equation 10.2)
= 0.4771 + log10 102
= 0.4771 + 2 (from Equation 10.1)
= 2.4771 to fo
our decimal places
(b)
log10 0.03 = log10 (3 ÷100)
= log10 3 − log10 100 (from Equation 10.3)
= 0.4771 − log10 10
2
= 0.
4771 − 2 (from Equation 10.1)
383
Maths for Science
(c)
= 2 × 0.
4771
Question 10.6
log10 (2x3 ) = log10 2 + log10 x3 (from Equation 10.2)
= log10 2 + 3 log110 x (from Equation 10.4)
Thus the three expressions that are equivalent are (i), (ii) and (vi).
Question 10.7
2.5 − 0.5 2.0
The gradient of the line = = = 2.0
1.0 − 0.0 1.0
This is the result expected.
Question 10.8
Taking the log to base 10 of both sides of the equation y = 2x3, as in Question
10.6, gives
= log100 2 + 3 log
10 x (from Equation 10.4)
Reversing the order of the two terms on the right-hand side gives
Question 10.9
n = n0eat
Taking the log to base 10 of both sides of the equation gives
= lo
384
Questions: answers and comments
Reversing the order of the two terms on the right-hand side gives
log10 n = at log10 e + log10 n0
= (a log10 e)t + log10 n0
Comparison with the general equation of a straight-line graph, y = mx + c,
shows that a graph of log10 n against t will be a straight line of gradient
a log10 e and intercept on the vertical axis of log10 n0.
Question 10.10
(a) ln 5 = 1.609
(b) The number whose natural logarithm is 5 is e5 = 148.4
Question 10.11
n = n0eat
Taking the log to base e of both sides of the equation gives
ln n = ln(n0e at )
= ln n0 + ln e at (from Equation 10.8)
= ln n0 + at (from Equation 10.7)
Reversing the order of the two terms on the right-hand side gives
ln n = at + ln n0
Comparison with the general equation of a straight-line graph, y = mx + c,
shows that a graph of ln n against t will be a straight line of gradient a and
intercept on the vertical axis of ln n0.
Question 10.12
If you had difficulty with any part of Question 10.12 you should refer
back to the section of the book indicated, and try the corresponding
Additional exercises on the accompanying website.
385
Maths for Science
(b) (i)
log10 4000 = log10 (4 ×1000)
= log10 4 + log10 1000 (from Equation 10.2)
2
= 0.6021 + log10 10
3
= 0.
6021 + 3 (from Equation 10.1)
(ii)
log1016 = log10 (4
2 )
= 2log10 4 ( from Equation 10.
4 )
= 2 × 0.
6021
= 1.204
0 to four significant figures
{See Section 10.3 Rules of logarithms. You may also find it helpful to look
back to Section 3.1.2 to remind yourself of the rules governing significant
figures and decimal places in calculations.}
(c) y = 3e−2x
Taking the log to base e of both sides of the equation gives
ln y = ln(3e −2x )
= ln 3 + ln e −2x (from Equation 10.8)
Reversing the order of the two terms on the right-hand side gives
ln y = −2x + ln 3
Comparison with the general equation of a straight-line graph, y = mx + c,
shows that a graph of ln y against x will be a straight line of gradient −2 and
intercept on the vertical axis of ln 3.
{See Section 10.3 Rules of logarithms and Section 10.5 Logarithms to
base e.}
Chapter 11
Question 11.1
(a) Of the 52 cards in the pack, 13 are hearts. So according to Equation 11.2,
the probability of a card drawn at random being a heart is
13 1
=
52 4
{This result also follows from noting that there are four suits, each with the
same number of cards, so one-quarter will be hearts.}
386
Questions: answers and comments
(b) Of the 52 cards in the pack, 26 are red (13 hearts and 13 diamonds). So
the probability of a card drawn at random being red is
26 1
=
52 2
{Or two of the four suits are red, so the probability is 2
4
= 12 .}
(c) Of the 52 cards in the pack, four are aces (one for each suit). So the
probability of a card drawn at random being an ace is
4 1
=
52 13
(d) Of the 52 cards in the pack, 12 are picture cards (three for each suit). So
the probability of a card drawn at random being a picture card is
12 3
=
52 13
Question 11.2
(a) For any one toss the probability of heads is always the same: 12 .
(b) For a single toss of the third coin, the probability of getting heads is 12
and that is unaffected by what has gone before. This is no different to tossing
the same coin three times in succession. Only foolish gamblers believe that
because heads have come up twice running the chances of tails coming up the
next time are thereby increased!
Question 11.3
(a) If two coins are tossed simultaneously, there are four possible outcomes,
all of which are equally likely:
outcome 1 H H
outcome 2 H T
outcome 3 T H
outcome 4 T T
The outcome of two tails can occur in only one way, so the probability of
getting two tails is 14 .
This result can also be found from the multiplication rule:
the probability that the first coin will show tails is 1
2
1
the probability that the second coin will show tails is 2
387
Maths for Science
1
(b) The probability of throwing a six with one die is 6
. So the probability of
getting a pair of sixes when throwing two dice is
1 1 1
× =
6 6 36
{An alternative method would be to list all outcomes, but it is quicker to use
the rule on this occasion.}
Question 11.4
Assuming the germination probabilities to be independent of one another:
(a) the probability of seeds of both A and B germinating is
1 1 1
× =
2 3 6
(b) the probability of the seeds of all three species germinating is
1 1 1 1
× × =
2 3 4 24
1
(c) the probability that a seed of A will not germinate is 2
2
the probability that a seed of B will not germinate is 3
3
the probability that a seed of C will not germinate is 4
Question 11.5
1
The probability of drawing any one particular card from the pack is 52 . This
is true for each of the three named cards. So the probability of drawing the
Jack of diamonds or the Queen of diamonds or the King of diamonds is
1 1 1 3
+ + =
52 52 52 52
Question 11.6
The situation is similar to the one described in Question 11.3. If two coins are
tossed simultaneously, there are four possible outcomes, all of which are
equally likely:
outcome 1 H H
outcome 2 H T
outcome 3 T H
outcome 4 T T
388
Questions: answers and comments
The outcome of a head and a tail can occur in two ways, so the probability of
getting a head and a tail is
2 1
=
4 2
This result can also be found from a combination of the multiplication and
addition rules. For the combination of one head and one tail:
1
the probability that the coin on the left will be a tail is 2
1
the probability that the coin on the right will be a head is 2
Question 11.7
The fraction of the atmosphere that is oxygen is
0.26 0.26
= = 0.21 to two significant figures
0.26 + 1 1.26
Expressed as a percentage, this fraction is 21%.
Question 11.8
For the ten measurements in Table 11.4,
mean = 1.122 mm to three decimal places
standard deviation σ = 0.123 mm to three decimal places
{If you obtained a value of 0.129 855 mm for the standard deviation, which
rounds to 0.130 mm to three decimal places, you have used the incorrect
function on your calculator or spreadsheet.}
Question 11.9
(a) For nine measurements, the median is the 5th measurement in the list (in
ascending or descending order). This is 7.8 cm.
(b) From Equation 11.3, or by using the statistical function on a calculator or
spreadsheet,
70.4 cm
mean = = 7.82 cm
9
389
Maths for Science
Question 11.10
The best estimate that can be made from these data of the mean number, μ, of
flowers per plant in the colony is the mean of the sample, x . In this case,
x = 7.3 flowers to one decimal place
{Note that it is normal practice to quote means and medians in this way, even
for quantities, such as numbers of flowers, which cannot really be fractional!}
The best estimate that can be made of the population standard deviation is the
sample standard deviation s. In this case,
s = 1.9 flowers to one decimal place
{If you obtained a value of 1.8 flowers for sample standard deviation, you
have used the incorrect function on your calculator or spreadsheet.}
Question 11.11
If you had difficulty with any part of Question 11.11 you should refer
back to the section of the book indicated, and try the corresponding
Additional exercises on the accompanying website.
390
Questions: answers and comments
chocolate and one plain chocolate are selected, the probability that both of
these will contain nuts is
2 3 6 1
× = =
9 8 72 12
{See Section 11.1.4 Combining probabilities.}
(d) The drink is made up of 8 + 7 + 2 + 1 = 18 parts, of which eight are
apple juice.
Thus the fraction of apple juice in the mixture is
8
= 44% to two significant figures
18
{See Section 11.1.5 Probability ratios and Section 1.2.6 Percentages.}
(e) (i) The mean number of children per household is
0 + 0 + 0 + 0 + 0 +1+1+ 2 + 3 + 5
= 1.2
10
(ii) Since there are 10 values (an even number), the median value is the mean
of the 5th and 6th values, i.e. the mean of 0 and 1. Thus the median number
of children per household is 0.5.
(iii) The mode is the most frequently occurring value, so the mode number of
children per household is 0.
{See Section 11.2.6 Different types of ‘average’.}
(f) For the ten measurements given
mean = 37.31° to two decimal places
standard deviation σ = 0.09° to two decimal places
{If you obtained a value of 0.099443° for standard deviation, which rounds to
0.10° to two decimal places, you have used the incorrect function on your
calculator or spreadsheet.}
{See Section 11.2.3 Mean and standard deviation for repeated measurements
(g) The best estimate that can be made from these data of the mean wing
length, μ, of all female meadow pipits in the population is the mean of the
391
Maths for Science
Chapter 12
Question 12.1
(a) The answer to this question depends on which significance level is used.
Employing the usual convention, i.e. rejecting the null hypothesis if P < 0.05,
the null hypothesis should be rejected on this occasion, since P < 0.01 means
that P must be less than 0.05. Therefore the alternative hypothesis should be
accepted. However, if it had been decided only to reject the null hypothesis if
P were less than 0.001, you would not be justified in categorically rejecting
the null hypothesis in this way.
(b) Employing the usual convention, the null hypothesis should be accepted,
since P > 0.05.
(c) This inequality is written in a way that is very unhelpful and ought to be
avoided. You are told that P > 0.01. But how much greater? If P > 0.05 then,
employing the usual convention, the null hypothesis must be accepted.
However, if P lies between 0.05 and 0.01 (i.e. 0.05 > P > 0.01) then,
employing the usual convention, the null hypothesis should be rejected and
the alternative hypothesis accepted. In the former situation, the result ought to
have been given as P > 0.05; in the latter it ought to have been given as
either P < 0.05 or 0.05 > P > 0.01.
Question 12.2
(a) Since the actual number of parasites per sheep is known, these data are at
the interval level.
(b) Since the sheep are classified into just two contrasting categories
(‘parasitised’ and ‘unparasitised’) these data are best treated as being at the
categorical level.
{Since there is an element of ranking here, you might have regarded these
data as being at the ordinal level. However, whether ‘unparasitised’ is ‘good’
or ‘bad’ does depend on whether you take the point of view of the sheep or
the parasites! ‘Parasitised’ and ‘unparasitised’ might correspond to the clear
cut categories ‘susceptible to parasites’ and ‘resistant to parasites’. In general,
ordinal level data are subdivided into more than two classes.}
(c) Since the degree of parasitisation is recorded, but not precisely how many
parasites there are on each sheep, these data are at the ordinal level.
Question 12.3
The total number of plants in the next generation was 636
(i.e. 185 + 305 + 146). If the ratio in a sample of 636 plants were
1 red-flowered : 2 pink-flowered : 1 white-flowered, then there would be
392
Questions: answers and comments
636
= 159 red-flowered plants
4
636
63
36
= 159 white-flowered plants
4
These are therefore the ‘expected’ numbers. The calculation of the test statistic
χ2 is shown in Table 12.16.
Question 12.4 n
The values of (RA)i, (RB)i, Di, Di2 and ∑ Di 2 are given in Table 12.17.
i=1
{Since there are two vertical distances of 7 cm, both are given the rank
3+ 4
= 3.5
2
The next vertical distance (9 cm) is given the rank 5.}
n
Substituting ∑ Di 2 = 21.5 (from Table 12.17) and n = 8 into Equation 12.2
gives i=1
6 × 21.5
rS = 1 −
8(82 − 1)
= 0.744
393
Maths for Science
Vertical distance/cm Rank (RA)i Mean water content/% Rank (RB)i Di = (RA)i − (RB)i Di2
0 1 76 1 0 0
4 2 83 3 −1 1
7 3.5 93 4 −0.5 0.25
9 5 80 2 3 9
7 3.5 102 6 −2.5 6.25
11 7 95 5 2 4
10 6 120 7 −1 1
13 8 130 8 0 0
n n
∑ Di = 0 ∑ Di 2 = 21.5
i=1 i =1
Reading across the row for 8 pairs of measurements in Table 12.7, it can be
seen that 0.05 > P > 0.01. Since P < 0.05, the null hypothesis must be
rejected at the 5% significance level and the alternative hypothesis accepted.
There is a statistically significant positive correlation between mean soil water
content and vertical distance from ridge crest.
{Although mean soil water content was significantly correlated with both
horizontal and vertical distance from the nearest ridge crest, the former
produced a value of rS that was both higher and more significant than the
latter (i.e. rS = 0.905, P < 0.01 compared to rS = 0.744, P < 0.05).}
Question 12.5
(a) These data are unmatched. There is no logical link between any particular
plant growing in one reserve and any particular plant growing in the other
reserve.
(b) These data are matched. For each sampling station along the stream, the
number of nymphs is known for two species of Stonefly.
Question 12.6
In this case x1 = 7.7, x2 = 7.2, s1 = 2.7, s2 = 2.1, n1 = 18 and n2 = 15.
Substituting for s1, s2, n1 and n2 into Equation 12.5 gives
= 5.
989
Substituting for SC2, n1 and n2 into Equation 12.4 gives
5.989 5.989
SED = + = 0.856
18 15
394
Questions: answers and comments
(18 − 1) + (15 − 1) = 31
The value of t (i.e. 0.584) is smaller than any of the critical values in the row
for 30 degrees of freedom (the nearest equivalent to 31) in Table 12.12. The
probability of obtaining a value of t as large as this by chance if the null
hypothesis were true is therefore greater than 0.1 (i.e. P > 0.1), probably
much greater. The difference in number of flowers per plant growing on either
side of this ridge is not statistically significant.
{Note: if you worked to a different number of decimal places in this question
you may have obtained a slightly different value for t. However, your
conclusion – that the difference in number of flowers per plant growing on
either side of this ridge is not statistically significant – should have been the
same.}
Question 12.7
If you had difficulty with any part of Question 12.7 you should refer
back to the section of the book indicated, and try the corresponding
Additional exercises.
(a) The null hypothesis is that the die is unbiased (and the alternative
hypothesis is that the die is biased in some way).
The ‘expected’ number of times each score would be obtained is
100
= 16.667 to 3 decimal places
6
The calculation of the test statistic χ2 is shown in Table 12.18.
395
Maths for Science
396
Questions: answers and comments
n
Substituting ∑ Di 2 = 102.5 (from Table 12.19) and n = 12 into Equation 12.2
gives i=1
n
6 ∑ Di
2
rS = 1 − i=1
n(n 2 − 1
)
6 × 102.
5
= 1−
12(122 − 1)
= 0.642
Reading across the row for 12 pairs of measurement in Table 12.7, it can be
seen that 0.05 > P > 0.01. Since P < 0.05, the null hypothesis must be
rejected at the 5% significance level and the alternative hypothesis accepted.
There is a statistically significant correlation between the length and the width
of the beans.
{See Section 12.4 The Spearman rank correlation coefficient.}
(c)(i) An appropriate null hypothesis would be that the mean test score is the
same before and after the students have studied Maths for Science. An
appropriate alternative hypothesis is that the mean test score is different before
and after the students have studied Maths for Science.
(ii) The data are at the interval level.
(iii) The two samples (before and after reading the book) are matched, so a
t-test for unmatched samples would not be appropriate. A different test should
be used.
{See Section 12.1 The principles of hypothesis testing, Section 12.2 Deciding
which test to use; levels of measurement and Section 12.5 The t-test for
unmatched samples.}
397
Maths for Science
Acknowledgements
Nick Adams, Jenny Barden, Tracy Bartlett, Greg Black, Jonathan Crowe,
Richard Jordan, Stephen Larkin, Lisa Mills, Tina Overton, David Vince and
Figure 1.3: Dr. Wolfgang Beyer, used under the Creative Commons
of the BIPM; Figure 4.6a: P.R. Deakin and Treak Cliff Cavern; Figure 4.6b:
Figure 12.1a: Skyscan Aerial Photography; Figures 12.1b and 12.2 inset: Mike
Every effort has been made to contact copyright holders. If any have been
398
Index
Index
A Celsius degree, 35, 38
density, 65, 84
cancellation, 12
dependent variable, 150
centi (prefix), 38
derivative, 177
CGS units, 85
derived function, 182
accuracy, 267
checking, 123
derived units, 67
al-Khwarizmi, 99
295
differentiation, 177
algebra, 58, 89
chord, 176
differentiation from first principles,
angles, 197
common denominator, 12
direct proportionality, 153
ångström (unit), 38
commutative, 8
antilog, 231
complex number, 4
E
approximately equals, 56
arc, 200
concentration, 83
elimination, 113
arccos, 210
arcsec, 200
arcsin, 210
arctan, 210
conversion factor, 74
equation of straight-line graph, 156
area, 49, 74
correlation, 299
equivalent fraction, 11
cosine, 207
estimated standard deviation of
atto (prefix), 37
average, 270
Avogadro’s number, 36
295
evaluate, 13
base number, 20
cube root, 25
exponential growth, 169
base units, 34
cystic fibrosis, 258
exponent, 20
expression, 90
BEDMAS, 26, 62
D extrapolation, 143
bimodal, 277
decimal notation, 4
factors, 109
281
degree Celsius, 35, 38
fractional exponent, 25
calculus, 173
degrees of freedom, 294
fractions, 4, 11, 18, 104
399
Maths for Science
kilowatt-hour (unit), 85
null hypothesis, 287
G number line, 3, 34
giga (prefix), 37
L numerator, 11
gradient, 147
Leibniz, Gottfried Wilhelm, 184
O
graphs, 137
levels of measurement, 290
observed numbers, 292
litre, 38, 76
ordinal level, 290
H logarithm, 229
origin (of graph), 143
hertz (unit), 68
log–linear graph, 243
pascal (unit), 68
histogram, 138
log–log graph, 238
percentage 17
hyperbola, 163
longitude, 199
peta (prefix), 37
hypotenuse, 205
lowest common denominator, 12
pH scale, 232
hypotheses, 286
photoelectric effect, 149, 158
Ma (million years), 38
pico (prefix), 37
I
magnitude, 220
playing cards, 255
improper fraction 13
mass, 35, 93
population mean, 280
258
mean, 270
powers, 20
inheritance, 259
Mendel’s peas, 262
precedence, 26
integer, 3
metre, 35
precise, 267
intercept, 156
micro (prefix), 37
precision, 44, 267, 276
interpolation, 142
milli (prefix), 37
probability, 254
210
probablilities, 257
Q
irrational number, 4
quadratic equations, 110
J nano (prefix), 37
400
Index
zero, 10
rational number, 4
statistically significant, 290
real number, 4
statistics, 266, 285
rearranging equations, 89
straight-line graph, 144
reciprocal, 21
Student’s t-test, 306
recurring decimal, 4
Student’s t-test (critical values),
refraction, 217
310
Richter scale, 43
subtend, 197
ridge-and-furrow, 285
sum, 8
roots, 25
rounding, 44, 52
T
rounding errors, 52
tangent (to a curved graph), 173
S tera (prefix), 37
sample, 280
term, 90
scalar, 220
test of difference, 287
second (time), 35
true mean, 280
SI units, 34, 67
similar, 206
units, 34, 50, 57, 64, 66, 123, 132
sine, 207
skewed, 276
variables, 21, 59
solutions, 59
solving equations, 59
W
Spearman rank correlation
watt (unit), 68
weight, 35, 93
word equation, 58
speed, 77
standard form, 31
401