Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Thermal Notes

Download as pdf or txt
Download as pdf or txt
You are on page 1of 55

Contents

1 Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Measurement of temperature . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Common temperature scales . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 The Kelvin temperature scale . . . . . . . . . . . . . . . . . . . . . . . 2
2 Thermal expansion of matter . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1 The linear expansion of solids . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Expansion of holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.3 The binomial theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.4 Cubical (or volume) expansion . . . . . . . . . . . . . . . . . . . . . . . 8
2.5 Thermal stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.6 The expansion of liquids . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.7 The expansion of gases . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.8 The relative thermal expansion of solids, liquids and gases . . . . . . . 10
2.9 An important relationship between α and β . . . . . . . . . . . . . . . 10
2.10 The effect of expansion on density . . . . . . . . . . . . . . . . . . . . . 11
2.11 The anomalous behaviour of water near 4 ◦C . . . . . . . . . . . . . . . 11
3 Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.2 Heat capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.3 Changes of phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.4 Heat units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.5 Distinction between heat and temperature . . . . . . . . . . . . . . . . 14
4 Calorimetry calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
5 Mechanisms of heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
5.2 Conduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
5.3 Convection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
5.4 Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
5.5 The spectral distribution of black body radiation . . . . . . . . . . . . 22
6 The gas laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
6.1 Revision of some basic terminology . . . . . . . . . . . . . . . . . . . . 24
6.2 Equation of state of an ideal gas . . . . . . . . . . . . . . . . . . . . . . 26
6.3 Dalton’s law of partial pressures . . . . . . . . . . . . . . . . . . . . . . 27
6.4 The pressure due to a vapour . . . . . . . . . . . . . . . . . . . . . . . 27
6.5 The pressure due to a mixture of gases and a vapour . . . . . . . . . . 28
7 Elementary kinetic theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
7.1 Basic assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
7.2 The pressure exerted by an ideal gas . . . . . . . . . . . . . . . . . . . 29

i
7.3 The root-mean-square speed . . . . . . . . . . . . . . . . . . . . . . . . 30
7.4 The distribution of molecular speeds . . . . . . . . . . . . . . . . . . . 32
8 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
8.2 The internal energy of an ideal monatomic gas . . . . . . . . . . . . . . 33
8.3 The first law of thermodynamics . . . . . . . . . . . . . . . . . . . . . 33
8.4 The work done by a gas . . . . . . . . . . . . . . . . . . . . . . . . . . 34
8.5 The molar specific heat capacities of monatomic gases . . . . . . . . . . 35

ii
List of Examples

1 Expansion of a bridge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2 Expansion of a steel scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3 Expansion of a steel ring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
4 Expansion of an aluminium ring . . . . . . . . . . . . . . . . . . . . . . . . . . 7
5 Surface expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
6 Heat generated by the human body during illness . . . . . . . . . . . . . . . . 13
7 Heat required to convert ice to steam . . . . . . . . . . . . . . . . . . . . . . . 14
8 Mixture of hot and cold water . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
9 Temperature of a steel ball . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
10 Final temperature of a mixture of water and ice . . . . . . . . . . . . . . . . . 16
11 Specific latent heat of fusion for ice . . . . . . . . . . . . . . . . . . . . . . . . 16
12 Conduction of heat through glass . . . . . . . . . . . . . . . . . . . . . . . . . 18
13 Temperature at the interface of two materials . . . . . . . . . . . . . . . . . . 18
14 Thermal conduction in a compound bar . . . . . . . . . . . . . . . . . . . . . . 19
15 Rate of heat radiated from a body . . . . . . . . . . . . . . . . . . . . . . . . . 23
16 Energy absorbed by a body . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
17 Energy radiated by a black body . . . . . . . . . . . . . . . . . . . . . . . . . 23
18 Cooling rate of a black body . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
19 The number of moles in a given mass of substance . . . . . . . . . . . . . . . . 25
20 Examples of unit conversions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
21 The mean-square, root-mean-square and mean speed of molecules . . . . . . . 30
22 The root-mean-square speed of molecules . . . . . . . . . . . . . . . . . . . . . 31
23 The increase in internal energy for boiling water . . . . . . . . . . . . . . . . . 36
24 Heating of Argon at constant volume, then at constant pressure . . . . . . . . 37

iii
iv
Thermal physics

1 Temperature
Suppose that you are not feeling well. You decide to take your temperature with a clinical
thermometer and it reads 39 ◦C. This is above 37 ◦C (which is regarded as normal) and you
therefore have a fever. The thermometer provides a quantitative measure of how hot or cold
things are and therefore a basis for comparison of your current temperature to its level when
you are well.

1.1 Measurement of temperature


The process of taking a temperature or reading a thermometer is a common experience.
Besides the obvious function of providing a basis for comparing the hotness or coldness of
different objects, however, do the numbers have any more fundamental meaning? To ask an
even more basic question, how do hot objects differ from cold objects? What is this quantity
we call temperature?
Most of us have an intuitive sens of what hot and cold mean, but our senses (apart from
being limited) can also be misleading. The pain that we feel when we touch something that is
very hot can be similar to, and hard to distinguish from, the pain that we feel when we touch
a very cold object.
What, then, are we doing in making a temperature measurement? The process is ultimately
one of comparison, and the relative terms hotter and colder actually have more meaning than
hot and cold themselves. A closer look at what happens when we use a thermometer will
clarify these issues.
The traditional clinical thermometer consists of a sealed glass tube partially filled with some
fluid, usually mercury. The inside diameter of the tube is very small, but it widens at the
bottom into a reservoir that contains most of the fluid. You usually place the thermometer
under your tongue and then wait a few minutes until the thermometer reaches the same
temperature as the inside of your mouth. Initially the mercury rises in the thin tube, and
when the level stops changing you read thermometer.
Why does the mercury rise? As we shall see, most materials expand when heated. The
fluid in the reservoir expands more than the glass does and the only place for the liquid to
go is up the tube. Here we use the physical property of the thermal expansion of the liquid
to give us an indication of temperature; by placing marks along the tube we can create a
temperature scale.
In principle, any physical property which changes with temperature could be used to
construct a temperature scale. Important examples of such properties (called thermometric
properties) are:

1. the volume of a liquid,

1
2. the resistance of a wire,
3. the voltage of a thermocouple,
4. the volume of a fixed mass of gas at constant pressure,
5. the pressure of a fixed mass of gas at constant volume,
6. the colour of a filament,
7. the length of a solid.
Implicit in this process of creating a temperature scale is the idea that if two objects are in
contact with each other long enough so their thermophysical properties do not change with
time, then the two objects have the same temperature. When we take a patient’s temperature,
we wait until the mercury stops rising before reading the thermometer. This idea is really
part of the definition of temperature because it determines when two (or more) objects have
the same temperature. When the thermometric properties are no longer changing, the objects
are said to be in thermodynamic equilibrium. Two or more objects that are in thermal
equilibrium with one another have the same temperature. This assumption is sometimes
referred to as the zeroth law of thermodynamics (with respect to the first law), because it
underlies the definition of temperature.
The zeroth law of thermodynamics states that if two thermodynamic systems are each in
thermal equilibrium with a third, then all three are in thermal equilibrium with each other.

1.2 Common temperature scales


A number of different temperature scales have been de-
vised, two popular choices being the Celsius and Fahren-
heit scales. The figure alongside illustrates these scales.
100 C

212 °F Historically, both scales were defined by assigning two
temperature points on the scale and dividing the dis-
tance between them into a number of equally-spaced in-
tervals. One point was chosen to be that at which ice
37 ◦C 98.6 °F
melts under one atmosphere of pressure (the ‘ice point’),
and the other was that at which water boils under one
0 ◦C 32 °F
atmosphere (the ‘steam point’). On the Celsius scale, an
ice point of 0 ◦C and a steam point of 100 ◦C were se-
lected. On the Fahrenheit scale these points were 32 °F
and 212 °F respectively. (See Problem A1)
(Note: The zero points on both the Fahrenheit and Celsius scales were arbitrarily selected
and have no fundamental significance. In fact, Celsius originally chose the ice point as 100 ◦C
and the boiling point as 0 ◦C!)

1.3 The Kelvin temperature scale


Although the zero points of both the Celsius and Fahrenheit scales were arbitrarily selected,
it is possible to define an absolute zero, which does have real fundamental significance. This
was not appreciated until 100 years or so after the original development of the Fahrenheit
and Celsius scales. The first suggestion that we might be able to define an absolute zero for
our temperature measurements came from studying changes in pressure and volume that take
place in gases when the temperature is changed.
If the volume of a gas is held constant while increasing the temperature, the pressure of
the gas will increase uniformly; it is thus a useful and important thermophysical property.

2
Measurements of the pressure of a gas at different temperatures can be made with a constant
volume gas thermometer, a diagram of which is shown alongside. It consists of a fused quartz
bulb connected by a capillary to a mercury manometer.
mercury A small quantity of gas is trapped in the bulb at low
pressure. The height of the mercury column may be ad-
justed so that the mercury level in the left branch of the
h U-tube is at a fixed reference mark to keep the confined
reference gas at constant volume. The pressure of the confined gas
level is the atmospheric pressure minus the difference in heights
of the mercury columns.
gas A plot of the gas pressure as a function of the Celsius
temperature results in a graph such as that shown in Fig-
ure 1. If the temperatures and pressures involved are not
flexible too high, an interesting feature emerges from plots of this
tube
nature. The graphs are all straight lines and, when extrap-
olated backwards to zero pressure, they all intersect the temperature axis at the same point.
This is true regardless of the nature of the gas used (oxygen, nitrogen, helium, etc.) or the
quantity of gas involved.
Pressure

b
b

b
b
b

b
b

b b
b
b
b
b

−273 ◦C −200 ◦C −100 ◦C 0 ◦C 100 ◦C 200 ◦C


0K 73 K 173 K 273 K 373 K 473 K
Temperature
Figure 1: Graph of pressure versus temperature for various ideal gases.

Careful experiments reveal that the point at which the graphs intersect the temperature
axis lies at −273.15 ◦C. Since a negative pressure has no meaning, this suggests that the
temperature cannot go lower than −273.15 ◦C. This temperature is referred to as absolute
zero. The Kelvin (named after Lord William Thomas Kelvin), or absolute temperature scale
has its zero at this point and uses the same size degree as the Celsius scale. Thus, the Kelvin
temperature T and the Celsius temperature t are related by
T = t + 273.15. (1)
The SI base unit for temperature is the kelvin (K).
According to Figure 1 the pressure of an ideal gas is directly proportional to the Kelvin
temperature. Thus for an ideal gas at constant volume
T2 P2
= , (2)
T1 P1

3
where the T1 and T2 are in Kelvin. By choosing a unique reproducible state, Equation (2)
can be used to define the Kelvin temperature. Such a state is the triple point of water. At
the triple point of water, solid water (ice), liquid water and water vapour can all coexist. The
triple point of water occurs at a temperature of 0.01 ◦C and a water-vapour pressure of 610 Pa.
The triple-point temperature is then defined to have the value Ttriple = 276.16 K.

P
T = 273.16 at constant volume. (3)
Ptriple

2 Thermal expansion of matter


It is well known that most materials expand when heated and contract when cooled. A change
in the external dimensions of a body implies corresponding changes in the average separation
between the atoms or molecules composing the body. We therefore seek the reasons for ther-
mal expansion in solids in the form of the intermolecular or inter-atomic potential energy
curve which determines the mean separation between the constituent particles. The continu-
ous curve in the graph in Figure 2 shows a typical interatomic potential energy (V (r)) of two
neighbouring atoms as a function of their separation (r). In the vicinity of the equilibrium
V (r)
mean
separation
r0 r

Figure 2: Interatomic potential energy as a function of their separation.

position r0 the potential curve is parabolic, the restoring force is proportional to the displace-
ment r − r0 from equilibrium and the vibrations are simple harmonic oscillations about r0 . If
the shape of the potential curve were to remain parabolic (as indicated by the dotted line)
for all r, then a rise in temperature would cause the atoms to oscillate with greater amplitude
about their equilibrium positions, but in view of the symmetry of the vibrations about r0
the average interatomic distance would remain r0 , and the body would neither expand nor
contract with temperature change. Thermal expansion in solids occurs because of departures
of the real potential energy curve from parabolic shape. Thus, when the temperature of the
solid rises, so the atomic vibrations increase in amplitude: the atoms make excursions into
the non-parabolic portions of the potential curve. In the vast majority of cases the departures
from parabolic shape are of the from shown in Figure 2, and the vibrating atoms spend more
time at the large than at the small separations. On a time average the mean interatomic
separation is larger than r0 the greater the temperature. The body as a whole expands.

4
2.1 The linear expansion of solids
The increase in any one dimension of a solid is called a linear expansion. Experiment shows
that the amount by which a material expands or contracts depends on two factors:

1. the magnitude of the temperature change ∆T ,

2. the initial linear dimension of interest (e.g. length, width, breadth).

For example, consider a metal rod of initial length ℓ0 at an initial temperature T0 . If this rod
is heated to a new temperature T , what will be the change in length ∆ℓ?
From points 1. and 2. above we know that

∆ℓ ∝ ℓ0
or ∆ℓ ∝ ℓ0 (T − T0 )
∆ℓ ∝ (T − T0 )

Introducing a proportionality constant α gives

∆ℓ = αℓ0 (T − T0 )
∆ℓ
α= , (4)
l0 (T − T0 )

where α is called the mean coefficient of linear expansion (units: ◦C−1 ).


The mean coefficient of linear expansion is defined as the fractional change in length
(∆ℓ/ℓ0 ) per degree Celsius change in temperature.
Equation (4) can be written in the form

∆ℓ = ℓ − ℓ0 = αℓ0 (T − T0 )
ℓ = ℓ0 + αℓ0 (T − T0 )
∴ ℓ = ℓ0 [1 + α(T − T0 )], (5)

which shows how ℓ varies with temperature. This result is valid for moderate temperature
changes (a few hundred degrees Celsius) only.

2.2 Expansion of holes


Consider a homogeneous body of any shape (including
cavities and holes), free to expand under an increase in
temperature. Let ℓ represent the distance between any
two points in the body.
When the temperature changes, ℓ changes by an
amount ∆ℓ. Experiment shows that, for a given body,
ℓ ∆ℓ/ℓ is the same for all pairs of points of the body, i.e.
the increase in length is strictly proportional to the orig-
inal length ℓ; so that the body changes size, but not
shape. The sizes of holes or cavities in the body expand
or contract as if the holes were filled with the material of
the body.

5
2.3 The binomial theorem

In many physics problems (especially in thermal expansion) we need to expand the quantity
(1 + x)n , where x is a number very much less than unity and n is either a positive or negative
number. Using the binomial theorem, this quantity can be written as

n(n − 1) 2 n(n − 1)(n − 2) 3


(1 + x)n = 1 + nx + x + x + ··· .
2! 3!

Since x ≪ 1, the terms in x2 and x3 are negligible and

(1 + x)n ≈ 1 + nx for x ≪ 1. (6)

For example:

1. (1 + 0.001)5 ≈ 1 + 5 × 0.001 = 1.005.


(Compare this value with the more precise answer of 1.005010.)

1
2. = (1 + 0.01)−1 ≈ 1 + (−1) × 0.01 = 1 − 0.01 = 0.99.
1 + 0.01
(Compare this value with the more precise answer of 0.990099.)

Example 1: Expansion of a bridge


By how many millimetres will a steel bridge of 100 m span increase in length when its tem-
perature rises from −10 ◦C to 40 ◦C? Take α (steel) = 1.1 × 10−5 ◦C−1 .

Solution:
The initial length and change in temperature are ℓ0 = 100 × 103 mm and ∆T = 40 −
(−10) = 50 ◦C. So the change in length is then

∆ℓ = αℓ0 ∆T = 1.1 × 10−5 × 105 × 50 = 55 mm.

Example 2: Expansion of a steel scale


A steel scale is correctly graduated at 0 ◦C. When measured by means of this scale on a day
when the temperature is 30 ◦C, a brass rod appears to be 98.90 cm long. What is the actual
length of the brass rod at 30 ◦C? (α (steel) = 1.1 × 10−5 ◦C−1 .)

6
Solution:
Since the scale is correctly graduated at 0 ◦C, each centimetre division of the scale is
ℓ0 = 1.000 cm in length. At 30 ◦C, each division has length ℓ, which from Equation (5) is

ℓ = 1.000 × 1 + 1.1 × 10−5 × 30 = 1.000 33 cm.
The brass rod measures 99.80 of these expanded centimetre divisions, the actual length is
99.80 × 1.000 33 = 99.83 cm.
This highlights the importance of a temperature-controlled workshop environment for pre-
cision engineering measurements where micrometers, verniers, etc., are used.

Example 3: Expansion of a steel ring


A steel tyre of internal diameter 99.90 cm at 15 ◦C is to be fitted on to a wheel whose external
diameter is 100.00 cm. To what minimum temperature must the tyre be heated before it can
be slipped on? (α (steel) = 1.1 × 10−5 ◦C−1 )
Solution:
Assume that the tyre can slip onto the wheel when its internal diameter is 100.00 cm.
For the diameter to reach this length, the change in temperature ∆T can be found using
Equation (5):
ℓ = ℓ0 (1 + α∆T )

100.00 = 99.90 × 1 + 1.1 × 10−5 × ∆T .
This gives ∆T = 91 ◦C, which means the tyre must be heated to 15 + 91 = 106 ◦C.

Example 4: Expansion of an aluminium ring


A man wishes to fit an aluminium ring on a rod of iron (diameter 1.0 cm), but it is 0.001 cm
too small in diameter. How much should its temperature be raised before it will just slip on?
(αAl = 2.5 × 10−5 ◦C−1 , αFe = 1.1 × 10−5 ◦C−1 .) Subsequently he wishes to remove it again,
but now he has to heat both metals together. Through how many degrees must this be done?
Solution:
The initial diameter of the aluminium ring is
ℓ0 = 1.000 − 0.001 = 0.999 cm.
Rearrange Equation (5) to make ∆T the subject (and let ∆ℓ = ℓ − ℓ0 ). This yields
1 1
∆T = ∆ℓ = × 0.001 = 40 ◦C.
ℓ0 α 0.999 × 2.5 × 10−5

Both materials expand when heated together; but for each Celsius degree increase in
temperature the aluminium expands more than the iron. We require the temperature
at which the two (unstressed) diameters become equal. Hence put
 
ℓ0,Al 1 + αAl ∆T = ℓ0,Fe 1 + αFe ∆T ,
which yields
ℓ0,Fe − ℓ0,Al 1.000 − 0.999
∆T = = = 71.6 ◦C.
ℓ0,Al αAl − ℓ0,Fe αFe 0.999 × 2.5 × 10−5 − 1.000 × 1.1 × 10−5

7
Example 5: Surface expansion
Consider a rectangular slab of material having initial length a0 and width b0 . The slab is
heated from an initial temperature T0 to a new temperature T . Calculate

(a) the increase in area of the slab,


(b) the fractional increase of the area.

Solution:
The initial area A0 is given by A0 = a0 b0 . The final area is (see Equation (5))

A = ab = a0 (1 + α∆T ) × b0 (1 + α∆T )
= a0 b0 (1 + α∆T )2
≈ A0 (1 + 2α∆T ),

where the binomial theorem was used. The increase in area ∆A is therefore

∆A = A − A0 ≈ 2A0 α∆T.

The fractional increase in area is


∆A
= 2α∆T.
A0
The factor 2α is sometimes called the coefficient of superficial expansion.

2.4 Cubical (or volume) expansion


Substances expand in all three directions when heated: length, breadth and height. They
experience a change of volume known as cubical (or volume) expansion. Experiments show
that if the temperature change ∆T is not too great (no greater than a few hundred ◦C), the
increase in volume ∆V is proportional to
1. ∆T ,
2. the original volume V0 .
By analogy with the definition of α, the coefficient of cubical (or volume) expansion is defined
∆V
β= . (7)
V0 ∆T

The mean coefficient of cubical expansion is defined as the fractional change in volume
(∆V /V0 ) per degree Celsius change in temperature.
The coefficient of cubical expansion β is defined for all three states of matter: solids, liquids
and gases. Note that the term linear expansion has no meaning for a liquid or a gas, since the
fluid always takes the shape of its container.
It follows from the definition in Equation (7) that
 
V = V0 1 + β T − T0 , (8)

where V is the volume at some temperature T .

8
2.5 Thermal stress
In some situations, the ends of a rod or slab of material are rigidly fixed, which prevents
thermal expansion or contraction. If the temperature should change, large compressive or
tensile stresses would develop; these are sometimes called thermal stresses. To calculate the
thermal stress we divide the problem into two steps.
1. The rod/slab expands (or contracts) by an amount ∆ℓ = αℓ0 ∆T .
2. A force is applied to compress (or expand) the material back to its original length.
The force per unit area (or stress) is related to the fractional change of length (the strain) by
F ∆ℓ
=Y ,
A ℓ0
where Y is Young’s modulus for the material. This gives rise to an equation for the thermal
stress as
F
= Y α∆T. (9)
A

For a three-dimensional object, the stress exerted on the object is the pressure exerted
over the entire surface of the object. The bulk modulus (B) of a material is defined as
∆P
B= . (10)
∆V /V0

2.6 The expansion of liquids


Liquids have no shape and therefore no fixed dimensions. Consequently, as has already been
stated, cubical expansion is the only concern when it comes to liquids. Moreover, they always
have to be held in some containing vessel. Owing to the expansion of the vessel, the apparent
expansion of the liquid is less than the real value.
Suppose that a glass vessel with a large bulb and thin neck is filled B
C
to a mark A at some temperature T0 . Let the volume up to this mark
A be V0 . Now, if the temperature of the system is raised to T , and A
assuming that the glass does not expand at all, the level of the liquid
should rise to B. The increase in volume from A to B is then the real
expansion of the liquid, given by
∆V = βr V0 ∆T
In practice, however, the bulb also expands and the liquid rises not to
B but to C. (Here the small second-order effect of the expansion of the narrow neck is ignored.)
The expansion from A to C is then given by βa V0 ∆T , where βa is the apparent coefficient of
cubical expansion. The expansion of the glass vessel is given by βves V0 ∆T . Hence
real expansion of liquid = apparent expansion of liquid + real expansion of vessel

βr V0 ∆T = βa V0 ∆T + βves V0 ∆T
βr = βa + βves ,
or
βa = βr − βves . (11)

9
2.7 The expansion of gases
The volume occupied by a gas is very sensitive to both temperature and pressure. Therefore,
if one wishes to measure the volume of a given mass of gas as a function of temperature (i.e.
its thermal expansion) the pressure must be kept constant in order to ensure that volume
changes are due to only changes in temperature. The zero coefficient of cubical expansion of
a gas at constant pressure β0 is defined as
V − V0
β0 = , (12)
V0 T
where V0 is the volume of a given mass of gas at 0 ◦C and V is the volume at a temperature T
of the same mass of gas at the same pressure. The result of many careful experiments indicates
that β0 is the same for all gases provided the pressure is low enough so that the gas may be
regarded as ideal. This value of β0 is
1 ◦ −1
β0 = C (= 3.66 × 10−3 ◦C−1 )
273.15
for a fixed mass of gas at constant pressure.

2.8 The relative thermal expansion of solids, liquids and gases


Typical orders of magnitude of the thermal expansion for each of the three states of matter
are given here:
β for a typical solid lies between 10−5 and 10−6 ◦C−1
β for a typical liquid lies between 10−4 and 10−5 ◦C−1
β0 for any gas at constant (low) pressure is about 10−3 ◦C−1
So the thermal expansion of a gas is about a thousand times greater than that of a typical
solid for a given temperature change.

2.9 An important relationship between α and β


Consider a cube of side ℓ0 at a temperature T0 , and of side ℓ at a different temperature T .
From Equation (5)

ℓ3 = ℓ30 (1 + α∆T )3 = ℓ30 (1 + 3α∆T + 3α2 ∆T 2 + α3 ∆T 3 ) (13)


≈ ℓ30 (1 + 3α∆T ),

since α∆T ≪ 1. But l3 = V , the final volume of the cube and l03 = V0 is the initial volume.
The volume expansion is therefore

V = V0 (1 + 3α∆T ).

Comparing this formula with Equation (8) gives

β = 3α. (14)

This result only applies to isotropic solids, that is, a solid for which α is the same in all
directions. Many solids (especially crystals) are anisotropic.

10
(a) (b) (c)
Figure 3: The contributions to the cubic expansion due to the (a) linear, (b) quadratic, and
(c) cubic terms in Equation (13).

2.10 The effect of expansion on density


When a substance is heated its volume increases but its mass is unchanged. Hence, in general,
density decreases with temperature. For a fixed mass of a substance, and from the definition
of density ρ, the mass m can be written as m = ρ0 V0 = ρT VT , where the subscripts 0 and T
refer to temperatures T0 and T respectively. Therefore
V0 V0
ρT = ρ0 = ρ0
VT V0 (1 + β∆T )
from Equation (8). Thus
ρ0
ρT = . (15)
(1 + β∆T )

2.11 The anomalous behaviour of water near 4 ◦C


Most substances expand when heated; however a few do not. Water is such a substance. If
water at 0 ◦C is heated, its volume decreases until the temperature reaches 4 ◦C. Above 4 ◦C,
water behaves normally and its volume increases as the temperature increases. Because a
given mass of water has a minimum volume at 4 ◦C, its density (mass per unit volume) is
greatest at this temperature, as the figure below shows.
The fact that water has its greatest density
at 4 ◦C, rather than at 0 ◦C, has important conse- 1000.0
quences for the way in which a lake freezes. When
the air temperature drops, the surface layer of 999.9


the water is chilled. When the water tempera-


Density kg m−3

ture reaches 4 ◦C, the surface layer becomes more


999.8
dense than the warmer water below it. This dense
water sinks, pushing up warmer water from below
999.7
which in turn is chilled at the surface. This pro-
cess continues until the temperature of the entire
lake reaches 4 ◦C. Thereafter, further cooling of 999.6
0 2 4 6 8 10
the surface water below 4 C makes the water less

Temperature ( C)

dense than the lower layers and thus remains on


the surface. Continued cooling of the surface layer to 0 ◦C leads to the formation of ice which
floats on the water. Below this sheet of ice the water temperature is above 0 ◦C. The ice
sheet acts as an insulator which inhibits the loss of heat energy from the body of the lake.

11
Furthermore, heat from the ground below helps prevent the entire lake from freezing. As a
result, lakes do not usually freeze solid, even during prolonged colds spells, so fish and other
aquatic life survive beneath the ice during winter.

3 Heat
3.1 Introduction
We know from experience that heat and temperature are related but they are not the same.
A burning match has a much higher temperature than a kettle full of boiling water but the
match cannot cause a severe burn whereas the hot water most certainly can. Suppose we pour
some of this boiling water into a cup. The water in the cup and in the kettle are at about
the same temperature. However, we could melt more ice with the water in the kettle than we
could with the water in the cup because the water in the kettle can give out more heat. It is
possible for a material to be at a high temperature and give out little heat, to be at a high
temperature and give out a large quantity of heat, to be at a low temperature and give out a
little heat, or to be at a low temperature and give out a large quantity of heat.
In matter the individual molecules possess various forms of energy, viz. translational kinetic
energy, rotational kinetic energy, vibrational kinetic energy and potential energy (the latter
originating from the forces which act between the atoms of a molecule). The sum of the various
forms of energy is called the internal energy of the substance. When two bodies at different
temperatures are placed in contact they eventually reach thermal equilibrium because the
internal energy from the hot object ‘flows’ into the cooler object. Energy in transit is called
heat.
Note that substances do not contain heat, they contain internal energy. The word ‘heat’ is
only used when referring to the energy in transit due to a difference in temperature between
parts of a system.

3.2 Heat capacity


Let Q be the amount of heat which is added to a body of mass m. Suppose the body undergoes
no change in phase and that its temperature rises by ∆T .
The heat capacity is defined by
Q
C= (16)
∆T
(SI units: J ◦C−1 ).
The specific heat capacity c of the body is defined by
Q
c= (17)
m∆T
(SI units: J kg−1 ◦C−1 ).

3.3 Changes of phase


It is important to realise that there are many situations in which the addition or removal of
heat from a system does not cause a temperature change but results only in a change of phase
(e.g. the temperature of melting ice remains at 0 ◦C; the temperature of boiling water is 100 ◦C
— in both cases we assume standard atmospheric pressure).

12
Suppose heat Q added to a mass m changes its phase without causing a temperature
change. Then we define the specific heat of transformation L as
Q
L= (18)
m
(SI units: J kg−1 ).
The equations

Q = mc∆T (19)
Q = mL, (20)

follow directly from the defining Equations (17) and (18), respectively. They are important
equations, especially for doing calorimetry calculations.
The following points should be especially noted:
1. The transformation referred to in Equation (18) could be any one of the following pro-
cesses: melting (or fusion), freezing, vaporization, condensation, sublimation.

2. Specific heats of transformation are sometimes called ‘specific latent heats’ or simply
‘latent heats’.

3. During melting, energy is required to melt the solid and during freezing, energy is released
from the liquid;

Q = mLf

(f denotes fusion).

4. For the liquid-gas phase change, energy is required during vaporization and energy is
released during condensation;

Q = mLv

(v denotes vaporization).

3.4 Heat units


Since heat is energy in transit, the SI unit of heat is the joule (J). However, dietitians and
nutritionists use the Calorie (spelt with a capital C) to specify the energy content of foods.
The conversion is: 1 Calorie = 4.1868 J.

Example 6: Heat generated by the human body during illness


During a bout of ‘flu’ a 60 kg student ran a temperature of 39 ◦C (instead of the normal 37 ◦C).
Assuming that the human body is mostly water (for which cwater = 4180 J kg−1 ◦C−1 ), how
much heat is required to raise the body temperature to 39 ◦C?
Solution:
The change in temperature ∆T is 39 − 37 = 2 ◦C. From Equation (19):

Q = mc∆T = 60 × 4180 × 2 = 501.6 kJ.

This corresponds to about 120 (food) Calories.

13
Example 7: Heat required to convert ice to steam
Calculate the heat required to convert 10 g of ice at −20 ◦C into 10 g of steam at 130 ◦C at a
pressure of 1 atmosphere. Assume the following quantities:

For ice ci = 2091 J kg−1 ◦C−1


For water cw = 4.18 × 103 J kg−1 ◦C−1
For steam cs = 2005 J kg−1 ◦C−1
For water Lf = 3.34 × 105 J kg−1
Lv = 2.25 × 106 J kg−1

Solution:
There are five processes involved to convert 10 g of ice at −20 ◦C into 10 g of steam at
130 ◦C: add heat to the 10 g of ice to raise its temperature from −20 ◦C to 0 ◦C, add heat
to melt the ice at 0 ◦C, add heat to raise the temperature of the water from 0 ◦C to 100 ◦C,
add heat to vaporize the water at 100 ◦C and finally add heat to raise the temperature of
the steam from 100 ◦C to 130 ◦C.

To heat 10 g of ice from −20 ◦C to ice at 0 ◦C requires

Q1 = mcice ∆T = 0.010 × 2091 × 20 = 418.2 J.

To melt 10 g of ice at 0 ◦C to water at 0 ◦C requires

Q2 = mLf = 0.010 × 3.34 × 105 = 3340.0 J.

To heat 10 g of water from 0 ◦C to water at 1000 ◦C requires

Q3 = mcwater ∆T = 0.010 × 4180 × 100 = 4180.0 J.

To vaporize 10 g of water at 100 ◦C to steam at 100 ◦C requires

Q4 = mLv = 0.010 × 2.25 × 106 = 22 500.0 J.

To heat 10 g of steam from 100 ◦C to steam at 130 ◦C requires

Q5 = mcsteam ∆T = 0.010 × 2005 × 30 = 601.5 J.

The total heat needed then is

Q = Q1 + Q2 + Q3 + Q4 + Q5 = 418.2 + 3340.0 + 4180 + 22500.0 + 601.5 = 31 040 J

Note: 72 % of the heat required is used to convert the water into steam.

3.5 Distinction between heat and temperature


It is important that you take some care to distinguish between the terms heat and temperature.
Clearly the two concepts are related but they are not the same. Unfortunately, everyday
usage often confuses these terms for which physicists have adopted a more precise meaning.
However, we are now in a position to complete our definition of temperature and to make a
clear distinction between these two concepts.

14
When two objects at different temperatures are placed in contact, heat will flow sponta-
neously from the object with the higher temperature to the object with the lower temperature.
Heat added increases the temperature; heat removed decreases the temperature. The amount
of heat that must be added or removed to produce a change in temperature depends on the
mass of the material involved and its specific heat capacity.
Temperature is a quantity that tells us in which direction heat will flow. If the two objects
are at the same temperature, no heat will flow. If they are at different temperatures, the
direction of heat flow is from the higher to the lower temperature. The quantity of heat that
is transferred depends on the temperature difference between the two materials as well as upon
their masses and specific heat capacities.
Heat
Heat is the net energy which is transferred spontaneously from regions of high temperature
to regions of low temperature.

Temperature
Temperature is the quantity that indicates whether or not, and in which direction, heat will
flow. Objects at the same temperature are in thermal equilibrium, and no heat will flow
from one to the other.

4 Calorimetry calculations
4.1 Introduction
Calorimetry means ‘measuring heat’ and calorimetry calculations therefore involve calculations
with heat. Suppose we bring into contact substances which are at different temperatures (e.g.
pouring cold milk into your hot tea). One might be interested in the final temperature after
the substances have been thoroughly mixed. The basic principle, which is very simple, is based
on the law of conservation of energy. The heat ‘lost’ by the hot substance(s) equals the heat
‘gained’ by the cold substance(s). In equation form:

Heat ‘lost’ = Heat ‘gained’ (21)

Very often we assume that the heat transferred to the surroundings is negligible and in practice
we would attempt to minimise such transfers.

Example 8: Mixture of hot and cold water


Suppose 200 g of water at 40 ◦C is mixed with 120 g of water at 16 ◦C. Calculate the resulting
temperature of the mixture.
Solution:
Use Equations (21) and (19) (Q = mc∆T )

Heat ‘lost’ by the hot water = Heat ‘gained’ by the cold water
0.200 × cwater × (40 − T ) = 0.120 × cwater × (T − 16)
40 − T = 0.6 × (T − 16)
∴ T = 31 ◦C

15
Example 9: Temperature of a steel ball
A 100 g stainless-steel ball is heated for some time in a Bunsen flame. It is then rapidly
plunged into 250 g of water, and causes a rise of temperature from 12 ◦C to 30 ◦C. Find the
temperature of the heated stainless-steel ball. Take csteel = 460 J kg−1 ◦C−1 .
Solution:
Let T be the temperature of the heated stainless-steel ball. Then

Heat ‘lost’ by the steel ball = Heat ‘gained’ by the water


mball × csteel × ∆T = mwater × cwater × ∆T
0.100 × 460 × (T − 30) = 0.250 × 4180 × (30 − 12).

Solving the above for T gives a temperature for the heated steel ball of 439 ◦C.

Example 10: Final temperature of a mixture of water and ice


Suppose that 100 g of water at 60 ◦C and 10 g of ice at −10 ◦C are mixed together. What is the
final temperature T of the mixture? Take cice = 2090 J kg−1 ◦C−1 and Lf = 3.34 × 105 J kg−1 .
Solution:
While the water loses heat, so that its temperature drops from 60 ◦C to some final tem-
perature T , the ice gains heat to (i) raise its temperature from −10 ◦C to 0 ◦C, (ii) melt at
0 ◦C, and (iii) to raise its temperature from 0 ◦C to the final temperature T of the mixture.
The heat required for the ice to undergo those three processes must all come from the heat
lost by the hot water, so that:

Heat ‘lost’ by water = Heat ‘gained’ by ice (to warm and melt it) and to warm it to T.

Hence

mw × cw × (60 − T ) = mi × ci × (0 − (−10)) + mi × Lf + mi × cw × (T − 0)
0.100 × 4180 × (60 − T ) = 0.010 × 2090 × 10 + 0.010 × 3.34 × 105 + 0.010 × 4180 × T,

which gives T = 46.8 ◦C.

Example 11: Specific latent heat of fusion for ice


A copper calorimeter having a mass of 24.40 g when empty is two-thirds filled with water at
21 ◦C and the total mass is found to be 130.56 g. Small chips of ice at 0 ◦C are added and
the mixture is stirred. At one stage, when all the ice had just melted, the lowest temperature
reached was found to be 10.5 ◦C and the total mass of the calorimeter and contents was 143.16 g.
From these data calculate the specific heat of fusion for ice, given that cCu = 418 J kg−1 ◦C−1 .
Solution:
The calorimeter and the warm water it contained both lose energy. This energy loss must
equal the energy gained by the ice to melt it and to raise the temperature of the ice-water
after melting. By Equation (21) we write:

mCu cCu ∆TCu + mw cw ∆Tw = mice Lf + mice cw ∆Tice,w ,

16
where the mass of the mass of the water in the calorimeter before the ice is added and the
mass of the added ice are:

mw = (130.56 − 24.4) × 10−3 = 106.16 × 10−3 kg,


mice = (143.16 − 130.56) × 10−3 = 12.6 × 10−3 kg.

Therefore

24.4 × 10−3 × 418 × (21 − 10.5) + 106.16 × 10−3 × 4180 × (21 − 10.5) =
12.6 × 10−3 × Lf + 12.6 × 10−3 × 4180 × (10.5 − 0),

which gives Lf = 3.344 × 105 J kg−1 .

5 Mechanisms of heat transfer


5.1 Introduction
Heat is transmitted from one place to another by three fundamentally different processes:
conduction, convection and radiation. We consider each process in turn.
Conduction is the process whereby energy is transferred directly through a body (or be-
tween two bodies in contact) due to interatomic or inter-molecular collisions. Bulk motion of
the material plays no role in the transfer.
Conduction is the process whereby energy is transferred from one place to another by the
bulk motion of a fluid.
Radiation is heat transfer by electromagnetic radiation (such as sunlight) with no need for
matter to be present in the space between the bodies.

5.2 Conduction
Suppose you hold one end of a metal rod (length ∆x; cross-sectional area A) in a flame. Heat
reaches your hand by conduction through the rod. Initially the temperature at points along
the rod increases with time. But after a while these temperatures stop changing — this is
called ‘the steady state’. Experiment shows that, once the steady state has been reached, the
rate of heat flow along the rod Q/t (here Q is the heat conducted along the rod in time t) is
proportional to
Q
1. the cross-sectional area A, i.e. ∝ A, and
t
∆T
2. the temperature gradient across the ends of the rod. Here, ∆T is the difference in
∆x
Q ∆T
temperature between the ends of the rod. So ∝ .
t ∆x
Combining these results gives
Q ∆T
= λA , (22)
t ∆x
where the constant of proportionality λ is called the thermal conductivity of the material.
In Equation (22) the temperature gradient ∆T∆x
is taken as a positive quantity.

17
Q
The heat power is measured in J s−1 which is equivalent to a Watt (W). Therefore, the
t
SI unit of λ is
[Q/t] W
[λ] = = 2 ◦  = W m−1 ◦C−1
[A][∆T /∆x] m Cm −1

Values of λ for some common materials are shown in Table 1. Notice that metals (good
conductors) have relatively large conductivities.

Material λ/W m−1 ◦C−1


silver 420
copper 390
aluminium 210
ice (0 ◦C) 2.2
glass 0.8
wood (oak) 0.15
air 0.0256

Table 1: The thermal conductivity of some common materials.

Example 12: Conduction of heat through glass


A glass window pane has an area of 3.0 m2 and a thickness of 6.0 mm. If the temperature
difference between its faces is 25 ◦C, how much heat is conducted through the glass every
hour?
Solution:
The rate of heat conduction is
Q ∆T
= λglass A
t ∆x
25
= 0.8 × 3.0 ×
6.0 × 10−3
4
= 10 W.
Therefore, in one hour the total heat conducted is
Q = 104 × 3600 = 3.6 × 107 J.

Example 13: Temperature at the interface of two materials


Two identical rods are joined together as shown. One is made of copper and the other from
aluminium. The end of the copper rod is held at −200 ◦C and the end of the aluminium is at
100 ◦C.
−200 ◦C Cu Al 100 ◦C
T
(a) Calculate the temperature of the copper-aluminium junction in the steady state, assum-
ing no heat loss through the sides (which are said to be lagged).

(b) Show that the ‘effective’ thermal conductivity of the two rods is 273 W m−1 ◦C−1 .

18
Solution:
(a) Let T be the steady-state temperature of the junction. By conservation of energy
heat flux through Cu = heat flux through Al
   
∆T ∆T
∴ λCu ACu = λAl AAl .
∆x Cu ∆x Al
But ACu = AAl and ∆xCu = ∆xAl (rods are identical). Hence
λCu ∆TCu = λAl ∆TAl
390[T − (−200)] = 210[100 − T ],
which gives T = −95 ◦C.
∆T
Note that the temperature gradient ∆x
must always be positive so
∆T = higher temperature − lower temperature.
(b) Assume the two joined rods are now one single material of thermal conductivity λ. In
this case the temperature T at the middle of the rod can be obtained by symmetry.
Thus T = 50 ◦C and ∆T = 150 ◦C. Then, as for part (a),
λCu ∆TCu = λAl ∆TAl = λ∆T,
which yields λ = 273 W m−1 ◦C−1 .

Example 14: Thermal conduction in a compound bar

A compound bar (shown alongside in cross section) 2.0 m long


is constructed of a solid aluminium core 1.0 cm in diameter sur-
rounded by a copper casing whose outer diameter is 2.0 cm. The
outer surface of the bar is thermally insulated. One end is main- Cu Al
tained at 100 ◦C and the other at 0 ◦C.

(a) Find the total heat flux in the bar.


(b) What fraction is carried in each material?
Solution:
(a) By conservation of energy, the total heat flux must be equal to the heat flux carried
by the aluminium (Al) core plus the heat flux carried by the copper (Cu) casing, i.e.
     
Q ∆T ∆T
= λAl AAl + λCu ACu ,
t total ∆x Al ∆x Cu
where
πd2 π(10−2 )2 π
AAl = = = × 10−4 m2
4 4 4
and
π(d2o − d2i ) π × (4.0 − 1.0) × (10−4 ) 3π
ACu = = = × 10−4 m2 .
4 4 4
The total heat flux is then
     
Q π 100 − 0 3π 100 − 0
= 210 × × 10 × −4
+ 390 × ×
t total 4 2 4 2
= 0.824 + 4.585 = 5.419 W.

19
(b) The fractions carried by each metal are:

0.824
% carried by Al = × 100 = 15.2 %
5.419
and
4.595
% carried by Cu = × 100 = 84.8 %.
5.419

5.3 Convection
Convection is the transfer of heat by mass motion of a fluid from one region of space to another.
Familiar examples include heaters in the home, the cooling systems of car engines and the
heating of the body by the flow of blood. If the fluid is circulated by a blower or pump, the
process is called forced convection. If the flow is caused by differences in density due to
thermal effects (such as hot air rising), the process is called natural convection.
Try to explain the following in terms of convective heat flow:

• How do vultures and some birds of prey manage to gain very high altitudes?

• How do woollen clothes keep us warm?

• How does one explain land and sea breezes?

Convective heat transfer is a very complex and mathematically difficult process and there is
no simple equation as there is for conduction. Here are a few experimental facts:

• The rate of heat transfer due to convection is directly proportional to the surface area.
This the reason for large surface areas of radiators and cooling fins.

• The viscosity of fluids slows natural convection near a stationary surface, giving a surface
film that on a vertical surface typically has about the same insulating value as 1.3 cm of
plywood. Forced convection decreases the thickness of this film, increasing the rate of
heat transfer. This is the reason for the ‘wind-chill factor’; you get cold faster in a cold
wind than in still air with the same temperature.

• The rate of heat transfer due to convection is found to be approximately proportional


to the 5/4 power of the temperature difference between the surface and the main body
of fluid.

5.4 Radiation
The transfer of energy by conduction and convection involves the participation of material
media. However, the transfer of energy by the process of radiation need not involve the
intervention of material media. An example is the radiation energy from the sun to the earth
where by far the greatest part of the space between these two bodies is a very good vacuum.
This radiant energy consists of electromagnetic waves travelling at the speed of light.
It can be shown experimentally that all bodies above absolute zero (0 K) radiate energy
and that the rate at which energy is radiated depends on the temperature of the body and
the nature of its surface. Also, when radiant energy falls on the surface of a body some of
the energy is absorbed and the remainder is either reflected or transmitted. A black body

20
is defined as one which would absorb all of the radiant energy incident upon it. Although
there is no real body satisfying this condition, a body whose surface is coated with soot is
very nearly an ideal black body.
The absorptivity α of a surface for radiant energy is defined as the fraction of the total
incident radiation absorbed by the surface. It follows that the absorptivity of a black body
is 1. For all other bodies α < 1.
It is known experimentally that bodies which are good absorbers of radiation are also
good emitters of radiation and that poor absorbers are poor emitters. It has been found
that the rate at which energy is radiated (i.e. the power) from a surface A at an absolute
temperature T is:
Q
= σεAT 4 , (23)
t
where σ is a constant called the Stefan-Boltzmann constant and is equal to
5.670 × 10−8 W m−2 K−4 , ε is a dimensionless quantity known as the emissivity of the body.
The emissivity is defined as the rate at which energy is emitted from the body to that emitted
by an identical black body at the same temperature. Clearly ε = 1 for a black body. For all
other bodies ε < 1. Furthermore, it can be shown that the absorptivity α and the emissivity ε
of a body are equal at the same temperature — a fact known as Kirchhoff’s law. Equation (23)
is known as the Stefan-Boltzmann radiation law.
Any object not only emits energy by radiation, but it also absorbs energy radiated by
other bodies. If an object of emissivity ε and area A is at a temperature T1 , it radiates at
a rate of σεAT14 , as given by Equation (23). If the object is surrounded by an environment
at a temperature T2 , the rate the surroundings radiate energy is proportional to T24 and thus
the rate that energy is absorbed by the object is proportional to T24 . The net rate of radiant
energy transfer (i.e. the ‘heat flow rate’) from the object is given by

Q
= σεA T14 − T24 .

(24)
t
Notice in this equation that the rate of heat absorption by an object was taken to be σεAT24 ;
that is, the proportionality constant is the same for both emission and absorption. This
must be true to correspond with the experimental fact that equilibrium between object and
surroundings is reached when they come close to the same temperature; that is, the net rate of
radiant energy transfer Q/t in Equation (24) must equal zero when T1 = T2 , so the coefficients
of each term must be the same.
Because both the objects and the surroundings radiate energy, there is a net transfer of
energy from one to the other unless everything is at the same temperature. From Equation (24)
it is clear that if T1 > T2 , the net flow of heat is from the body to the surroundings, so the
body cools; but if T1 < T2 , the net heat flow is from the surroundings into the body and
its temperature rises. If different parts of the surroundings are at different temperatures,
Equation (24) becomes more complicated.
Heating of an object by radiation from the sun cannot be calculated using Equation (24)
since T2 in this equation refers to the temperature of the environment surrounding the object,
whereas the sun is essentially a point source. It must, therefore, be treated as an additional
source of energy. Heating by the sun is calculated using the fact that 1350 J of energy strikes
the atmosphere of the earth from the sun per second per square metre of area at right angles
to the sun’s rays. This number, 1350 W m−2 , is called the solar constant which we denote
by the symbol I. The atmosphere may absorb as much as 70 % of this energy before it reaches

21
the ground, depending on the cloud cover. On a reasonably clear day, perhaps 1000 W m−2
reaches the earth’s surface.

su
n ’s
ra
ys

effective area
θ A cos θ
θ
Area A

An object of emissivity (absorptivity) ε with an area A facing the sun absorbs heat at the rate
Q
= IεA cos θ, (25)
t
where θ is the angle between the sun’s rays and a line perpendicular to the area A (as shown).
That is, A cos θ is the effective area at right angles to the sun’s rays.

5.5 The spectral distribution of black body radiation


The figure alongside shows the emitted power per
unit wavelength interval per square metre of black
body radiation as a function of wavelength λ for var-
10 1650 K
ious temperatures. Note that for a given tempera-
9 ture T1 there is a wavelength λ1 at which the inten-
sity is a maximum. At a higher temperature T2 the
8 power density is greater at all wavelengths and the
particular wavelength λ2 at which the intensity is a
7 maximum is shifted towards shorter wavelengths.
E in arbitrary units

Wien’s displacement law states that


6
1450 K
λ1 T 1 = λ2 T 2 or λmax T = constant (26)
5

4 It may be shown experimentally that λmax T =


1250 K
2.897 × 10−3 K m.
3 This suggests a convenient method for determin-
ing the temperature of a black body (or one which
2 approximates a black body). One examines the ra-
diation from the body with a spectrometer and the
1
wavelength λmax corresponding to the greatest in-
tensity is determined. A thermometer which uses
0 1 2 3 4 5 6 radiation as the thermometric property is called a
λ in µm pyrometer. This is how we can accurately measure
from the earth the temperature of distant celestial
objects such as the sun.
Radiation emitted by a body as a result of its temperature is called thermal radiation. All
bodies emit and absorb radiation from their surroundings. A black body (ε = 1) above abso-
lute zero emits a continuous spectrum of radiation. The energy emitted per unit wavelength

22
interval is shown in the figure above. At ‘ordinary’ temperatures most bodies are visible to
us, not by their emitted light, but by the light which they reflect. If no light shines on them
we cannot see them. At high temperatures, however, bodies become self-luminous. Even
at temperatures as high as several thousand degrees Kelvin, most of the emitted radiation
is invisible to humans as it falls in the infra-red portion of the spectrum. Therefore we say
self-luminous bodies are ‘hot’.

Example 15: Rate of heat radiated from a body


A person stands unclothed in a room whose walls are at a temperature of 15 ◦C. Calculate the
rate at which heat is radiated by the body assuming a skin temperature of 34 ◦C and ε = 0.70.
Take the surface area of the body to be 1.5 m2 .
Solution:
From Equation (24):

Q
= σεA TB4 − TW4

t
= 5.67 × 10−8 × 0.70 × 1.5 × 3074 − 2884


= 119 W.

Note: the absolute temperatures TB and TW are in Kelvin.

Example 16: Energy absorbed by a body


A girl sunbathes on a beach on a clear day. Suppose she lies flat on the sand and that the
sun’s rays make an angle of 30° with the horizontal. Assume that ε = 0.60, the area of her
body exposed to the sun is 0.80 m2 and that I = 1000 W m−2 . How much radiant energy does
her body absorb after half an hour of sunbathing?
Solution:
From Equation (25), the rate of energy absorption is

Q
= IεA cos θ
t
= 1000 × 0.60 × 0.80 × cos 60° (since θ is perpendicular to A)
= 240 W

Hence after 30 minutes

Q = 240 × 30 × 60 = 432 kJ.

Example 17: Energy radiated by a black body


How many days does it take a perfect black-body cube (1.0 cm on a side at 30 ◦C) to radiate
the amount of energy that a 100 W light bulb uses in one hour?
Solution:
The energy Q radiated by the bulb after 1 hour is

Q = P t = 100 × 60 × 60 = 3.6 × 105 J.

23
From Equation (23), the rate of energy radiated is

Q
= σεAT 4
t
3.6 × 105
∴ = 5.67 × 10−8 × 1 × 6 × 10−4 × 3034 ,
t

where the surface area A of the cube is 6 × 10−4 m2 . Solving the above for t gives an
equivalent of 14.5 days.

Example 18: Cooling rate of a black body


A blackened copper sphere 15 mm in diameter is inside an evacuated enclosure with blackened
walls maintained at 27 ◦C. Find the rate of cooling of the sphere when its temperature is
127 ◦C. For copper, take ρ = 8840 kg m−3 and c = 385 J kg−1 ◦C−1 .
Solution:
Since the density is given by mass/volume, the mass m of the sphere is
 3
4 15 × 10−3
m= π × 8840 = 1.56 × 10−2 kg.
3 2

The heat radiated by the sphere is given by Equation (24):


 2
Q 15 × 10−3
× 4004 − 3004

= 5.67 × 10−8 × 1 × 4π ×
t 2
= 0.701 W. (27)

∆T
Since Q = mc∆T , the rate of cooling t
can be found by taking Equation (19) and
dividing both sides by t:

Q ∆T
= mc
t t
∆T Q/t
∴ =
t mc
0.701
=
1.56 × 10−2 × 385
= 0.12 ◦C s−1 .

6 The gas laws


6.1 Revision of some basic terminology
Atomic and molecular mass
The unified atomic mass unit (symbol: u) is defined as one twelfth of the mass of an
unbound neutral atom of carbon-12 in its nuclear and electronic ground state (at rest).

24
The atomic mass of an element is the mass of an atom of that element on a scale on which
the mass of an atom of 126C = 12 u. To convert the atomic mass of an element to kilograms
we use the result

1 u = 1.660 538 921(73) × 10−27 kg (28)

Hence, the mass of 126C is 12 × 1.66 × 10−27 = 1.99 × 10−26 kg.


The molecular mass of a molecule is the sum of the atomic masses of its constituent atoms.
For instance, the molecular mass of the SO3 molecule is 32 + 3 × 16 = 80.

The mole and Avogadro’s number


One mole of any substance contains as many particles (or atoms) as there are atoms in 12 g of
the isotope 12C. Experiment shows that 12 g of 12C contains 6.022 × 1023 atoms. This number
is called Avogadro’s number NA .
One mole on any substance has a mass in grams which is equal to its relative atomic or
molecular mass (e.g. 1 mole of SO3 has a mass of 80 g).

Example 19: The number of moles in a given mass of substance


Calculate the number of moles in 100 g of SO3 .
Solution:
80 g of SO3 contains 1 mol (one mole) of SO3 .
1
∴ 1 g contains 80
mol.
100
∴ 100 g contains 80
= 54 mol.

Example 19 leads to the result for the number of moles

mass (g) m
n= = . (29)
molar mass (g) M

Standard temperature and pressure (STP)


• Standard temperature = 273.15 K.

• Standard pressure = 101 325 Pa = 1 Atm = 760 mmHg.

• Experiment shows that one mole of any gas at STP occupies 22.4 litres.

Absolute and gauge pressure


The pressure at a given depth h below the surface of a fluid of density ρ is given by

P = P0 + hρg, (30)

where P0 is the atmospheric pressure at the surface of the fluid.


The pressure P in Equation (30) is the total or absolute pressure at the depth h. The
pressure (hρg) due to the liquid alone is the gauge pressure.

25
6.2 Equation of state of an ideal gas
Experiment shows that if the temperatures are not too low and the pressures are not too high,
all real gases show the same simple behaviour. This suggests the concept of an ideal gas —
one that will have this simple behaviour under all conditions.
Methods exist for measuring the pressure P , volume V , temperature T and the mass m of
a gas. Experiments carried out on a gas at low density lead to the following results:
1. For a fixed mass of gas at constant temperature

P1 V1 = P2 V2
or P V = constant (Boyle’s Law).

2. For a fixed mass of gas at constant pressure


V1 V2
=
T1 T2
V
or = constant (Charles’ Law),
T
where T is the absolute temperature.

3. For a fixed mass of gas at constant volume


P1 P2
=
T1 T2
P
or = constant (Gay-Lussac’s Law), (31)
T
where T is the absolute temperature.
The three laws listed above can be combined to yield the result that for a fixed mass of an
ideal gas
P1 V1 P2 V2 PV
= or = constant, (32)
T1 T2 T
where T is the absolute temperature.
Since the mass of the gas m is fixed (and constant), Equation (32) can be written in the
form
PV
= mr, (33)
T
where r is a constant for a particular gas.
We find from experiment that for different gases r = R/M where R is a constant which is
the same for all gases and M is the relative atomic or molecular mass. This gives
PV R
=m . (34)
T M
From Equation (29) (n = m/M ) Equation (34) leads to
PV
= nR or P V = nRT, (35)
T

26
where T is the absolute temperature. The numerical value of the universal gas constant
R = 8.314 J mol−1 K−1 .
The behaviour of real gases conforms closely to Equation (35) except at high densities.
Equation (35) gives the relationship between the variables n, P , T and V and it is called the
equation of state of an ideal gas (or the ideal gas equation).

Example 20: Examples of unit conversions


(a) Express P = 709 mmHg in Pa.
(b) Express P = 33 atm in Pa.
(c) Express V = 22.4 ℓ in m3 .
Solution:
101325
(a) 709 mmHg = 709 × = 9.45 × 104 Pa.
760
101325
(b) 33 atm = 33 × = 3.34 × 106 Pa.
1
(c) 22.4 ℓ = 22.4 dm3 = 22.4 × 10−3 m3 = 2.24 × 10−2 m3 .

6.3 Dalton’s law of partial pressures


Consider a container of volume V filled with two non-interacting ideal gases A and B. The
partial pressures of each gas are
nA RT nB RT
PA = and PB = .
V V
The total pressure of the mixture is P = nRT /V , where n = nA + nB , and so
nRT
P =
V
RT
= (nA + nB )
V
nA RT nB RT
= + = PA + PB .
V V
This result can obviously be extended to any number of component gases.
Dalton’s law of partial pressures
The total pressure of a mixture of non-interacting gases is equal to the sum of the partial
pressures of the component gases:

Ptotal + P1 + P2 + P3 + . . . (36)

6.4 The pressure due to a vapour


Let us imagine that we have put a small quantity of liquid into an otherwise empty (evacuated)
cylinder fitted with a piston, and that we take a series of readings of pressure and volume,
keeping the temperature constant. We shall suppose that there is initially no free liquid.
A graph of pressure versus volume will then take the from shown in the figure below. At
a point such as A, where the volume is large and the pressure is small the substance is all in

27
the form of a vapour. As the vapour is compressed the variation of pressure with volume is
represented by the curve AB. Over this region the vapour is said to be unsaturated, and the
vapour obeys Boyle’s law to a good approximation. An unsaturated vapour is characterized
by the fact that a small isothermal compression does not result in any condensation of liquid.
When point B is reached the vapour becomes
saturated and the liquid appears in the cylinder

Pressure
in increasing quantities as it is compressed, but the D
pressure remains fixed. A decrease in volume does
not lead to an increase in pressure, but only the liquid
condensation of more liquid. The fixed pressure C
which is observed when a saturated vapour is in S.V.P. B
liquid &
equilibrium with its fluid is called the saturation saturated unsaturated
vapour pressure (S.V.P.) of the substance for the vapour vapour
particular temperature.
A
At point C the substance is entirely liquid, and a
small change in volume gives a very large change in Volume
pressure due to the low compressibility of a liquid.
Note that a liquid boils when its S.V.P. is equal to the pressure on its free surface. Thus
if a liquid is open to the atmosphere it boils at the temperature for which the S.V.P. equals
the atmospheric pressure.

6.5 The pressure due to a mixture of gases and a vapour


A mixture of a gas and an unsaturated vapour is fairly accurately described by the ideal
gas laws, right up to the point at which saturation occurs. Boyle’s law, for example, may to
a good approximation be used to describe the behaviour of a gas and an unsaturated vapour.
This is not so once the vapour is saturated and an excess of liquid is present. If a gas and
saturated vapour are compressed at constant temperature the vapour continues to exert a
partial pressure equal to the constant S.V.P. at the prevailing temperature, but the partial
pressure of the gas will obey Boyle’s law. Thus, if P is the total pressure of the mixture, and
V its volume, with f the constant S.V.P. at the prevailing temperature

(P − f )V = constant.

7 Elementary kinetic theory


7.1 Basic assumptions
The kinetic theory of gases attempts to explain the behaviour of gases in terms of the micro-
scopic properties of the individual atoms or molecules. We make certain reasonable assump-
tions about the nature of an ideal gas and then derive various theoretical results which are
consistent with the experimentally observed gas laws. These assumptions are:
1. A gas consists of a large number of identical particles which are in continual, random
motion.
2. The particles move in straight lines and observe Newton’s laws of motion.
3. The particles exert no forces on each other except at the instant of collision.
4. The collisions are perfectly elastic and of negligible duration.

28
5. The volume of the particles is negligible compared with the volume of the container.

7.2 The pressure exerted by an ideal gas


We shall use these assumptions to derive an expression for the pressure exerted by an ideal
gas in terms of the average kinetic energy of the particles. The pressure of a gas is due to the
particles bombarding the walls of the container. Whenever a particle bounces off a wall, its
momentum component perpendicular to the wall is reversed. The force exerted on the wall is
then the rate of change of the particle’s momentum. The pressure is obtained by calculating
the net force due to all of the particles and dividing by the area of the wall.

A2
A1

x
z

Consider a gas of N molecules each of mass m in a cubical vessel of side l. Suppose a molecule
has a velocity v which can be resolved into components vx , vy and vz , in the directions of the
edges of the box. If this molecule collides with wall A1 (the wall facing in the direction of the
+x axis), it rebounds with its x component of the velocity reversed. There will be no effect
on vy and vz , so the change in the molecule’s momentum will be

∆p = mvx − (−mvx ) = 2mvx

perpendicular to the wall.


Suppose this molecule reaches the wall A2 (the wall facing in direction of the −x axis)
without striking any other molecule along the way. The time required to cross the cube will
be l/vx . At A2 it again reverses direction and returns to A1 . Hence, assuming no collision in
between, the molecule will again strike the wall A1 after a time of 2l/vx . So the rate of change
of momentum is

∆p 2mvx mvx2
= = .
∆t 2l/vx l

But this is simply the impulsive force on the wall exerted by this one molecule. To obtain the
total force on wall A1 due to N molecules, one can write
m 2 2 2 2

F = v1x + v2x + v3x + · · · + vN x , (37)
l
2
where v1x is the square of the x component of the velocity for molecule 1, and so on.

29
The pressure P on the wall is the total force per unit area. Hence

F F 1 m 2 2 2 2

P = = 2 = 2 v1x + v2x + v3x + · · · + vN x
A l l l 
2 2 2 2
N m v1x + v2x + v3x + · · · + vN x
= 3 .
l N
2 2 2 2
 2
The quantity v1x + v2x + v3x + · · · + vN x /N is the average value of vx for all the particles in
the container. Let this value be denoted by vx2 . The pressure is then

N m vx2
P = .
l3
Furthermore, for any particle v 2 = vx2 + vy2 + vz2 . Since N is large, it is reasonable to suppose
that the average values of vx2 , vy2 and vz2 are equal and exactly one third of the value v 2 . Hence
one can say
1 2
vx2 = v ,
3
so that

N m v2 1 N m v2
P = 3 = .
l 3 3 l3
Now l3 is simply the volume V of the container, so therefore

1 N m v2
P = (38)
3 V
or
1
P V = N m v2 . (39)
3

7.3 The root-mean-square speed


2
In Equation (39) v 2 is the mean-square speed. It is not the same as v , which is the
square of the mean speed. We define the root-mean-square speed vrms as the square root of
the mean-square speed, i.e.
q
vrms = v2 . (40)

Example 21: The mean-square, root-mean-square and mean speed of molecules


Given 6 molecules with the following speeds: 1, 2, 3, 4, 5 and 6 m s−1 , calculate

(a) their mean-square speed,


(b) their root-mean-square speed and
(c) their mean speed.

30
Solution:
12 + 22 + 32 + 42 + 52 + 62
(a) v 2 = = 15.2 m s−1
6
2
(b) vrms = v 2 = 15.2 m s−1 ∴ vrms = 3.90 m s−1
1+2+3+4+5+6
(c) v = = 3.50 m s−1
6
Note that vrms 6= v .

We now have two gas equations, namely P V = nRT (Equation (35)) observed experimen-
tally and P V = 31 N m v 2 = 13 N mvrms
2
(from Equations (39) and (40)) from kinetic theory. To
make kinetic theory consistent with experiment, the two results must be equated. Suppose the
number of moles is one, i.e. n = 1. Then the number of particles equals Avogadro’s number
NA (N = NA ), so P V = nRT = 13 N mvrms 2
becomes
2
RT = 31 NA mvrms , (41)
where m is the mass of a single particle. The molar mass M = NA m, so that
r
3RT
vrms = , (42)
M
where M is in kilograms and T must be in kelvin.
From Equation (41)
1 2 3R
2
mvrms = × T,
2NA
where 21 mvrms
2
is simply the average kinetic energy of a single gas particle. Hence we obtain
an important result that the average kinetic energy of a particle depends only on the absolute
temperature because 3R/(2NA ) is a constant, which has the same value for all gases (R/NA is
called the Boltzmann constant kb ).

Example 22: The root-mean-square speed of molecules


Calculate the root-mean-square speed of

(a) an oxygen molecule at 27 ◦C.


(b) a hydrogen molecule at 27 ◦C.
Solution:
(a) The molar mass of oxygen is 32 g and the absolute temperature of the gas is T =
27 + 273 = 300 K. Hence
r r
3RT 3 × 8.314 × 300
vrms = = = 484 m s−1 .
M 32 × 10 −3

(b) For hydrogen, with molar mass of 2 g


r
3 × 8.314 × 300
vrms = = 1936 m s−1 .
2 × 10 −3

Notice that the mass of hydrogen is 16 times smaller than the mass of oxygen. Since
both molecules are at the same temperature they both have the same kinetic energy.
However, for this to be so the speed of the hydrogen must be 161/2 = 4 times the
speed of the oxygen.

31
7.4 The distribution of molecular speeds
Consider a container filled with oxygen at 300 K. We have just calculated that the root-mean-
square speed is 484 m s−1 at this temperature. Does this mean that every oxygen molecule in
the container has this speed? The answer is no. However, it does mean that the majority of
the molecules travel near 484 m s−1 . Some molecules have considerably higher speeds whilst
others have much lower speeds. The graph in Figure 4 shows curves for the distribution of
speeds in a sample of molecular oxygen at T = 300 K and T = 1200 K. Note the labelling of
the axes.

vmp
vrms
Nv , number of molecules per unit
speed interval molecules/(m s−1 )

T = 300 K

vmp
vrms

T = 1200 K

395 789 967 v(m s−1 )


484

Figure 4: Maxwell distribution for T = 300 K and T = 1200 K.

The graph in Figure 4 is called the Maxwell distribution and is interpreted as follows:

1. More molecules travel at the most probable speed vmp than at any other speed. For
oxygen at 300 K this speed is about 395 m s−1 . Note that the most probable speed vmp
is close to the root-mean-square speed vrms but they are not equal.

2. Some molecules travel at speeds considerably greater than this but they are relatively
few in number (similarly for low-energy molecules).

3. Increasing the temperature causes the most probable speed to shift to the right (higher
values) although the number of molecules travelling at this speed is less than the number
at 300 K.

4. The area under both curves is the same and represents the total number of molecules
present in the container.

32
8 Thermodynamics
8.1 Introduction
The science of thermodynamics is concerned with those processes in which heat is converted
into work and other kinds of energy, and vice versa.
Fundamental to the study of thermodynamics is the concept of internal energy. In general,
a sample of matter may possess both kinetic and potential energy. Consider a simple molecule
such as oxygen. The pictures below show some of the translational, vibrational and rotational
modes of this molecule. Each mode carries a certain amount of energy, the sum of which gives
the internal energy, U , of the substance.
It is very difficult, if not impossible, to evaluate the
internal energies of substances via this approach, other
than for monatomic gases. However, we can easily translation
evaluate changes in the internal energy of a substance
by noting a change in temperature, pressure, volume vibration
or state. Such changes in the internal energy can be
brought about by the transfer of heat into or out of rotation
the system, or by the performance of work on or by
the system.

8.2 The internal energy of an ideal monatomic gas


The internal energy of a substance is the sum of the various kinds of energy possessed by its
atoms or molecules. A monatomic ideal gas (e.g. an inert gas) is composed of single atoms
which do not interact (since there are no forces between the particles). Hence there is no
potential energy and all of the atoms’ energy is kinetic. As a result the internal energy U is
simply the total kinetic energy of the N atoms which constitute the gas. So
2 2
U = N 21 mvrms

or 2U = N mvrms .

From Equation (39)


1 2
PV = 3
N mvrms
1
= 3
× 2U = 32 U
3
∴ U= 2
PV
3
= 2
nRT
3
or ∆U = 2
nR∆T. (43)

In practice, the only gases for which this result is valid are the inert gases since they are
monatomic.

8.3 The first law of thermodynamics


The first law of thermodynamics is simply a statement of the conservation of energy. In the
mathematical formulation of this law, take:
Q as positive for heat added to the system
(I)
Q as negative for heat removed from the system

33
and
W as positive for work done on the system
(II)
W as negative for work done by the system.
With this sign convention, the change in internal energy ∆U may be written as

∆U = Q + W, (44)

where ∆U , Q and W are all measured in joules.

8.4 The work done by a gas


Consider an air-tight cylinder fitted with a (frictionless) piston and containing some gas as
shown below. If the system is heated, the gas expands and does work on an external agent.
Pressure

P constant

F = PA

Volume
s Vi Vf

For the gas inside the piston (i.e. the system), take Q as positive and W as negative.
If the pressure of the gas remains constant during the expansion (an isobaric process),
then the work done by the system can be expressed as

W = force × distance
= P A × s = P × As
= P ∆V

= P Vf − Vi ,

since the quantity As is the change in the volume (∆V = Vf − Vi ) of the expanding gas.
Now, if Vf > Vi , the work done W for the system is negative. Conversely, if Vf < Vi , W is
positive since work has been done on the system. Using the sign convention (II)

W = −P ∆V for an isobaric process. (45)

Note, this result is only true provided the pressure remains constant. For an ideal gas (P V =
nRT )

W = −P ∆V = −nR∆T for an isobaric process. (46)

For isochoric processes (i.e. constant volume) the gas cannot do any work since it cannot
expand. Hence

W = 0 for an isochoric process. (47)

34
For an isothermal process (i.e. constant temperature)
V2 V2
nRT V2
Z Z
W =− P dV = − dV = −nRt ln .
V1 V1 V V1

Since the internal energy of an ideal gas is proportional to the temperature (Equation (43)), it
remains constant throughout an isothermal process. Hence the change in the internal energy
∆U is zero and the first law of thermodynamics gives

∆U = Q + W = 0
∴ Q = −W.

8.5 The molar specific heat capacities of monatomic gases


From the section of heat capacity (Section 3.2), the temperature of a substance changes as a
result of heat flow. The change in temperature ∆T and the amount of heat Q are related by

Q = mc∆T,

where the specific heat capacity c is measured in J kg−1 K−1 . For gases it is more convenient
to express the amount of material as a number of moles n, rather than an amount of mass.
The above expression is replaced by

Q = nC∆T,

where C is the molar specific heat capacity of the gas measured in J mol−1 K−1 .
For gases — which can expand significantly when heated — there is a distinction between
the molar specific heat capacity at constant pressure CP and the molar specific heat capacity
at constant volume CV .

Constant Volume
When adding an amount of heat Q to a gas at constant volume all of the heat is used to change
the internal energy, since there can be no work done. From the first law of thermodynamics
and Equation (43)

∆U = Q + W = Q
∴ Q = 23 nR∆T.

But the heat that changes a substance’s temperature is

Q = nCV ∆T
∴ nCV ∆T = 32 nR∆T
CV = 23 R. (48)

Constant Pressure
Let the gas now expand at constant pressure. Some of the heat is used to do work (W =
−P ∆V = −nR∆T ). In order to bring about the same change in internal energy as before, a

35
certain amount of heat (Q = nCP ∆T ) must be added:

Q = nCP ∆T = ∆U − W
= 32 nR∆T − (−nR∆T )
= 25 nR∆T
∴ CP = 25 R. (49)

From Equations (48) and (49)

CP − CV = R. (50)

In deriving Equations (48) and (49), Equation (43) was used, which is true for monatomic
gases. Hence Equations (48) and (49) are only valid for inert gases. A more sophisticated
treatment is needed for molecules containing more than one atom. However, it can be shown
that Equation (50) is generally valid.

Example 23: The increase in internal energy for boiling water


Calculate the increase in the internal energy when 1.000 kg of boiling water at 100.0 ◦C is
converted into steam at the same temperature. Atmospheric pressure is 1.013 × 105 Pa and
the latent heat of vaporisation is 2.260 × 106 J kg−1 .
Solution:
Use the first law of thermodynamics

∆U = Q + W,

where Q = mLv and W = −P ∆V .


First find the change in the volume ∆V as the water changes to steam. 1.00 kg of water
occupies a volume of 1 litre = 10−3 m3 . 1.00 kg of steam occupies a volume V , given by

nRT
V =
P 
1000
18
× 8.314 × 373
=
1.013 × 105
= 1.701 m3 .

So the change in volume ∆V = 1.701 − 1.00 × 10−3 = 1.700 m3 . The change in internal
energy

∆U = Q + W
= mLv + (−P ∆V )
= 1.000 × 2.260 × 106 − 1.013 × 105 × 1.700
= 2.088 × 106 J.

Note: For this example, the result ∆U = 32 nR∆T cannot be used as it is only valid for a
monatomic, ideal gas, which steam is not.

36
In the above example, heat is used to do work on the surroundings (i.e. to ‘push back’ the
atmosphere as the volume increases). The ratio W/Q gives
W P ∆V
=
Q mLv
1.722 × 105
=
2.260 × 106
= 7.619 × 10−2 .
Therefore about 7.6 % of the heat does work on the surroundings and 92.4 % is used to change
the internal energy of the water.

Example 24: Heating of Argon at constant volume, then at constant pressure


One mole of argon at STP is heated at a constant volume until its pressure is tripled. It is
then heated at constant pressure to twice its original volume. Assume ideal gas behaviour.

(a) Illustrate these changes on a pressure-volume graph.


(b) How much heat is added to the gas in the entire process?
(c) Calculate the work done by the gas.
(d) Find the total change in the internal energy.

Solution:

(a) Let the three different states of the gas be labelled A, B and C. In state A, the gas is
at STP. Recall that one mole of any gas at STP occupies a volume of 22.4 litres. The
pressure is 101 325 Pa. The temperature is 273 K. In state B, the volume remains the
same (22.4 litres) and the pressure increases to 3 × 101 325 Pa. In state C, the volume
is doubled to 2 × 22.4 litres while the pressure is 3 × 101 325 Pa.
The pressure-volume graph is

3 B b b
C
P/atm

1 A b

V /ℓ
22.4 44.8

(b) Since ideal gas behaviour is assumed, the equation of state can be used:
PA VA PB VB PC VC
= = .
TA TB TC
From states A to B (constant volume)
PA PB
=
TA TB
PB
∴ TB = TA = 3 × 373 = 819 K.
PA

37
From states B to C (constant pressure)

VB VC
=
TB TC
VC
∴ TC = TB = 2 × 819 = 1638 K.
VB
The heat added to the system

Q = QAB + QBC
= nCV ∆TAB + nCP ∆TBC
= 1 × 32 R × (819 − 273) + 1 × 25 R × (1638 − 819)
= 2.38 × 104 J.

(c) For an isochoric process the work done is zero. Hence WAB = 0. For an isobaric
process the work done

WBC = −P ∆V
= −3 × 101 325 × (2 × 22.4 × 10−3 − 22.4 × 10−3 )
= −6.81 × 103 J.

The total work done by the gas is W = WAB + WBC = −6.81 × 103 J.
(d) The total change in the internal energy of the gas

∆U = Q + W
= 2.38 × 104 − 6.81 × 103 = 1.70 × 104 J.

38
TUTORIAL QUESTIONS
A. Thermometry and thermal expansion
Useful formulae
T = t + 273.15 relationship between Celsius and Kelvin temperature scales
(ℓ − ℓ0 )
α= mean coefficient of linear expansion of a solid
ℓ0 ∆t
(V − V0 )
β= mean coefficient of cubical expansion for solids and liquids
V0 ∆t
β = 3α for solids
βreal = βapparent + βvessel for liquids in a vessel
ρ0 the temperature dependence of the density of normal solids
ρ=
1 + β(t − t0 ) and liquids
V − V0 1 ◦ −1 zero coefficient of cubical expansion of a gas at constant pres-
β0 = = C
V0 t 273.15 sure
P
T (P ) = 273.16 for constant-volume gas thermometer
P0
Problems
A1 (a) Water freezes at 0 ◦C and boils at 100 ◦C at standard atmospheric pressure. The
corresponding temperatures in Fahrenheit are 32 ◦F and 212 ◦F. Prove that
TF = 59 TC + 32 ◦ ,
where TF and TC are the temperatures in Fahrenheit and Celsius respectively.
(b) At what temperature do the Fahrenheit and Celsius scales register the same tem-
perature? (−40 ◦ )
A2 If the ideal gas temperature at the steam point is 373.15 K, what is the limiting value
of the ratio of the pressures of a gas at the steam point and at the triple point of water
when the gas is kept at constant volume? (1.366)
A3 A constant volume gas thermometer reads 50 mmHg at the triple point temperature of
water.
(a) What will be the pressure when the thermometer measures a temperature of 300 K?
(b) What temperature corresponds to a pressure of 68 mmHg? (54.9 mmHg, 371.5 K)
A4 The values of the thermometric properties of different thermometers at the triple point
(T.P.) temperature of water, the normal boiling point (N.B.P.) of water and the normal
boiling point of nitrogen are listed in the table below.

Cu-Ni Pt resis. H2 const. vol. H2 const. vol.


thermocouple thermometer thermometer thermometer
ε/mV R/ohm P/atm P/atm
T.P. of water 2.98 9.83 6.80 1.00
N.B.P. of water 5.30 13.65 9.30 1.37
N.B.P. of nitrogen −0.10 1.96 1.82 0.29

39
Calculate the temperature of the normal boiling points of water and nitrogen which
would be recorded by the different thermometers if they were calibrated in accordance
with the general expression in terms of which temperature is defined quantitatively.
(N.B.P. of water (kelvin) 486 379 374 374
N.B.P. of nitrogen (kelvin) -9.2 54.5 73.1 79.2)

A5 To ensure a tight fit, the aluminium rivets used in aeroplane construction are made
slightly larger than the rivet holes and are cooled by ‘dry ice’ (solid CO2 ) before being
driven. If the diameter of a hole is 6 mm at 20 ◦C, what should be the diameter of a rivet
at 20 ◦C if its diameter is equal to that of the hole when the rivet is cooled to −78 ◦C,
the temperature of dry ice? β (aluminium) = 7.50 × 10−5 K−1 . (6.01 mm)

A6 At 20 ◦C a brass cylinder fits exactly into a circular hole of 5.58 cm diameter drilled into
a block of iron. What will be the width of the gap between brass and iron when they
are both cooled to −70 ◦C? α (brass) = 1.89 × 10−5 K−1 , α (iron) = 1.20 × 10−5 K−1 .
(1.8 × 10−2 mm)

A7 At room temperature a steel rod is 3.000 cm in diameter and a brass ring has an interior
diameter of 2.992 cm. By how much must the temperature of the rod and the ring
be changed so that the ring just slides on the rod at the final common temperature?
β (steel) = 3.60 × 10−5 K−1 , β (brass) = 5.67 × 10−5 K−1 . (389 ◦C)

A8 An aluminium measuring rod which is correct at 5 ◦C measures a certain distance as


88.42 cm at 35 ◦C. Determine the correct value for this distance. α (aluminium) =
2.50 × 10−5 K−1 . (88.49 cm)

A9 A thin copper wire is bent into a large circular ring with a small gap between the ends
of the wire. The gap is 1 mm wide at 20 ◦C. α (copper) = 1.66 × 10−5 K−1 .
(a) What will the width of the gap be if the temperature of the wire is increased
uniformly to 120 ◦C?
(b) Would it be possible to close the gap merely by cooling the ring? Explain your
answer. (1.002 mm)

A10 A simple pendulum consisting of a bob attached to a piece of brass wire has a length of
1.2 m at 15 ◦C. Calculate the change in the period of the pendulum when the temperature
p
changes from 15 ◦C to 20 ◦C. The period of the pendulum is given by T = 2π L/g,
where L is the length of the pendulum. α (brass) = 1.89 × 10−5 K−1 . (1.0 × 10−4 s)

A11 A spherical concave mirror is made of aluminium. What will be the fractional change
in the focal length of the mirror if its temperature is uniformly increased by 50 ◦C?
α (aluminium) = 2.50 × 10−5 K−1 . (1.25 × 10−3 )

A12 The coefficient of linear expansion of glass is 9.0 × 10−6 ◦C−1 . If a bottle has a volume
of 50.000 cm3 at 15 ◦C find its capacity at 25 ◦C. (50.014 cm3 )

A13 What mass of mercury must be put in a 100 cm3 glass bottle so that the remaining
volume of the bottle does not vary with temperature? Take the density of mercury at
the initial temperature as ρ = 1.36 × 104 kg m−3 , β (mercury) = 1.82 × 10−4 K−1 and
α (glass) = 8.0 × 10−6 K−1 . (0.179 kg)

40
A14 Calculate the density of mercury at 100 ◦C, given that the density at 0 ◦C is
1.36 × 104 kg m−3 and that its real coefficient of expansion is 1.82 × 10−4 K−1 .
(1.34 × 104 kg m−3 )

A15 The density of gold is 19.3 × 103 kg m−3 at 20 ◦C and the coefficient of linear expansion
is 14.3 × 10−6 K−1 . Calculate the density if gold at 90 ◦C. (19.24 × 103 kg m−3 )

A16 A density bottle is filled with glycerine at a temperature of 20 ◦C. If the bottle
has a volume of 50.0 cm3 at 0 ◦C, what mass of glycerine is contained in the bottle?
Take ρ (glycerine) = 1.26 × 103 kg m−3 , βr (liquid) = 5.20 × 10−4 K−1 and β (glass) =
2.40 × 10−5 K−1 . (6.303 × 10−2 kg)

A17 A glass vessel is filled with exactly 1000.0 cm3 of turpentine at 20 ◦C. What volume
will overflow if the temperature is raised to 86 ◦C? The coefficient of linear expansion
of the glass is 9 × 10−6 ◦C−1 and the coefficient of cubical expansion of turpentine is
97 × 10−5 ◦C−1 . (62.2 cm3 )

A18 A graduated pyrex flask is filled with motor oil up to the 100 cm3 mark at 20 ◦C.
What will be the volume of the oil at 120 ◦C as read on the graduations on the flask?
β (pyrex glass) = 9.6 × 10−6 K−1 , β (motor oil) = 9.0 × 10−4 K−1 . (109 cm3 )

A19 A body floats in a liquid the temperature of which is changed from 0 ◦C to 50 ◦C. What
is the fractional change in the volume of the liquid displaced by the body if the mean
coefficient of cubical expansion of the liquid is 4 × 10−4 K−1 ? (2 × 10−2 )

A20 A brass cylinder having a mass of 500 g weighs 374 g when fully submerged in a liquid at
20 ◦C. When the temperature of the liquid and cylinder is raised to 120 ◦C, the cylinder
weighs 379.3 g in the liquid. Calculate the coefficient of cubical expansion of the liquid.
β (brass) = 5.7 × 10−5 K−1 . (5.0 × 10−4 K−1 )

A21 The bulb of a mercury-in-glass thermometer has a volume of 0.5 cm3 and the distance
between successive degree marks is 2 mm. Find the cross-sectional area of the bore of
the stem. α (glass) = 1.0 × 10−5 K−1 , β (mercury) = 1.8 × 10−4 K−1 . (3.8 × 10−8 m2 )

A22 If a force of 8.9 × 103 N is required to extend a certain cast-iron bar 30 cm long by
0.25 mm at a constant temperature (about 25 ◦C), what is the least force that would
prevent it contracting when cooled from 50 ◦C to 0 ◦C? α (cast iron) = 1.1 × 10−5 K−1 .
(5.9 × 103 N)

A23 Estimate the change in pressure required to prevent water from expanding when its tem-
perature is raised from 10 ◦C to 20 ◦C. For water, β (10−20 ◦C, 1 atm) = 1.5 × 10−4 K−1 ,
B (1 − 25 atm, 15 ◦C) = 2 × 109 Pa where B is the bulk modulus. (30 atm)

A24 A sample of gas with a zero coefficient of cubical expansion β0 = 3.662 × 10−3 ◦C−1 has
a volume of 1000 cm3 at 0 ◦C and 760 mmHg. Calculate the volume V20 and V140 (at
constant pressure) occupied by this sample of gas at 20 ◦C and 140 ◦C. Show that

V140 − V20
β= 6= β0 .
V20 ∆t

(V20 = 1073.24 cm3 , V140 = 1512.68 cm3 , β = 3.412 × 10−3 ◦C−1 6= β0 )

41
B. Heat and calorimetry
Useful formulae
Q = M c∆T for an energy transfer not involving a phase change
Q = ML for an energy transfer involving a phase change
Energy lost = Energy gained in calorimetry experiments
∆Q ∆T for heat conduction in the steady state
= λA
∆t ∆x (∆T /∆x must be positive)
∆Q

H = ∆t = σεA T 4 − T04 for heat radiation
λ1 T 1 = λ2 T 2 Wien’s law
λmax T = 2.897 × 10−3 K m Wien’s law

Problems
Unless otherwise stated in the question, use the following data:
For water ρ = 1000 kg m−3
c = 4180 J kg−1 K−1
Lf = 3.35 × 105 J kg−1
Lv = 2.26 × 106 J kg−1
1 calorie = 4.18 J
R = 8.314 J mol−1 K−1
Stefan-Boltzmann constant σ = 5.67 × 10−8 W m−2 K−4
Standard temperature = 273.15 K
Standard pressure = 101 325 Pa = 760 mmHg
B1 A 50 g ice cube is taken from a refrigerator where its temperature was −10 ◦C and is
dropped into a glass of water at 0 ◦C. If no heat is exchanged with the surroundings,
how much water will freeze on to the cube? c (ice) = 2.3 × 103 J kg−1 K−1 . (3.43 g)
B2 A tumbler has a mass of 110 g and contains 190 g of water at 21 ◦C. An ice cube of mass
27 g at 0 ◦C is added and the water is stirred. Calculate the temperature of the water
when all the ice has melted, assuming that the heat exchanged with the surroundings is
negligibly small. c (glass) = 8.4 × 102 J kg−1 K−1 . (9.6 ◦C)
B3 500 g of ice at −10 ◦C are placed in a vessel containing 1 litre of water at 20 ◦C. If the
thermal capacity of the vessel and the heat exchange with the surroundings are both
negligibly small, what will be the mass of the remaining ice after equilibrium between
the ice and the water has bee established? c (ice) = 2.1 × 103 J kg−1 K−1 . (281 g)
B4 A block of dry ice of mass 2 kg at its sublimation temperature is used for making ice.
(a) What is the maximum mass of water, initially at 20 ◦C, which can be frozen to ice
at 0 ◦C? Assume the vaporized carbon dioxide does not exchange heat with the
water.
(b) If all the vaporized carbon dioxide is collected, what will its volume be at 0 ◦C and
atmospheric pressure, viz. 1.01 × 105 Pa?
Ls (dry ice) = 5.74 × 105 J kg−1 , Molar mass of CO2 = 44 g mol−1 . (2.75 kg; 1.0 m3 )

42
B5 A kettle with a 1 kW heater element contains 1 litre of water at 20 ◦C.
(a) If heat transfer to the surroundings is negligibly small, how long will it be before
the water reaches its boiling point at 100 ◦C? Neglect the thermal capacity of the
kettle.
(b) At what rate, in g/min, will the water evaporate if the power is left on?
(5 min 34 s; 27 g/min)

B6 A 400 g mass of brass which has been heated to a uniform temperature of 500 ◦C is
lowered into a vessel of thermal capacity 100 J K−1 and which contains 1 litre of water
at 20 ◦C. Some steam is generated locally and escapes into the atmosphere. The final
temperature of the water and the brass is 25 ◦C. What mass of water has evaporated?
The heat losses to the surroundings by the vessel and the water are negligibly small. For
brass: c = 390 J kg−1 K−1 . (20 g)

B7 A thermos flask contains liquid oxygen at its boiling point temperature, viz. −182.9 ◦C. A
mass of 25.2 g of mercury at 21.5 ◦C is poured into the liquid oxygen and the gaseous oxy-
gen produced is found to have a volume of 3068 cm3 at 21.5 ◦C and 748 mmHg, the room
temperature and the atmospheric pressure respectively. Calculate the specific heat of
fusion of mercury, assuming that some liquid oxygen remains in the flask when the evap-
oration ‘ceases’. N.M.P. (mercury) = −38.8 ◦C, c (liquid mercury) = 142 J kg−1 K−1 ,
c (solid mercury) = 126 J kg−1 K−1 , ρ (oxygen at STP) = 1.43 kg m−3 , Lv (oxygen) =
2.43 × 105 J kg−1 . (1.19 × 104 J kg−1 )

B8 A casting weighing 50 kg is taken from an annealing furnace where its temperature was
480 ◦C and plunged into a tank containing 360 kg of oil at a temperature of 27 ◦C. The
final temperature is 38 ◦C. The specific heat capacity of the oil is 2090 J kg−1 K−1 . What
is the specific heat capacity of the casting? Neglect the heat capacity of the tank itself
and any heat losses. (374 J kg−1 K−1 )

B9 A calorimeter of thermal capacity 25 J K−1 contains 100 g of water at 0 ◦C. A 200 g copper
cylinder and a 300 g lead cylinder, both at 100 ◦C, are placed in the calorimeter. Calculate
the final temperature attained, assuming that heat losses are negligible. c (copper) =
385 J kg−1 K−1 , c (lead) = 127 J kg−1 K−1 (21 ◦C)

B10 A copper calorimeter of mass 300 g contains 500 g of water at a temperature of 15 ◦C. A
560 g block of copper, at a temperature of 100 ◦C, is dropped into the calorimeter and the
temperature is observed to increase to 22.5 ◦C. Neglect heat losses to the surroundings.
(a) Find the specific heat capacity of copper.
(b) What is the water equivalent of the calorimeter? (381 J kg−1 K−1 ; 27.3 g)

B11 When 100 g of liquid ‘X’ at 44 ◦C were poured into a certain calorimeter weighing 100 g
and initially at 20 ◦C, the final temperature of the calorimeter and liquid was observed to
be 40 ◦C. In a second experiment the same calorimeter initially contained contained 20 g
of water at 20 ◦C, and the addition of 100 g of the liquid X at 36 ◦C to the calorimeter
and water yielded a system with a final temperature of 30 ◦C. What is the specific heat
capacity of liquid X? (2.09 × 103 J kg−1 K−1 )

B12 A liquid of density 850 kg m−3 passes through a ‘flow calorimeter’ at the rate of
8.0 cm3 s−1 . Heat is added by means of a 250 W electric heating coil and a temper-

43
ature difference of 15 ◦C is established in steady state conditions between inflow and
outflow points. Find the specific heat capacity of the liquid. (2.45 × 103 J kg−1 K−1 )

B13 A car weighing 103 kg travels on a level road at a speed of 72 km h−1 . It is brought to
rest by applying the brakes which act equally on its four wheels. The brake shoes, of
which there are two per wheel, each weighing 0.5 kg, are of steel of specific heat capacity
500 J kg−1 K−1 . (The shoe lining is ignored in this problem.) If 40 % of the work done
against friction is transformed into internal energy of the brake shoes, calculate their
rise in temperature. (40 ◦C)

B14 Given that the Howick waterfalls are 95 m high, and assuming that no heat is transferred
between the water and its surroundings, what temperature difference would you expect
between the water at the top of the falls and in the pool at the bottom? (0.22 ◦C)

B15 What is the minimum speed at which a lead bullet, at a temperature of 30 ◦C, must
strike a target in order that its temperature will rise to its melting point temperature of
327 ◦C on impact? c (lead) = 132 J kg−1 K−1 (280 m s−1 )

B16 An evacuated glass tube 1 m long contains a 10 cm long mercury column. If the tube is
rotated and raised 50 times so that each time the column of mercury falls down the full
length of the tube from one end to the other, what temperature change will the mer-
cury experience assuming that heat losses are negligible? c (mercury) = 138 J kg−1 K−1 .
(3.2 ◦C)

B17 Rays from the sun fall on a convex lens of 15 cm diameter and are brought to focus
on a blackened calorimeter containing 25 g of ice at 0 ◦C. If 46 % of the incident solar
radiation reaches the earth’s surface and the energy absorbed by the lens is negligibly
small, how long will it take before all the ice is melted? The solar radiant flux per m2
incident on the earth’s atmosphere (the solar constant) Is = 1.35 kW m−2 .
(12 min 41 s)

B18 The temperature in an air-conditioned room is maintained constant. Compare the rates
at which heat is conducted through brickwork of an exterior wall and the glass of the
windows, per unit area, if the thicknesses of the wall and windows panes are 20 cm
and 3.5 mm respectively. λ (brick) = 0.50 W m−1 K−1 , λ (glass) = 0.70 W m−1 K−1
(1.25 × 10−2 )

B19 One litre of water which is initially at 20 ◦C is heated by a 500 W immersion heater.
(a) Assuming that all the heat supplied by the heater is absorbed by the water, how
long will it take before the water starts to boil at 100 ◦C?
(b) At what rate, expressed in cm3 s−1 , will the water evaporate when boiling?
(c) If the immersion heater has a surface area of 0.01 m2 in contact with the water and
the thickness of its metal casing is 0.5 mm, what is the difference in temperature
between the inner and outer surfaces of the casing. The thermal conductivity of
the metal is 200 W m−1 K−1 ? (11 min 9 s; 0.22 cm3 s−1 ; 0.1 ◦C)

B20 A sealed plastic bag filled with 300 g of ice at 0 ◦C is suspended in a draught of air at
20 ◦C. The surface area of the bag is 400 cm2 and the plastic material is 0.2 mm thick.
Assume that the outer surface of the plastic is at the temperature of the air and calculate
(a) the time it will take for all the ice to melt, and

44
(b) the initial rate at which the temperature of the water rises after all the ice has
melted.
The thermal capacity of the plastic bag is negligible compared with that of the water.
λ (plastic) = 0.03 W m−1 K−1 . (13 min 55 s; 0.1 ◦C s−1 )

B21 A layer of ice 15 cm thick has formed on the surface of a deep pond. If the temperature
of the upper surface of the ice is constant and equal to that of the air which is at −12 ◦C,
determine the approximate time it will take for the thickness of the ice to increase by
0.2 mm. ρ (ice) = 910 kg m−3 , λ (ice) = 2.09 W m−1 K−1 . (5 min)

B22 Water is contained in an aluminium put on a hot-plate. The water which is at 100 ◦C
boils away at a rate of 20 g per minute.
(a) What is the rate of heat transfer to the water?
(b) If the base of the pot has a surface area of 80 cm2 and a thickness of 1.0 mm, what
is the temperature at the lower surface of the base?
λ (aluminium) = 210 W m−1 K−1 (753 W; 100.4 ◦C)

B23 The surface area of the tungsten filament of a 60 W lamp is 2 cm2 . If it is assumed
that all transfer of energy from the filament is by radiation and that the emissivity
of tungsten has a constant value of 0.3, determine the operating temperature of the
filament. (2049 K)

B24 The temperature of the filament of an incandescent lamp bulb is 2400 K. The emissivity
of the filament is 0.35.
(a) If the power of the lamp bulb is 100 W, what is the surface area of the filament?
(b) If the same lamp is operated at 200 W, what will the temperature of the filament
then be? (1.56 cm; 2833 K)

B25 Compare the rates at which heat is transferred from a body when at 1000 ◦C and when
at 500 ◦C, if it is situated in an enclosure at 20 ◦C. Assume that the emissivity of the
surface of the body does not change with temperature. (7.5)

B26 A tungsten filament of total emissivity 0.32 has a diameter of 0.10 mm and length 0.25 m.
At what temperature should it be operated if it is rated at 100 W? Where in the spectral
distribution of its radiation is the maximum likely to occur? (λmax T = 2.897 × 10−3 m K)
(2.89 × 103 K; 1 µm)

B27 The power per m2 radiated by a tungsten surface at a temperature of 2.45 × 103 K
is 0.50 MW m−2 . Calculate the power per m2 radiated by a black body at the same
temperature, and hence the average emissivity of tungsten at this temperature. What
does this enable you to say about the absorptivity of tungsten? (ε = 0.24 = α)

B28 Data needed for this question:


Radius of sun rs = 6.95 × 108 m,
Mean distance from the sun to the earth R = 1.50 × 1011 m,
Solar constant = 1353 W m−2 ,
σ = 5.67 × 10−8 W m−2 K−4 .
(a) Assuming that the sun radiates as a black body at temperature T , write an expres-
sion for its total power radiation.

45
(b) Write an expression for the intensity of solar radiation at a large distance R from
the sun.
(c) Use the data to estimate the sun’s surface temperature. (5774 K)
B29 A black body has a surface area of 10 cm2 and is maintained at a temperature of 227 ◦C.
(a) Calculate the radiant flux emitted by the body.
(b) Is this radiant flux equivalent to the rate at which heat is transferred from the
body? Explain your answer briefly. (3.5 W)
B30 A cylindrical metal can 10 cm high and 5 cm in diameter contains liquid helium at 4 K.
At this temperature the specific heat of vaporization of liquid helium is 2.09 × 104 J kg−1 .
The can is situated inside an evacuated chamber, the walls of which are maintained at
the normal boiling point temperature of nitrogen, viz. 77 K. A small diameter tube leads
from the helium can to the outside atmosphere.
If the radiant emittance of the can is 0.2 times that of a black body at 4 K, how much
helium is lost per hour? (1.3 g)

C. The ideal gas law and kinetic theory


Useful formulae
For a fixed mass of gas:

P1 V1 = P2 V2 at constant temperature Boyle’s law

V1 /T1 = V2 /T2 at constant pressure Charles’ law

P1 /T1 = P2 /T2 at constant volume Guy-Lussac’s law

P1 V1 P2 V2
= Generalization
T1 T2
P V = nRT or
Ideal gas law
P V = N kT where k = R/NA
1
2
2
mvRMS = 32 kT for a monatomic ideal gas

Ptot = P1 + P2 + P3 + · · · Dalton’s law of partial pressures

Problems
Unless otherwise stated in the question, use the following data:
Universal gas constant R = 8.314 J mol−1 K−1
Boltzmann constant k = 1.381 × 10−23 J K−1
Avogadro’s constant NA = 6.022 × 1023 mol−1
STP conditions T = 273.15 K (0 ◦C)
P = 1.013 × 105 Pa (1 std atm or 760 mmHg)
Molar volume of a gas at STP = 22.414 × 10−3 m3 mol−1
C1 A certain mass of an ideal gas occupied a volume of 4.00 m3 at 758 mmHg. Calculate its
volume at 635 mmHg if the temperature remains unchanged. (4.77 m3 )

46
C2 A given mass of ideal gas occupies 38 cm3 at 20 ◦C. If the pressure is held constant what
is the volume occupied at a temperature of 45 ◦C? (41.2 cm3 )

C3 A faulty barometer with some air above the mercury indicated a pressure of 74.0 cmHg
when the length of the unoccupied closed end of the tube was 20.0 cm. When the
barometer tube was pushed down into the mercury until the unoccupied length became
10 cm the barometer had a height of 72.0 cmHg. If the temperature remained constant,
deduce the true atmospheric pressure. (76.0 cmHg)

C4 On a certain day the true atmospheric pressure was 76.0 cmHg, and a faulty barometer
with some air in the tube above the mercury indicated a barometric height of 74.0 cmHg.
The length of the mercury-free glass tube above the mercury was 20 cm. When the tube
was pushed down into the mercury the length of the free space was reduced to 10 cm.
What was the new height of the mercury column in the tube? The temperature remained
constant for the duration of these observations. (72.0 cm)

C5 Given that the density of air at STP is 1.294 kg m−3 , determine the mass of air in your
physics laboratory, which has the dimensions 29 m × 11 m × 3 m, when the barometric
pressure is 71.0 cmHg and the temperature 20 ◦C (1.08 × 103 kg)

C6 What volume will 1.216 g of SO2 (with a molecular mass of 64.1) occupy at 18 ◦C and
755 mmHg if it behaves as an ideal gas? (456 cm3 )

C7 The pressure of air in a reasonably good vacuum might be 2 × 10−5 mmHg. If the
effective molar mass of air is 28, what mass of air exists in a 250 cm3 volume at this
pressure and 25 ◦C? How many molecules would this volume contain?
(7.53 × 10−9 g; 1.62 × 1014 )

C8 An evacuated tank with a capacity of 20 ℓ is filled with helium at room temperature,


viz. 27 ◦C. The mass of the tank is thereby increased by 0.10 kg. What is the pressure
registered on the tank’s pressure gauge? The atmospheric pressure is 1.0 × 105 Pa and
the molar mass of helium is 4.0 g (3.02 × 106 Pa)

C9 A 1.0 ℓ vessel is filled with helium at a pressure of 2 atmospheres and at a temperature


of 27 ◦C. The vessel is heated until the pressure of the helium is doubled. In the process
the volume of the vessel has changed by 1 %. Calculate (a) the final temperature of the
helium, and (b) the mass of helium present in the vessel. Molar mass of helium is 4 g.
(333 ◦C; 0.325 g)

C10 A 50 ℓ gas bottle contains 10 kg of oxygen. The bottle is fitted with a gas flow controller
which is adjusted so that oxygen is drawn at a constant rate of 200 mℓ per minute,
measured at 1 atmosphere and 20 ◦C.
(a) What is the initial pressure of the oxygen in the bottle?
(b) For how many days will the flow rate be maintained?
Molar mass of oxygen is 32 g. (1.52 × 107 Pa; 26 days)

C11 A flask contains 1.0 g of oxygen at an absolute pressure of 10 atmospheres and at a


temperature 47 ◦C. At a later time it is found that because of a leak the pressure has
dropped to 5/8 of its original value and that the temperature has decreased to 27 ◦C.
(a) What is the volume of the flask?

47
(b) What mass of oxygen has leaked out between the two observations?
Molar mass of oxygen is 32 g. (82 cm3 ; 0.33 g)

C12 A gas cylinder has a volume of 5.0 ℓ and contains sufficient hydrogen to exert a pressure of
100 atmospheres at 0 ◦C. Calculate the pressure in the cylinder when 9.0 g of hydrogen
have leaked out and the temperature is 39 ◦C. The density of hydrogen at STP is
0.089 g ℓ−1 . (91 atm)

C13 A vessel communicates with the atmosphere through a capillary tube. If the vessel is
initially at a temperature of 27 ◦C, to what uniform temperature must it be raised so
that 1/4 of the air molecules initially in the vessel will be expelled to the atmosphere.
Assume that the volume of the vessel remains unchanged. (127 ◦C)

C14 A vertical cylindrical tank 1.0 m high has its top end closed by a tightly fitting frictionless
piston of negligible weight and thickness. The air inside the cylinder is at an absolute
pressure of 1 atmosphere. The piston is depressed by pouring mercury on it slowly.
How far will the piston descend before mercury spills over the top of the cylinder? The
temperature of the air is maintained constant. ρ (mercury) = 13.6 × 103 kg m−3 and
g = 9.8 m s−2 . (24 cm)

C15 Two vessels of equal volumes are joined by a narrow tube and are filled with air at STP
and sealed. Calculate the pressure when one vessel is maintained at 0 ◦C and the other
is heated to 100 ◦C. The volume of the joining tube and the change in volume of the
vessel at 100 ◦C may be neglected. (88 cmHg)

C16 Two vessels each of volume 100 cm3 , one being at 27 ◦C and the other at 227 ◦C, contain
different mixtures of carbon dioxide and hydrogen at atmospheric pressure. The con-
tents of the two vessels are allowed to mix through a short connecting tube while the
temperature throughout becomes uniform at 27 ◦C. The final proportion by volume of
carbon dioxide in the mixture in both vessels is found to be 50 %. If the initial propor-
tion by volume of carbon dioxide in the vessel at 27 ◦C was 53 %, what was the initial
proportion by volume of carbon dioxide in the hotter vessel? Any change in the volume
of the containing vessel can be neglected. (45 %)

C17 A flask with a volume of 250 cm3 containing krypton gas at a pressure of 500 mm of
mercury is connected by a tube of negligible volume and a stopcock (initially closed) to
a second flask of volume 450 cm3 containing helium at a pressure of 950 mmHg. Calculate
the final pressure in the system if the stopcock is opened and the gases are allowed to
mix at constant temperature. (789 mmHg)

C18 A balloon of volume 600 m3 and negligible mass is to be filled with helium to a pressure
of 1 atmosphere at constant temperature.
(a) If the helium is stored in 50 ℓ cylinders at a pressure of 120 atmospheres, how many
cylinders will be required?
(b) If the atmospheric temperature is 27 ◦C, what mass of helium is contained in the
balloon?
(c) Calculate the load that can be supported by the balloon at an altitude where the
pressure is 0.5 atmosphere and the temperature −13 ◦C, given that the density of
air at STP is 1.3 kg m−3 . The balloon volume remains constant.
The molar mass of helium is 4 g. (100; 97 kg; 313 kg)

48
C19 A submarine sank at a point where the depth of water was 73 m. The temperature of
the water at the surface is 27 ◦C and at the bottom it is 7 ◦C. The density of the sea
water may be taken as 1.03 × 103 kg m−3 and the pressure at the surface as standard
atmospheric pressure.
(a) If a cylindrical diving bell 2.40 m high and open at the bottom is lowered to this
depth, to what height will the water rise within it when it reaches the bottom?
(b) At what gauge pressure must compressed air be supplied to the bell while on the
bottom to expel all water from it? (2.13 m; 7.27 atm)

C20 If the mixture of air and a saturated vapour is compressed at constant temperature to
a tenth of its original volume, and the pressure is increased from 15 cmHg to 60 cmHg,
what is the saturated vapour pressure? (10 cmHg)

C21 On a certain day the temperature was 28 ◦C and the barometric pressure was 76.0 cmHg.
A faulty barometer with some excess water droplets visible above the mercury indicated
a pressure of 71.0 cmHg. (The S.V.P of water at 28 ◦C is 3.0 cmHg.) The free length
of the tube above the mercury in this faulty barometer was 30.0 cm. If the barometric
tube were then pushed down into the mercury until the free length above the mercury
became 20.0 cm, what would be the new height of the mercury column? (70.0 cm)

C22 A closed vessel contains a mixture of air and water vapour in contact with excess water.
The pressure in the vessel at 27 ◦C and 60 ◦C are respectively 77.7 cmHg and 98.1 cmHg.
If the S.V.P. of water at 27 ◦C is 2.7 cmHg, what is the S.V.P. at 60 ◦C? (14.8 cmHg)

C23 Calculate the mass of oxygen contained in 2 ℓ of oxygen collected at 15 ◦C and 75.3 cmHg
by bubbling the oxygen through water. The S.V.P. of water at 15 ◦C is 13 mmHg and
the molecular mass of oxygen is 32. (2.637 g)

D. Thermodynamics
Useful formulae
∆U = Q + W First law of thermodynamics

∆U = 32 nR∆T Change in internal energy of an ideal monatomic gas

W = −P ∆V Work done at constant pressure (isobaric process)

W =0 Work done at constant volume (isochoric process)

V2
W = −nRT ln Work done at constant temperature (isothermal process)
V1
CV = 23 R molar specific heat capacity at constant volume

CP = 25 R Molar specific heat capacity at constant pressure

Relationship between molar specific heat capacities at


CP − CV = R
constant pressure and volume

49
Problems

D1 (a) Steam at a constant (absolute) pressure of 20 atmospheres is admitted to the cylin-


der of a steam engine. The length of the stroke is 60 cm and the diameter of the
cylinder is 20 cm. How much work is done by the steam per stroke?
(b) 1.0 kg of water, when converted to steam at atmospheric pressure occupies a volume
of 1.67 m3 . Calculate the work done against atmospheric pressure.
(−3.8 × 104 J; 1.69 × 105 J)
D2 A vessel with rigid, thermally insulating walls is divided into two equal parts by a
partition. One part contains a gas and the other is evacuated. What will be the change
in the internal energy of the gas if the partition is suddenly broken? Explain your answer.
D3 A steel tank contains 25 g of hydrogen at 15 ◦C and at a pressure of 1 atmosphere. The
hydrogen is heated to a temperature of 80 ◦C at constant volume. Determine (a) the heat
supplied to the hydrogen, and (b) the change in its internal energy. cV (hydrogen) =
1.00 × 104 J kg−1 K−1 . (1.63 × 104 J; 1.63 × 104 J)
D4 Given the following data on hydrogen determine the Boltzmann constant (do not assume
a value for R):
cP = 1.431 × 104 J kg−1 K−1
cV = 1.014 × 104 J kg−1 K−1
NA = 6.022 × 1023 mol−1
Molar mass of hydrogen = 2 g (1.38 × 10−23 J K−1 )
D5 The specific heat capacity of nitrogen at constant pressure is 1040 J kg−1 K−1 and its
density at STP is 1.25 kg m−3 . Calculate the constant volume heat capacity of nitrogen
assuming that it behaves as an ideal gas. (743 J kg−1 K−1 )
D6 Calculate the values of the principal specific heat capacities of hydrogen if their ratio
is 1.41 and the density of hydrogen at STP is 0.089 g ℓ−1 . Assume that the hydrogen
behaves as and ideal gas. (cV = 1.02 × 104 J kg−1 K−1 ; cP = 1.43 × 104 J kg−1 K−1 )
D7 One mole of argon gas, initially at STP expands isobarically to twice its initial volume.
(a) What is the final temperature of the gas?
(b) What is the work done by the gas in expanding?
(c) By how much does the internal energy of the gas change?
(d) Does the thermal energy leave or enter the gas? Calculate the quantity.
(546 K; −2.27 × 103 J; 3.40 × 103 J; 5.67 × 103 J)
D8 A mass of 100 g of benzene is initially at a temperature of 20 ◦C.
(a) How much heat is required to vaporize the benzene completely at its normal boiling
point temperature of 80 ◦C?
(b) Determine the amount of work done by the benzene during the vaporization process.
(c) Hence calculate the change in internal energy of the benzene during the process.
For benzene: c = 1.7 × 103 J kg−1 K−1
Lv = 3.9 × 105 J kg−1
ρ (liquid) = 9.0 × 102 kg m−3
ρ (vapour) = 2.7 kg m−3 . (4.92 × 104 J; 3.7 × 103 J; 4.55 × 104 J)

50
D9 Water is boiled at 100 ◦C at 1 standard atmosphere. Under these conditions 1 g of water
occupies 1 cm3 and 1 g of steam occupies 1.7 × 103 cm3 . Find the external work done
when 1 g of steam is formed at 100 ◦C, and calculate the increase in internal energy. For
water Lv = 2.26 × 106 J K−1 . (1.7 × 102 J; 2.1 × 103 J)
D10 (a) A student when resting produces about 60 kilocalories of heat per hour. Determine
the equivalent in watts.
(b) An electric power plant develops an average of 4 MW of power and burns 50 metric
tons of coal per day. The coal has a heat of combustion of 3.2 × 107 J kg−1 . What
is the overall efficiency of the power plant? (70 W; 22 %)
D11 An energetic athlete dissipates all the energy in a diet from 4000 kcal per day. If he were
to release this energy in the form of heat at a steady rate, how would the heat output
compare with that of a 100 W bulb if 98 % of the bulb’s output is heat? (ratio: 2.0)
D12 Two grams of helium (molecular mass = 4.0) expand isothermally at 77 ◦C and do 1600 J
of work. Assuming helium is and ideal gas, determine the ratio of the final volume of
the gas to the initial volume. (3:1)

P/Pa
D13 The pressure and volume of an ideal gas change
from A to B to C, as shown in the graph along- C B
side. Determine the total heat for the process and 4 × 105 b b

state whether the flow of energy is into or out of the A


2 × 105 b

gas. (80 kJ; outflow)

0.2 0.4 V /m3


D14 The drawing refers to one mole of a monatomic ideal gas and shows a process that has
four steps, two isobaric (A to B, C to D) and two isochoric (B to C, D to A). Complete
the following table by calculating ∆U , W and Q (including the algebraic signs) for each
of the four steps. Note that the gas has returned to its initial state at the end of the
process, so the value for the total ∆U can be predicted in advance without calculations.

∆U W Q
A to B
Pressure

B to C
A B
C to D
C 800 K D to A
D
400 K
200 K
Total
Volume

D15 A cylindrical aluminium rod has a length of 0.50 m and a radius of 3.0 cm. The temper-
ature of the rod is raised from 20 ◦C to 320 ◦C. How much work does the expanding rod
do on the surrounding air if the air pressure is 1.01 × 105 Pa? Take α for aluminium to
be 2.3 × 10−5 ◦C−1 . (2.96 J)
D16 A bubble from the tank of a scuba diver in a lake contains 3.5 × 10−4 moles of gas. The
bubble expands as it rises to the surface from a freshwater depth of 10.3 m. Assuming
the gas is an ideal gas and the temperature remains constant at 18 ◦C, find the amount
of energy that flows into the bubble. (0.59 J)

51

You might also like