Sunteți pe pagina 1din 102

PHYSICS 1 2002 S.

Nitsolov ii

TABLE OF CONTENTS
1 INTRODUCTION ...................................................................................................................................................1
1.1 INTRODUCTION .................................................................................................................................................1
1.2 MEASUREMENT ................................................................................................................................................1
1.3 THE INTERNATIONAL SYSTEM OF UNITS...........................................................................................................1
1.4 DIMENSIONS .....................................................................................................................................................2
1.5 SIGNIFICANT FIGURES ......................................................................................................................................3
2 KINEMATICS.........................................................................................................................................................4
2.1 INTRODUCTION .................................................................................................................................................4
2.2 FRAMES OF REFERENCE ....................................................................................................................................4
2.3 KINEMATICS OF A PARTICLE: DEFINITIONS ........................................................................................................4
2.4 TANGENTIAL AND NORMAL ACCELERATION .....................................................................................................7
2.5 MOTION WITH CONSTANT ACCELERATION ........................................................................................................8
2.6 ANGULAR KINEMATIC PARAMETERS .................................................................................................................9
2.7 ROTATION WITH CONSTANT ACCELERATION ...................................................................................................10
2.8 RELATIONS BETWEEN LINEAR AND ANGULAR KINEMATIC VARIABLES............................................................11
2.9 GALILEAN TRANSFORMATION .........................................................................................................................12
2.10 RELATIVE ROTATIONAL MOTION.....................................................................................................................13
3 NEWTON'S LAWS...............................................................................................................................................15
3.1 INTRODUCTION ...............................................................................................................................................15
3.2 THE LAW OF INERTIA ......................................................................................................................................15
3.3 MASS ..............................................................................................................................................................15
3.4 MOMENTUM AND THE PRINCIPLE OF CONSERVATION OF MOMENTUM .............................................................16
3.5 NEWTON'S SECOND LAW: FORCE .....................................................................................................................16
3.6 LAW OF ACTION AND REACTION .....................................................................................................................17
3.7 ANGULAR MOMENTUM AND TORQUE ..............................................................................................................17
3.8 DYNAMICS OF CIRCULAR MOTION ...................................................................................................................19
3.9 RELATIVE TRANSLATIONAL MOTION...............................................................................................................20
3.10 FUNDAMENTAL INTERACTIONS AND FORCES ..................................................................................................21
4 APPLICATION OF NEWTON'S LAWS ...........................................................................................................23
4.1 INTRODUCTION ...............................................................................................................................................23
4.2 APPLICATIONS OF THE LAWS OF MOTION ........................................................................................................23
4.3 GRAVITATION .................................................................................................................................................24
4.4 CONTACT FORCES ...........................................................................................................................................25
4.5 RESTORING FORCE ..........................................................................................................................................26
5 WORK AND ENERGY ........................................................................................................................................27
5.1 INTRODUCTION ...............................................................................................................................................27
5.2 WORK .............................................................................................................................................................27
5.3 POWER ............................................................................................................................................................28
5.4 KINETIC ENERGY ............................................................................................................................................28
5.5 POTENTIAL ENERGY ........................................................................................................................................29
5.6 RELATION BETWEEN FORCE AND POTENTIAL ENERGY ....................................................................................30
5.7 LAW OF CONSERVATION OF ENERGY ...............................................................................................................31
5.8 WORK OF NON-CONSERVATIVE FORCES ..........................................................................................................31
5.9 POTENTIAL ENERGY CURVES. EQUILIBRIUM ...................................................................................................31
6 SYSTEMS OF PARTICLES ................................................................................................................................33
6.1 INTRODUCTION ...............................................................................................................................................33
6.2 CENTRE OF MASS ............................................................................................................................................33
6.3 LINEAR MOMENTUM .......................................................................................................................................33
6.4 ANGULAR MOMENTUM ...................................................................................................................................35
6.5 KINETIC ENERGY ............................................................................................................................................38
6.6 WORK- ENERGY THEOREM ..............................................................................................................................39
6.7 PRINCIPLE OF CONSERVATION OF ENERGY ......................................................................................................39
6.8 COLLISIONS ....................................................................................................................................................40
7 MOTION OF RIGID BODY ................................................................................................................................42
7.1 INTRODUCTION ...............................................................................................................................................42
7.2 TRANSLATIONAL AND ROTATIONAL MOTION ..................................................................................................42
7.3 ROTATION OF A RIGID BODY ABOUT A FIXED AXIS ..........................................................................................42
7.4 ENERGY OF A RIGID BODY ...............................................................................................................................44
7.5 ANALOGY BETWEEN THE ROTATIONAL AND TRANSLATIONAL MOTION ..........................................................45
8 IDEAL GAS ...........................................................................................................................................................46
PHYSICS 1 2002 S. Nitsolov iii
8.1 INTRODUCTION ...............................................................................................................................................46
8.2 THERMAL EQUILIBRIUM AND TEMPERATURE ..................................................................................................46
8.3 THE IDEAL GAS EQUATION ..............................................................................................................................47
8.4 KINETIC THEORY OF THE IDEAL GAS ...............................................................................................................49
9 FIRST AND SECOND PRINCIPLE OF THERMODYNAMICS....................................................................52
9.1 INTRODUCTION ...............................................................................................................................................52
9.2 INTERNAL ENERGY AND WORK OF MANY-PARTICLE SYSTEM ..........................................................................52
9.3 WORK.............................................................................................................................................................52
9.4 HEAT ..............................................................................................................................................................53
9.5 FIRST LAW OF THERMODYNAMICS ..................................................................................................................53
9.6 HEAT CAPACITY ..............................................................................................................................................54
9.7 ENTROPY ........................................................................................................................................................55
9.8 CARNOT CYCLE...............................................................................................................................................57
10 ELECTRIC CHARGE AND ELECTRIC FIELD .............................................................................................59
10.1 INTRODUCTION ...............................................................................................................................................59
10.2 ELECTRIC INTERACTION AND ELECTRIC CHARGE ............................................................................................59
10.3 COULOMB'S LAW ............................................................................................................................................60
10.4 ELECTRIC FIELD ..............................................................................................................................................61
10.5 ELECTRIC POTENTIAL .....................................................................................................................................61
10.6 MOTION OF CHARGED PARTICLES IN ELECTRIC FIELD .....................................................................................63
11 GAUSS' LAW ........................................................................................................................................................65
11.1 INTRODUCTION ...............................................................................................................................................65
11.2 LINES OF FORCE AND EQUIPOTENTIAL SURFACES ............................................................................................65
11.3 FLUX OF THE ELECTRIC FIELD .........................................................................................................................66
11.4 FIELD PATTERNS .............................................................................................................................................68
12 CONDUCTORS AND DIELECTRICS IN AN ELECTRIC FIELD................................................................71
12.1 INTRODUCTION ...............................................................................................................................................71
12.2 CONDUCTOR IN AN ELECTRIC FIELD ................................................................................................................71
12.3 DIELECTRIC IN AN ELECTRIC FIELD .................................................................................................................72
12.4 ELECTRIC CAPACITANCE .................................................................................................................................75
12.5 ENERGY OF THE ELECTRIC FIELD ....................................................................................................................76
13 ELECTRICAL CURRENTS................................................................................................................................78
13.1 INTRODUCTION ...............................................................................................................................................78
13.2 CURRENT AND CURRENT DENSITY ..................................................................................................................78
13.3 OHM'S LAW .....................................................................................................................................................80
13.4 CONDUCTION OF ELECTRICITY IN METALS ......................................................................................................81
13.5 ELECTRIC POWER ............................................................................................................................................81
13.6 ELECTROMOTIVE FORCE .................................................................................................................................82
14 MAGNETIC INTERACTION .............................................................................................................................84
14.1 INTRODUCTION ...............................................................................................................................................84
14.2 MAGNETIC FIELD ............................................................................................................................................84
14.3 MOTION OF CHARGES IN A UNIFORM MAGNETIC FIELD ...................................................................................85
14.4 THE HALL EFFECT...........................................................................................................................................86
14.5 MAGNETIC FORCE ON AN ELECTRIC CURRENT.................................................................................................87
14.6 MAGNETIC TORQUE ON A CURRENT LOOP. ......................................................................................................88
15 SOURCES OF MAGNETIC FIELDS .................................................................................................................89
15.1 INTRODUCTION ...............................................................................................................................................89
15.2 MAGNETIC FELD OF A MOVING CHARGE..........................................................................................................89
15.3 MAGNETIC FIELD OF CURRENTS ......................................................................................................................90
15.4 AMPERE'S LAW ...............................................................................................................................................92
15.5 GAUSS' LAW....................................................................................................................................................93
15.6 MAGNETIC FIELD OF A SOLENOID ...................................................................................................................94
16 MAGNETIZATION OF MATTER .....................................................................................................................95
16.1 INTRODUCTION ...............................................................................................................................................95
16.2 MAGNETIC MOMENTS OF ELECTRONS AND ATOMS ..........................................................................................95
16.3 MAGNETISATION ............................................................................................................................................96
16.4 FERROMAGNETISM .........................................................................................................................................96
16.5 MAGNETIZATION VECTOR ...............................................................................................................................97
16.6 THE MAGNETIZING FIELD ................................................................................................................................98
16.7 ENERGY OF THE MAGNETIC FIELD ...................................................................................................................99
PHYSICS 1 2002 S. Nitsolov

1 INTRODUCTION

1.1 INTRODUCTION
The term physics comes from a Greek word meaning nature. From the time of the great ancient
philosopher Aristotle who defined physics as the study of all natural phenomena until now, the
subject of physics has undergone significant changes. Nowadays,
physics can be defined as the study of the most general properties of matter. Physics deals
with the structure and interactions of bodies, using the fundamental concepts of matter,
energy, space, time and motion and the relationships between them.
So-called classical physics developed until twentieth century is divided into mechanics,
thermodynamics and electromagnetism, which study motion, heat and electric and magnetic
phenomena, respectively. The revolution in physics in the end of nineteenth century modified the
basic concepts and principles and gave rise to theories of relativity and quantum mechanics.
These theories have provided a new revised view of natural phenomena, which is the base of
modern physics.
The progress in physics is due to the scientific method based on observation, hypothesis,
experiment, theory and prediction. The scientific investigation starts from observation, that is,
collection of data by use of the senses, or the extended senses and library research of the known
facts about the problem. The analysis of these data results in hypothesis, which is an assumption
having empirical consequences. The hypothesis can be confirmed only by experiment. An
experiment is a procedure carried out to study by observation under controlled conditions the
phenomenon in order to discover an unknown effect or law, to test or establish a hypothesis, or to
illustrate a known law.
The knowledge of particular phenomenon or process is generally incomplete unless it results in
quantitative information. The information in physics is expressed in mathematical form by means of
formulas and equations.
Because of its importance in sciences, an understanding of physics is required in many branches
of human activity and particularly in engineering. Physics develops the knowledge of the processes
and structure and functioning of materials, which are used by engineering to advance technology.
Thus, the goal of the course in physics for engineers is both an understanding of the concepts and an
ability to apply that knowledge in solving practical problems.

1.2 MEASUREMENT
Experimental method plays very important role in physics as well in all natural sciences. Any
experiment consists in the examination of a phenomenon under controlled conditions. To get
quantitative information from experiments and observations, we have to carry out measurements
of various physical properties like length, time, mass, temperature, electric current, voltage and so
on. The measurement is the act of determination of the numerical value of a certain physical
property by comparing its magnitude with its unit, which is a quantity of the same kind defined to
be exactly 1. The standard of unit is an object serving as reference with which all other examples
of the quantity are compared. For example, to measure a time interval we have to compare it with
the unit time called second. The standard of second was defined in 1967 as the duration of
9,192,631,770 vibrations of the light of specified wavelength emitted by a cesium-133 atom.

1.3 THE INTERNATIONAL SYSTEM OF UNITS


Since the physical quantities are not all independent they can be divided into fundamental and
derived. The derived quantities can be defined by mathematical relations involving only the
fundamental quantities. Thus the units of derived quantities are to be expressed in terms of the units
PHYSICS 1 2002 S. Nitsolov 2
and standards of the fundamental quantities. Any system of units includes (i) base quantities and
associated units and standards, (ii) method of forming smaller and larger and (iii) definitions of
derived quantities.
At present the International System of Units, with the abbreviation SI (for the French words
Systeme International) is officially adopted by the most of the world. It defines seven basic units for
seven most important quantities in physics. The complete list of basic SI units is given in Table 1.1.

Table 1.1
List of Base SI units

QUANTITY UNIT NAME UNIT SYMBOL


length meter m
time second s
mass kilogram kg
electric current ampere A
temperature kelvin K
amount of substance mole mol
luminous intensity candela cd

Table 1.2
List of Prefixes for SI Units

FACTOR PREFIX SYMBOL FACTOR PREFIX SYMBOL


10-21 zetta- Z 1021 zepto- z
10-18 atto- a 1018 exa- E
10-15 femto- f 1015 peta- P
10-12 pico- p 1012 tera- T
10-9 nano- n 109 giga- G
10-6 micro- µ 106 mega- M
10-3 milli- m 103 kilo- k
10-2 centi- c 102 hecto- h
10-1 deci- d 101 deka- da

To express the very large and very small quantities in SI, we can form larger and smaller units
using the prefixes listed in Table 1.2 in front of the base units. Attaching a prefix to a fundamental
or derived unit has the effect of multiplying by the associated power of ten. Thus, we can express
the wavelength of light of ruby laser as 0.000 000 694 m = 694 × 10-9 m = 694 nm.
The British engineering system is still in commercial use in the USA and in the most of English
speaking countries but it is in process of being be replaced by SI. The basic quantities in the British
system are length, mass and time. The correspondent base units are inch (defined to be 0.00254 m),
pound-mass (defined to be 0,4359237 kg) and second, respectively. The larger unit for length is the
foot equal to 12 inches.

1.4 DIMENSIONS
An equation connecting physical quantities such as x = y makes no sense, if x and y represent, for
example, mass and length since qualitatively different physical quantities can not be equated. The
dimension of a quantity is the physical property that it describes. The equality must connect only
quantities that have the same dimensions. The dimension of a particular quantity is denoted by
putting the associated symbol in brackets. For example, the dimension of x is denoted by [x]. We
use abbreviations [L], [M] and [T] for length, mass and time, respectively. Thus, if the symbol s is
used to denote a path, that is, a length, the equality [s] = [L] expresses the fact that the dimension of
s is length.
PHYSICS 1 2002 S. Nitsolov 3
Any physical quantity can be expressed as an algebraic combination of the basic SI quantities.
Similarly, the dimension of any quantity x can also be expressed as the algebraic combination of the
correspondent base dimensions of the kind
[x] = [LpM qT r] (1.1)
where p, q and r are positive or negative integers or fractions. If all the exponents are zero, the
quantity is dimensionless.
For example, the velocity in uniform motion is given by the equation
s
v= (1.2)
t
where s is the path and t is the time. The correspondent dimensional equation results in
[ L]
[v ] = = [ LT −1 ] (1.3)
[T ]
and we see that the unit of velocity is meter per second.
The analysis of dimensions provides a valuable check for various calculations. If the dimension
on one side of the equation is not the same as that on the other side, the equation is not correct.

1.5 SIGNIFICANT FIGURES


Physical quantities determined experimentally are never known with exact precision. Consequently,
in reporting the measurements it is necessary to indicate the degree of uncertainty. The precision of
a measured quantity can be expressed by the number of significant figures. A significant figure is
any digit in a numerical value of a quantity with exception of the leading zeros and trailing zeros
placed before the decimal point. For example, 0.307, 80.0, 3420 and 5.70 × 10-19 have three
significant figures. The farthest to right significant figure is called the least significant digit.
The number of significant figures with which a quantity is given depends on the accuracy of the
measurement. It can be assumed that the rounding off is done in such way that the error in the least
significant digit does not exceed 0,5. Thus, it can be inferred that the true values for numbers in the
previous example lie in the intervals (0,3065, 0,3075), (79.95, 80.05), (3419.5, 3420.5) and (5.695 ×
10-19, 5.705 × 10-19). If the least significant digit after the decimal point is zero, the exponential
notation should be used. For example, the quantity whose true value is in the interval (3419, 3421)
can be recorded as 3.420 × 103.

1.5.1 ADDITION AND SUBTRACTION


The least significant digit of the result is in the same position with respect to the decimal point as
the farthest to the left least significant digit of the input numbers. For example, writing the sum 11.0
kg + 1.13 kg + 2.18 kg in the form
11.0
+ 1.13
2.18
14.31
we see that the farthest to the left least significant digit is 0 in 11.0 and the least significant digit in
the result must have the sane position (highlighted) with respect to the decimal point. Then the
surrounding off gives 11.0 kg + 1.13 kg + 2.18 kg = 14.3 kg.

1.5.2 MULTIPLICATION AND DIVISION


The result has the same number of significant figures as are in the least precise input quantity. For
example, in the product (0.50 m)(124 s)/(11.2 m) the number of significant figures in the least
precise input quantity, 0.50 m, is two and the surrounding off the result on calculator
5,535714285714 gives (0.50 m)(124 s)/(11.2 m) = 5,5 s.
PHYSICS 1 2002 S. Nitsolov 4

2 KINEMATICS

2.1 INTRODUCTION
Motion is the most fundamental and common physical phenomenon. It consists in the change in
position of bodies or their parts relative to one another. The motion of a body is influenced by the
bodies surrounding it, i.e., by its interactions with them. Mechanics is a fundamental area of
physics, which studies the universal principles of motion. The aim of mechanics is to analyse and
predict the motion of bodies resulting from the different interactions. The nature of the particular
interactions is considered by other branches of physics.
Mechanics consists of two branches, kinematics and dynamics. Kinematics describes
geometrically the motion without considering the causes of motion. Dynamics studies the relation
between the motion and the causes of the motion. Sometimes a study of the equilibrium of bodies is
considered to be a separate branch of mechanics, called statics.

2.2 FRAMES OF REFERENCE


To describe the motion of a body means to
indicate its position in space at each instant in
time. The position of the body can be determined
only relative to another body. Because of that, an
observer must therefore indicate another body or
set of fixed bodies relative to which motion is
considered. To determine a time it is necessary to
use a clock. The combination of a body used as
reference and clock forms a frame of reference.
Usually, a set of coordinate axes is attached to the
frame of reference (Figure 2.1 ).
It should be noted that two observers associated
with two different frames of reference moving
relative to each other would make different
Figure 2.1 Frame of reference
observations of the motion of a given body. The
choice of frame of reference is determined by convenience. Usually, the motion of bodies is
referred to the Earth.

2.3 KINEMATICS OF A PARTICLE: DEFINITIONS

2.3.1 POINT-LIKE PARTICLE


In general, the position of a moving particle, P, in the
frame of reference XYZ with origin O (Figure 2.2) is
defined by its coordinates x, y, and z or by the
position vector r. A vector is a quantity that has both
a direction and magnitude. In this book we symbolise
the vectors with boldface letters. Introducing the unit
vectors i, j, k (i = j = k = 1) along the X- Y-
and Z-axis, respectively, we express the position
vector as
r =ix+jy+kz (2.4)
The position of a particle P may also be given by its
Figure 2.2 Position vector, curvilinear curvilinear coordinate, s, on the trajectory C. It is
coordinate and unit vectors.
PHYSICS 1 2002 S. Nitsolov 5
equal to the length of the path between an origin O' placed on the trajectory, and P, with sign "+" or
"-", depending on which side (positive or negative) of the origin the particle is on. As P moves, the
coordinates and position vector changes, that is, they are functions of time so that they may be
represented in the following form
x = x(t), y = y(t), z = z(t) (2.5)
r = r(t) (2.6)
s = s(t) (2.7)

2.3.2 DISPLACEMENT
Suppose that at time t the particle is at position P
with position vector r and at later time t', it is at
position P' with position vector r' (Figure 2.3).
The change in the particle's position between times
t and t' is described by the vector
∆r = r’ - r (2.8)
called the displacement of the particle during the
time interval ∆t = t’ - t. The vector of displacement
is directed from the initial position of the particle
to its final position. Using Eq. 2.4 we may write
the displacement in component form
∆r = i ∆x + j ∆y + k ∆z (2.9)
Figure 2.3 Vector of displacement and where ∆x = x' - x, ∆y = y' - y and ∆z = z' - z. The
displacement along the curve. displacement along the curve is defined by
∆s = s' - s. The magnitude of ∆s is equal to the length of the arc PP' and its sign depends on the
direction of motion of the particle along the trajectory.

2.3.3 VELOCITY AND SPEED


Velocity is a notion characterising how fast a given body moves. The average velocity of the
particle is defined as the ratio
∆r ∆x ∆y ∆z
v ave = =i +j +k (2.10)
∆t ∆t ∆t ∆t
where ∆r is the displacement during the time interval ∆t. From the definition, it follows that the
vector vave is parallel to the displacement ∆r.
Average speed is another quantity used to describe how fast the position of the particle changes.
It is defined as the ratio of the total distance covered along the path for a particular time interval to
that time interval, that is,
total distance
average speed = (2.11)
∆t
Since the average speed does not include direction it is always a positive scalar quantity. Unlike
velocity, speed refers only to how fast the distance increases, not to how fast the displacement
changes. Notice that the total distance travelled and the displacement along the path can be quite
different. For example, when the motion along a closed path ends at the same point it starts at, the
displacement is zero, but, clearly, the total distance covered is not zero.
The instantaneous velocity or simply, velocity at instant t is defined as the limiting value of the
average velocity as the time interval ∆t approaches zero:
∆r dr
v = lim v ave = lim = (2.12)
∆t → 0 ∆t → 0 ∆t dt
That is, the velocity of a particle is given by the first derivative of its position vector with respect to
time. In component form,
v = ivx + jvy + kvz (2.13)
where
PHYSICS 1 2002 S. Nitsolov 6
dx dy dz
vx = vy = vz = (2.14)
dt dt dt
If the coordinates are known functions of time, we can calculate the components of the velocity by
taking time derivatives.
The limiting process that defines the time
derivative is represented graphically in Figure 2.4.
As ∆t approaches zero, the positions P1, P2, P3, ...,
Pn of the particle and the correspondent
displacements relative to the limiting point P, are
shown. When Pn is very close to P, the
displacement coincides in direction with the
tangent, T, to the trajectory. Therefore, the velocity
is the vector tangent to the path. Furthermore,
when Pn is very close to P, the lengths of the arc
PPn and its chord are equal, that is
∆r=∆s as ∆t → 0 (2.15)
This relation may be transformed into the vector
equality
Figure 2.4 Limiting process ∆t→0.
∆r = uT∆s as ∆t → 0 (2.16)
where uT is a unit vector tangent to the path.
Substituting for ∆r from the previous equation into Eq. 2.12 yields
dr ds
v= = uT (2.17)
dt dt
Then the magnitude of the velocity may be expressed as
ds
v = (2.18)
dt
The instantaneous speed (or simply speed) is defined as the limit of the average speed as the time
interval approaches zero or,
∆s ds
speed = lim = (2.19)
∆t →0 ∆t dt
Combining Eq. 2.18 and 2.19 gives
speed = v (2.20)
That is, the speed is the magnitude of the velocity. In general, the speed and velocity vary with time
and we say that the motion is variable. If the speed is constant, the motion is often called uniform.

2.3.4 ACCELERATION
In general, when a particle is moving, its velocity
changes both in magnitude and in direction.
Acceleration is a notion, which characterises the
rate of change of the velocity. Suppose, that at
time t the particle is at point P with velocity v, and
at a latter time t' is at position P' with velocity v'
(Figure 2.5). The average acceleration is defined
by ratio of the change of velocity ∆v = v' - v and
elapsed time ∆t = t' - t,
Figure 2.5 The change in velocity between ∆v
times t and t'. a ave = (2.21)
∆t
The instantaneous acceleration, or simply acceleration, is defined by the limiting value of the
average acceleration as the time interval approaches zero, or
PHYSICS 1 2002 S. Nitsolov 7
∆v dv
a = lim a ave = lim = (2.22)
∆t → 0 ∆t →0 ∆t dt
Since the velocity is a time derivative of the displacement we may write
d 2r
a= 2 (2.23)
dt
In component form,
a = iax + ja y + kaz (2.24)
where
dv d2x dv y d 2 y dv d2z
ax = x = 2 , a y = = 2 , az = z = 2 (2.25)
dt dt dt dt dt dt
Since the velocity is tangent to the path its direction changes in the direction in which the trajectory
bends. The changes of the velocity and acceleration are always pointing toward the concavity of the
trajectory.

2.4 TANGENTIAL AND NORMAL ACCELERATION


Let us consider the acceleration of a moving
particle at point P of its trajectory (Figure 2.6).
Through the point P, a circle is drawn. This circle,
called tangential circle, most accurately
approximates the curve in a small neighbourhood
of the point. The tangential circle has common
tangent, T, with the curve at the point P. Its centre,
C, radius, R and the outward normal N to the curve
passing trough the centre are called centre of
curvature, radius of curvature and principal
normal for the point P, respectively.
Figure 2.6 Tangential and normal components We shall resolve the acceleration into two
of the acceleration. components, parallel to the tangent and to the
principal normal. They are called tangential and
normal or centripetal acceleration. To find the components of the acceleration we proceed from
Eq. 2.23. Its derivative with respect to time is
dv d dv du
a= = ( vuT ) = uT + v T (2.26)
dt dt dt dt
To calculate the time derivative of uT we consider
the motion of the particle from point P to P' in a
small time interval ∆t. The trajectory of motion in
a small neighbourhood of P may be considered an
arc of the tangential circle (Figure 2.7). As the
particle moves from P to P' the unit tangent vector
rotates together with the radius joining the particle
with the centre C. It can be seen from the figure
that vectors uT and uT' make the same angle θ as
the initial and final direction of the radius passing
through the particle. Therefore, the isosceles
triangles ∆PP'C and ∆AA'P are similar. From the
similarity it follows that AA'/AP = PP'/PC, so that
Figure 2.7 The change of the unit tangent PP ′
vector. ∆u T = (2.27)
R
Then, equating the magnitude of the displacement PP' to the length of the arc PP' and dividing by
∆t, we have
PHYSICS 1 2002 S. Nitsolov 8
∆u T 1 ∆s
= (2.28)
∆t R ∆t
As ∆t → 0, then θ → 0 also and the vector ∆uT is directed along the normal. Writing v = ds/dt, we
have
du T v
= uN (2.29)
dt R
where uN is a unit vector normal to the path. Thus, substituting Eq. 2.29 into 2.26, we obtain
finally the following equation:
a = aT u T + aN u N (2.30)
where
dv v2
aT = aN = (2.31)
dt R
are magnitudes of the tangential and normal acceleration, respectively. From this result we see that
the acceleration is a combination of normal and tangential vectors and lies in the plane of the
tangential circle. It always points towards the concavity of the trajectory. The magnitude of the
tangential acceleration is related to the change of the magnitude of the velocity. If the motion is
uniform, then aT = 0 and there is no tangential acceleration. The magnitude of the normal
acceleration is related to the change of the direction of the velocity. The direction of the normal
acceleration always points toward the centre of curvature of the path. When the motion is
rectilinear, the radius of curvature is infinite, and aN = 0, so that there is no normal acceleration.

2.5 MOTION WITH CONSTANT ACCELERATION


We shall consider a curvilinear motion in the case where the acceleration is constant, both in
magnitude and direction, a = constant. Let us assume that at the initial instant, t = 0, values of the
vector position and velocity are r(0) = ro and v(0) = vo. In order to determine the velocity we must
integrate the components of the acceleration in Eq. 2.25. Since a = const. the integration of the
x-component yields
vx (t ) = ∫ ax dt = c1 + ax t (2.32)
where c1 is a constant. Substituting this result into the initial condition vx (0) = vox , we obtain
c1 = vox so that vx = vox + ax t . Similar equations can be obtained for the y- and z- scalar components
of the velocity. Therefore, we may write
v = vo + at (2.33)
To find the vector position, we must integrate this result. From Eq. 2.33, we have
a t2
x = ∫ vx dt = c2 + vox t + x (2.34)
2
where c2 is a constant. From the initial condition x(0) = xo , we find c2 = xo and
a t2
x = xo + vox t + x (2.35)
2
It is easy to see that similar equations are valid for the y- and z- components. Thus, we obtain that
the displacement is given by
at 2
r − ro = vo t + (2.36)
2

The first term in the right-hand side is a vector parallel to the initial velocity and corresponds to a
uniform rectilinear motion of velocity vo. The second term is a vector parallel to the acceleration
and represents a uniformly accelerated motion with zero initial velocity and acceleration a.
Therefore, the displacement vector which is the sum of these two vectors must always lie in the
plane formed by vectors vo and a. Therefore, the uniformly accelerated motion can be considered as
a superposition of uniform rectilinear motion and uniformly accelerated motion with zero initial
PHYSICS 1 2002 S. Nitsolov 9
velocity; it is always in the plane defined by the initial velocity and acceleration.
According to the definition 2.10 the average velocity of uniformly accelerated motion for a time
interval (0, t) is given by
1
v o t + at 2
r − ro 2 1 v +v
v ave = = = v o + at = o (2.37)
t t 2 2
In order to determine the trajectory, we choose a
frame of reference so that the origin O coincides
with the position of the particle at the initial
instant to = 0. The trajectory of motion lies in
XY-plane and the Y-axis is parallel and opposite
to the acceleration (Figure 2.8). Then, Eq. 2.36
may be simplified to
1
r = v o t + at 2 (2.38)
2
The components of the position vector are given
Figure 2.8 Motion with constant acceleration. by
1
x = vox t − ax t 2 = vo cos θ t (2.39)
2
since ax = 0 and
1 1
y = voy t − a y t 2 = vo sin θ t − at 2 (2.40)
2 2
where θ is the angle between X-axes and vo.
Eliminating t from these equations, we have
a
y = tanθ x − 2 2 x 2 (2.41)
2vo cos θ
The result is an equation of a parabola with its
axis parallel to the Y-axis. Therefore, in general,
Figure 2.9 Trajectory in uniformly accelerated the trajectory of a uniformly accelerated
particle is parabolic. For example, this is the
motion, when θ varies from -900 to 900.
motion of a projectile launched with some initial
velocity at an angle θ, if we neglect the effect of air resistance. From Eq. 2.41 the trajectory
depends on the angle θ. Several trajectories for different equispaced values of θ are shown in Figure
2.9 where the directions of the initial velocities are indicated by arrows.

2.6 ANGULAR KINEMATIC PARAMETERS

2.6.1 ANGLE OF ROTATION


Circular motion is motion whose path is circle.
It is the motion of any particle of rotating objects
like wheel and flywheel. Consider a particle
moving in circle with centre C and radius R
(Figure 2.10). The radius joining the particle P
and the radius passing trough a fixed point A
make angle θ. This angle, called angle of
rotation, determines the position of the particle
on the circle. Usually, the angle θ is measured in
radians, (abbreviated rad). Its sign depends on the
Figure 2.10 Rotational motion of a particle. sense of rotation. It is positive when the rotation
is counter-clockwise and negative for rotation in the opposite sense.
PHYSICS 1 2002 S. Nitsolov 10
2.6.2 ANGULAR DISPLACEMENT
When the rotating particle changes its angular
position from θ at instant t to θ' at instant t'
(Figure 2.11), the increment of the angle of
rotation
∆θ = θ'- θ (2.42)
is called angular displacement. When we
consider infinitely small angular displacements, it
is useful to introduce a vector of angular
displacement, ∆θ. It has magnitude ∆θ and
Figure 2.11 Angular displacement, velocity direction perpendicular to the plane of circle. The
and acceleration.
sense of ∆θ is given by the right hand rule.
When the fingers of the right hand are directed along the path of the particle the thumb points to the
direction of the vector. To be more precise, the quantity ∆θ is not a true vector, but a
pseudovector. The direction of pseudovectors is related to that of rotation. When the type of system
of coordinates changes from a right-hand system to a left-hand one, the directions of pseudovectors
are reversed, while the directions of the true vectors remain unchanged. It should be noted that only
infinitely small angles of rotation can be considered as vectors.

2.6.3 ANGULAR VELOCITY


The instantaneous angular velocity, or simply, angular velocity at instant t is defined as
θ '− θ ∆θ d θ
ω = lim = lim = (2.43)
t ′→t t ′ − t ∆t →0 ∆t dt
The angular velocity is also a pseudovector. As it is shown in Figure 2.11, ω has the same direction
as the infinitesimal angular displacement dθ and its magnitude is equal to the time derivative of the
angle θ. The angular velocity is expressed in radians per second, rad⋅s-1. Remind that radian is
dimensionless unit. For that reason, sometimes angular velocity is expressed as s-1.
The instantaneous angular speed, or simply, angular speed is defined as the magnitude of
the angular velocity, so that
dθ dθ
angular speed = ω = = (2.44)
dt dt

2.6.4 ANGULAR ACCELERATION


The angular acceleration is defined as
d ω d 2θ
α= = 2 (2.45)
dt dt
It is also a pseudovector like the angular velocity. The angular acceleration and velocity have the
same direction when ω increases and are antiparallel when ω decreases.

2.7 ROTATION WITH CONSTANT ACCELERATION

2.7.1 EQUATION OF MOTION


When the circular motion is uniformly accelerated the angular acceleration a is constant. We shall
suppose that at the initial instant, t = 0, ω(0) = ωo and θ(0) = θo. Integrating Equation 2.45 with
respect to time as in the general case of uniformly accelerated motion, we obtain
ω = ωo + αt (2.46)
θ = θo + ωot + αt2 (2.47)
We see that these results are analogous to equations (2.33) and (2.36), respectively.
PHYSICS 1 2002 S. Nitsolov 11
2.7.2 PERIOD AND FREQUENCY
The motion with constant angular velocity is of special interest. Since α = 0, we find from Eq. 2.47
that the angle of rotation is a linear function of time
θ = θo + ωot (2.48)
If the magnitude of θ changes by 2π, the particle comes again to the same position. The motion
repeats itself every time the particle completes one revolution. That sort of motion is called
periodic. The period, T, of the rotational motion is the time required for one complete revolution,
that is, a rotation through the angle 2π. Since angle of rotation of 2π corresponds to one period, we
have

T= (2.49)
ω
The frequency ν is defined as the number of revolutions per unit time. Thus, period and the
frequency are reciprocal of each other
1
T= (2.50)
ν
and are related to the angular frequency by the equation

ω = 2πν = (2.51)
T

2.8 RELATIONS BETWEEN LINEAR AND ANGULAR KINEMATIC


VARIABLES
d In this section, we shall relate the linear
variables r, ∆r, v and a to the angular variables
P' ∆θ, ω and α. We consider a particle moving along
R' a circular path of radius R (Figure 2.12) and
dθ dr
O choose the origin O to be coincident with the
R P
centre of the circle (i.e., r = R). When a particle
Figure 2.12 Vectors of linear and angular rotates through an angle ∆θ, it covers distance ds,
displacement. given by the equation
ds = Rdθ (2.52)
Then, according to Equation (1.9), the magnitude of a very small displacement dr of the particle is
given by
dr = ds = Rdθ (2.53)
In order to express the relation between dr and
dθ in vector form, we shall use the properties of
vector product called also cross product. The
vector product a×b of the vectors a and b is
defined as a vector of magnitude
a×b = absinα
Figure 2.13 Vector product of vectors a and b
where α is the angle between the vectors Figure
2.13 . It is seen from the figure that the magnitude of a × b is equal to the area of the parallelogram
with sides a and b making angle α. The direction of the vector product is perpendicular to the plane
of a and b and is defined by the right hand rule: if the fingers of the right hand are directed along
the shortest path of rotation from the first multiplier, a, to the second multiplier, b, then the
extended thumb points in the direction of the vector a×b. Since a × b = −b × a the vector product is
not commutative.
We may, therefore, express the infinitesimal displacement as
dr = dθ×R (2.54)
The instantaneous velocity of circular motion may be expressed, according to (1.6), dividing the
infinitesimally small displacement dr by the corresponding time interval dt. Thus, we have
PHYSICS 1 2002 S. Nitsolov 12

v= × R = ω× R (2.55)
dt
Differentiating with respect to time, we obtain that the acceleration is given by
dv d ω
a= = × R + ω × v = α × R + ω × (ω × R ) (2.56)
dt dt
To simplify the double vector product, we apply the general formula a × (b × c) = b(a.c) - c(a.b),
taking into account that vectors ω and R are perpendicular. The result is
a = α × R - ω2R (2.57)
Since α is perpendicular to R, the magnitude of α × R is equal to αR and its direction coincides
with the tangent to the circle. Comparing Eq. 2.57 with the general decomposition a = aT uT + aN u N
dv v2
(Eq. 2.30) where aT = and aN = , we express the tangential and normal components of the
dt R
acceleration as
dω ν2
aT = α R = R aN = = ω2R (2.58)
dt R

2.9 GALILEAN TRANSFORMATION


Since motion is relative, it is important to know
how the quantities, which describe it, change when
passing from one frame of reference to another.
Let XYZ and X'Y'Z' (Figure 2.10) be two frames of
reference whose motion relative to each other is
translational. In translational motion any straight
line associated with the moving body remains
parallel to itself, that is, the frames do not rotate
relative to each other. For simplicity we shall
assume that at initial instant, t = 0, the two frames
of reference coincide. Therefore, their axes remain
parallel. We will compare the observations of
observers associated with different frames of
Figure 2.14 Two frames of reference moving reference. When an observer associated with XYZ
with relative translational motion. sees O' moving with position vector R and
velocity u = dR/dt, another observer associated
with X'Y'Z' sees O moving with position vector R' = - R and velocity u' = - u. Let P be a moving
particle with position vectors r = OP and r' = O'P relative to XYZ and X'Y'Z', respectively. As can
be seen from Figure 2.10, the vectors R, r and r' are related by the equation
r' = r - R (2.59)
This vector equality expresses the coordinates relative to X'Y'Z' in terms of coordinates relative to
XYZ. Assuming that the time flows the same in both frames, we have
t' = t (2.60)
Differentiating Eq. 2.59 with respect to time, we obtain
dr ′ dr ′ d r d R
= = − (2.61)
dt dt ′ dt dt
or
v' = v - u (2.62)
where v' and v are the velocities of the particle measured relative to XYZ and X'Y'Z'. This is the rule
for comparing the velocity as measured by the two observers. Taking the time derivative once
again, we have
dv dv ′ du
= − (2.63)
dt dt dt
or
PHYSICS 1 2002 S. Nitsolov 13
a' = a - aO (2.64)
where a' and a are the accelerations of the particle measured by observers associated with XYZ and
X'Y'Z' and aO is the acceleration of the origin O' relative to O. This is the rule for comparing the
accelerations as measured by the two observers.
In the special case when the relative translational motion is uniform rectilinear, u is constant so
that R = ut. Then we may write
r' = r - ut (2.65)
t' = t (2.66)
v' = v - u (2.67)
a' = a (2.68)
This set of equations is called the Galilean transformation of coordinates, time, velocity and
acceleration. According to the Galilean transformation
if two frames of reference are in relative uniform translational motion, then the acceleration
of a body remains invariant when passing from one frame of reference to another.
The Galilean transformation is correct only for velocities, which are small in comparison with the
velocity of light.

2.10 RELATIVE ROTATIONAL MOTION


Consider two frames of reference XYZ and X'Y'Z'
with coincident origins O and O' rotating relative to
each other (Figure 2.15). An observer in XYZ notes
ω that X'Y'Z' is rotating with angular velocity ω. The
position vector, velocity and acceleration of the
particle at point P determined by the observers in
XYZ and X’Y’Z’ are r, v, a and r’, v’, a’,
respectively. It can be shown that the transformation
laws for velocities and accelerations are given by
Figure 2.15 Frames of reference in relative v' = v - (ω × r) (2.69)
rotational motion. a' = a - 2(ω × v') - ω × (ω × r) (2.70)
The second and third terms in the right-hand side of
the last equation are called Coriolis acceleration and
centrifugal acceleration. It is possible to apply
Newton's laws to the rotating reference frame
introducing the correspondent inertial forces, called
Coriolis and centrifugal force, respectively.
Motion relative to the Earth (Figure 2.16) may be
analysed using the equations given above. The
angular velocity of the Earth is directed along the
axes of rotation from South to North. The period of
revolution of the Earth is approximately 24 hours.
Figure 2.16 Centrifugal acceleration on the
The angular velocity of the Earth is then
Earth surface.
ωE = 2π/(24 × 60 × 60 s) = 7.292 × 10-5 rad s-1.

2.10.1 CENTRIFUGAL ACCELERATION


All bodies on the Earth experience centrifugal
acceleration because of its rotation about its axis. The
centrifugal acceleration is directed outward, as shown
in Figure 2.16. It slightly reduces the effective
horizontal plane acceleration of gravity (from 9.83 m s-2 at the North
Pole to 9.78 m s-2 at the Equator).
Figure 2.17 Coriolis acceleration in the
Northern hemisphere.
PHYSICS 1 2002 S. Nitsolov 14
2.10.2 CORIOLIS ACCELERATION
The vector of Coriolis acceleration, - 2ω × v', is perpendicular to the vectors of the velocity v' and
angular velocity ω. For that reason in the
Northern hemisphere the Coriolis force tends to
deviate the path of a body moving horizontally
toward the right (Figure 2.17). The deviation in
the Southern hemisphere is to the left. For
example, rivers always wash out their right banks
in the Northern hemisphere. It is known that the
Danube erodes only the Bulgarian riverside, not
the Romanian one.
The Coriolis effect was demonstrated by
Figure 2.18 Rotation of a plane of oscillations. Foucault in 1851. He used a pendulum with large
length. The path of the bob is shown in Figure
2.18. The motion of the bob is deviated
continuously to the right. Therefore, at the end of
the first oscillation the pendulum starting from
point A in direction AA" goes to point A, not A".
On its return it reaches point B, not B''. As a
result the plane of oscillations rotates clockwise
and the path has the shape of a rosette. In the
Southern hemisphere the direction of rotation is
counter-clockwise.
The Coriolis effect also manifests itself in the
Northern hemisphere Southern hemisphere
rotational motion of storms around low-pressure
Figure 2.19 Rotational motion of storms due to centres. The wind initially blowing radially
Coriolis effect. toward the centre is deviated by Coriolis force, as
shown in Figure 2.19.
PHYSICS 1 2002 S. Nitsolov 15

3 NEWTON'S LAWS

3.1 INTRODUCTION
The study of motion in connection with its causes is called dynamics. Dynamics analyses the
relation between the properties of the body and the effects of the body's environment on the one
hand and the changes of the motion of the body on the other hand. The principles or laws of
motion were first stated by Sir Isaac Newton in 1687. Newton's laws should be considered as
extrapolations of the generalized results of numerous observations and experiments. They can not
be proved by any direct experiments. Nevertheless we are justified in using them because
everything they predict agrees with observed facts. These laws are valid for relatively large bodies
moving with velocities small compared with that of light.

3.2 THE LAW OF INERTIA


Newton's first law of motion states that
every body continues in its state of rest or of uniform motion in a straight line, unless it is
compelled to change that state by forces impressed upon it.
It is convenient to restate it using the notion of a free particle. It is a particle which does not
interact with other bodies or the interactions with the rest of the world cancel, giving zero net
interaction. In other words,
a free particle always moves with constant velocity,
or,
v = const. (3.1)
A free particle moves without acceleration in a straight line with constant velocity or is at rest (zero
velocity). The property of body to conserve its state of rest or uniform rectilinear motion is called
inertia. The Newton's first law is also called law of inertia.
This law and the rest Newton’s principles are obeyed not in any frame of reference. Thus, the
law of the inertia may be considered as a definition of the kind of frames of reference in which the
laws of mechanics hold (a free particle moves with constant velocity). From this point of view the
first Newton’s law can be expressed as follows: it is possible to find reference frame in which given
free particle has constant velocity. Such a frame of reference is called an inertial frame of
reference. We shall assume that the inertial frame of reference is associated with a free particle or
body and does not rotate since the rotation implies acceleration. According to Gallilean
transformations, the acceleration remains invariant in all frames of reference moving uniformly in a
straight line relative each other and a free particle moving without acceleration in one of these
frames will move with constant velocity in any other frame. Therefore, a frame of reference moving
uniformly in a straight line relative to an inertial frame will also be an inertial reference frame and
any frame moving with acceleration relative to an inertial reference frame is not an inertial
reference frame. It is accepted that a reference frame with axes fixed to the distant stars is an inertial
reference frame. The Earth rotates and a reference frame associated with the Earth's surface is not
strictly inertial. The acceleration of such a frame, however, is so small that it can be considered
inertial without too much error.

3.3 MASS
Mass was defined by Newton as the quantity of matter of a body. To find the mass of a body, we
must compare it with a standard body, whose mass is taken as unity. There are two methods of
measuring mass based on effects of gravitation and inertia.
PHYSICS 1 2002 S. Nitsolov 16
3.3.1 GRAVITATIONAL MASS
The notion of mass could be defined operationally using an equal-arm balance. The equality of
mass of two bodies is independent of the position and instant when the balance compares them. If
one body is composed of standard units the mass of the other one can be obtained. Mass obtained
by the use of gravitation is called the gravitational mass of the body.

3.3.2 INERTIAL MASS


The mass of a body may be defined dynamically as follows. Let us consider an isolated system of
two interacting particles of mass m1 and m2, which are subject only to their mutual interaction and
are isolated from the rest of the world. Assume, for simplicity, that the interaction of the particles is
of short duration and may be considered to be a collision. Before the interaction, at instant t, the
velocities of the particles are v1 and v2, respectively; after the interaction, at instant t', the velocities
are v1' and v2', respectively, (Figure 3.1). Experiments show that the increments of the velocities
∆v1 = v1' - v1 and ∆v2 = v2' - v2 are in opposite directions, that is, their unit vectors are related by
the equation
∆v1 ∆v 2
=− (3.2)
∆v1 ∆v 2
and the ratio of the magnitudes of the increments is
inversely proportional to the ratio of the masses of the
particles,
∆v1 m
= 2 (3.3)
∆v 2 m1
These equations may be combined to form one vector
Figure 3.1 Interaction of two particles equation,
m1∆v1 = -m2∆v2 (3.4)
This equation is valid for all interactions, not only for collisions. It can be used to determine the
mass m1, if the mass m2 and ratio |∆v1|/|∆v2| are known. The mass obtained in this way is called the
inertial mass. The inertial mass of a particle is a quantity that determines the changes of velocity in
interactions with other bodies. It is a measure of the inertia of a given body. Experiments show that
the inertial mass of a particle coincides with its gravitational mass. This is why we simply use the
term mass instead gravitational or inertial mass.

3.4 MOMENTUM AND THE PRINCIPLE OF CONSERVATION OF


MOMENTUM
The product of the mass and the velocity of a given particle is called its linear momentum or
simply momentum. Designating the momentum by p, we have
p = mv (3.5)
Using the notions of momentum and inertial frame of reference we may restate Newton's first law:
the momentum of a free particle is constant relative to an inertial frame of reference.
p = const. (3.6)
This statement is known as the principle of conservation of momentum. In Chapter 6 this
principle is generalised for systems of many particles.

3.5 NEWTON'S SECOND LAW: FORCE


Whenever there is a momentum change, such as in the case of a projectile, it is due to the
interactions of the body with other bodies. Such actions that can result in momentum change are
called forces. Newton's second law of motion gives the relation between the momentum change of
a particle and its interactions with other bodies. The force acting on a particle is related to its
momentum change by
dp
F= (3.7)
dt
PHYSICS 1 2002 S. Nitsolov 17
That is,
the rate of change of momentum of particle equals the force acting on the particle.
The force, as characteristic of an interaction, has some independent properties. It may be expressed
in terms of the distance between the two particles, their masses, charges etc. Such an expression is
called the force law. In the following chapters we shall consider particular force laws. The unit of
force in the SI system is the Newton, N = kg m s-2.
If the mass of the particle is constant, then we have
dp d (mv ) dv
F= = =m = ma (3.8)
dt dt dt
This is another more restrictive form of Newton's second law.
If the particle is free, there is no action on it, F = 0, and according to Eq. 3.8, p = const. This
result coincides with the law of inertia. It seems that, the law of inertia is contained in the Newton's
second law as a particular case. Nevertheless, the first law is formulated independently of the
second one because it postulates the existence of inertial frame of reference.

3.6 LAW OF ACTION AND REACTION


Dividing both sides of Eq. 3.4 by the corresponding time interval ∆t = t' - t, we obtain
∆p ∆p
1
= 2
(3.9)
∆t ∆t
Finding the limit of this equation as ∆t → 0, we may write
F12 = - F21 (3.10)
where F12 = dp1/dt is the force on particle 1 due to the interaction with particle 2 and F21 = dp2/dt is
the force on particle 2 due to the interaction with particle 1. This relation constitutes Newton's
third law of motion, also called the law of action and reaction:
the forces exerted by interacting particles on each other are equal in magnitude and opposite
in directions.
The law of action and reaction implies that the forces exerted by interacting bodies appear only
in pairs, called action-reaction pairs. The forces in such pair are equal and opposite; one of them
(either one) is called the action force and the other is called the reaction force. Action force and
reaction force always act on different bodies. Newton's third law is a consequence of the second law
and the principle of conservation of momentum and therefore it also must be considered relative to
an inertial frame of reference.

3.7 ANGULAR MOMENTUM AND TORQUE


In kinematics we described the circular motion by notions of angle of rotation, angular velocity and
angular acceleration, which are analogous to displacement, velocity and acceleration. Similar
analogy exists also in dynamics. To formulate the laws of motion of rotating bodies it has been
found convenient to use notions of angular momentum and torque, which correspond to linear
momentum and force. Note that these variables can be applied to describe any motion of a particle
not only to the circular motion.
PHYSICS 1 2002 S. Nitsolov 18
3.7.1 ANGULAR MOMENTUM
The angular momentum is a quantity characterising
the particle motion, which obeys law of conservation
like the angular momentum. Moreover, the angular
momentum conserves in cases in which the linear
momentum varies with time like motion under the
action of central force or circular motion.
Consider a particle with linear momentum p = mv
located at point P (Figure 3.2). We define the
angular momentum with respect to a fixed point O
for the particle by the relation
L = r×p = r×mv (3.11)
where r is the position vector of the particle relative
Figure 3.2 Angular momentum of a to O. Since the definition of angular momentum
particle. contains the position vector of the particle, the value
of L depends on the reference point O. According to
the properties of the vector product, the direction of the angular momentum is perpendicular to the
plane of r and v and the sequence of r, v and L forms a right-hand system. Its magnitude is
L = mrvsinθ = mr⊥v = mrv⊥ (3.12)
where θ is the angle between the vectors r and v, r⊥ is the component of r perpendicular to the
velocity (r⊥ = rsinθ)) and v⊥ is the component of the velocity perpendicular to r (v⊥ = vsinθ)).
Any one of these forms can be used to calculate the magnitude of the angular momentum. The SI
unit of angular momentum is kg⋅m2/s.

3.7.2 TORQUE
Taking the time derivative of Equation 3.11, which defines the angular momentum, we obtain
dL dr dp
= ×p +r× (3.13)
dt dt dt
dr
Since vectors = v and p = mv are parallel, then the first vector product in the right-hand side is
dt
dp
zero. Substituting = F , we may write
dt
dL
=τ (3.14)
dt
where the vector
τ = r×F (3.15)
is called the torque or moment of the force F with respect to a fixed point O. In SI the torque is
measured in N⋅m. Notice, that it is presupposed that L and τ are measured relative to the same point
fixed to an inertial frame of reference. The relation 3.14 may be expressed as
the time rate of change of the angular momentum of a particle is equal to the torque of the
force acting on that particle.
According to the definition, the torque is
perpendicular to the plane determined by vectors F
and r, and it is directed according to the right-hand
rule (Figure 3.3). The magnitude of torque is given
by
τ = rFsinθ (3.16)
where θ is the angle between vectors r and F. We
can rewrite Eq. 3.16 as
τ = r⊥F= rF⊥ (3.17)
Figure 3.3 Torque of a particle. where F⊥ is the component of F perpendicular to r
PHYSICS 1 2002 S. Nitsolov 19
(F⊥ = Fsinθ) and r⊥ is the component of r perpendicular to F (r⊥ = rsinθ)).
Eq. 3.14 gives the relationship between the
interactions of the particle with its environment
and its angular momentum. It is analogous to
Newton's second law, with the linear momentum p
replaced by L, and the force F replaced by the
torque τ. The torque characterises the ability of the
force to rotate a body about a fixed point. We can
think of the rotation of a particle attached to the
end of a massless rod that can rotate freely about
the other end attached to the origin O (Figure 3.4).
The acceleration of the particle due to the action of
Figure 3.4 A particle attached to the end of a the force F is directed perpendicular to the rod
massless rod that can rotate freely about the because the longitudinal component of the force is
other end. compensated by the reaction of the rod. Thus, we
see that the resultant motion of the particle caused
by the force is rotation about an axis passing trough the origin and perpendicular to the rod. That is,
the axis of rotation is parallel to the torque of the force relative to the origin.
Consider the motion of a particle subjected to a resultant force whose torque is zero. It is a
physically important case because, if the torque is zero, by the law 3.14, L = const, the angular
moment is independent of time and is conserved. This is the case of a free particle. The
conservation of angular momentum for a single particle is not an independent statement; it is a
consequence of the Newton's first law of motion. However, for more complicated physical systems
the law of conservation of angular momentum is fundamental principle.
The torque can also be zero for nonzero force when the force and position vector relative to the
reference point are parallel (or antiparallel). A particle moving with uniform circular motion has
only normal acceleration of constant magnitude and the force acting on it has also constant
magnitude and is directed to the center of the circle. Therefore, the torque of the force relative to the
centre of the circle is zero. In the case of a particle subjected to a central force, the torque relative to
the centre of force is zero because the force and r are parallel. In all these cases, the angular
momentum is constant and according to the properties of vector product, the vectors r and L = r × p
are always perpendicular. Since L = const, its direction does not vary and the position vector must
lie in a fixed plane perpendicular to L and passing trough the reference point O. Therefore, if the
angular momentum is constant, the trajectory is a plane curve perpendicular to L. This is an
important property of the motion under the action of central forces like electrical and gravitational
forces.

3.8 DYNAMICS OF CIRCULAR MOTION


Figure 3.5 shows a particle moving in circle with centre C and radius R. We consider the angular
momentum of the particle and the torque of the force
Z acting on it relative to a reference point O on the axes
v of rotation at distance d from the centre C. Using
F Equation 3.11, we represent the angular momentum
R as
L L = r×mv = m(OC + R)×v (3.18)
C
r Taking into account that the vectors OC (OC = d),
R and v = ω × R (Eq. 2.55) are mutually
perpendicular, we obtain the following
decomposition of the vector L into axial and radial
Y components:
O
L = LR + Lz
X LR = mOC×(ω×R) = - mωdR (3.19)
Figure 3.5 Angular momentum and torque of a Lz = mR×(ω×R) = mR2ω
particle moving along a circular path.
PHYSICS 1 2002 S. Nitsolov 20
The radial component is perpendicular to the axes of rotation and its magnitude is zero, if the origin
O coincides with the centre of the circle. The magnitude of the axial component of the angular
momentum
Lz = mR2ω (3.20)
is independent on the position of the origin O on the axes of rotation and is called the angular
momentum with respect to an axis of rotation. Equation 3.20 is similar to the definition of linear
momentum p = mv but instead of mass the property of inertia is described by the quantity
I = mR2 (3.21)
called moment of inertia of a particle with respect to the axes of rotation. The SI unit of moment of
inertia is kg⋅m2.
To find the equation of motion of a particle rotating in a circle, we take the time derivative of
Equation 3.20.
dLz d dω
= (Iω ) = I = Iα (3.22)
dt dt dt
We now use Equation 3.14 to substitute for dLz/dt. This gives
τz = Iα (3.23)
We see that the quantity τz called the torque or moment of the force F with respect to an axes Z is
analogous to the force in the Newton’s second law. To calculate τz, we resolve the force
F = Fxy + Fz acting on the particle into a component in the XY-plane, Fxy and an axial component, Fz
(Figure 3.6). Then the torque 3.15 can be written as
τ = r × F = r × Fxy + r × Fz (3.24)
Since the vector r × Fz is perpendicular to the
Z-axes the component of τ along Z-axes is equal to
τz = (r × Fxy)z (3.25)
Substituting r = OC + R and taking into account
that the vector OC × Fxy lies in the XY-plane and
R × Fxy is parallel to Z-axes, we have
τz = [(OC + R) × Fxy]z = R × Fxy (3.26)
The result can be written in the form
τz = RFxysinθ = R⊥Fxy (3.27)
where θ is the angle between R and Fxy and
R⊥ = Rcosθ is the perpendicular distance from the
Figure 3.6
axes of rotation to the line of action of the force
Fxy. Often R⊥ is called moment arm or arm of the force. The definition of the torque with respect
to an axis expresses the fact that the torque of a force is a measure of its ability to rotate a body
about a fixed axes and the rotating effectiveness of the force increases with the moment arm. For
instance, when we tighten a nut, we push or pull the handle of the spanner as far as possible from
the nut and apply our force in direction perpendicular to the handle.

3.9 RELATIVE TRANSLATIONAL MOTION

3.9.1 GALILEAN PRINCIPLE OF RELATIVITY


We mentioned above that the laws of motion must be considered relative to an inertial frame of
reference. We shall now analyse how the laws of motion transform when the inertial frame of
reference is replaced with another frame of reference. Let us consider an inertial frame of reference
XYZ and a frame of reference X'Y'Z' which is in translational motion with constant velocity u
relative to XYZ. According to Galilean transformation of velocity (Eq. 2.47), a free particle moving
relative to XYZ with constant velocity v, in X'Y'Z' has velocity v', given by
v' = v - u = const. - u (3.28)
When u is constant, the right-hand side of the above equation is also constant. Therefore, the law of
inertia is valid in each frame of reference, which is in uniform translational motion relative to an
inertial frame of reference. If we consider the law of inertia as a definition of an inertial frame of
PHYSICS 1 2002 S. Nitsolov 21
reference, we can conclude that
any frame of reference moving with a rectilinear uniform motion relative to an inertial frame
is also inertial.
Let us now discuss the relation between the force and the acceleration of a particle in the two
frames of reference. In the case of constant relative velocity, we can use Galilean transformation of
acceleration (Eq. 2.68), a' = a. Therefore, the forces F = ma and F' = ma', measured in XYZ and
X'Y'Z', respectively, are equal, F' = F. Thus, we can conclude that
the laws of motion are the same relative to all inertial frames of reference.
This statement is called Galilean or classical principle of relativity.

3.9.2 INERTIAL FORCES


Let us now discuss the relation between the force and the acceleration in the general case when the
velocity u of X'Y'Z' relative to the inertial frame of reference XYZ is not constant in time. Although
that Newton's laws of motion are valid only for inertial frames of reference, we use for both frames
the same definition of force, F = ma and F' = ma'. Substituting the values of a = F/m and a' = F'/m
into Eq. 2.64, we obtain
F' = F - mat = F - Fi (3.29)
We see that the force measured by an observer attached to the non-inertial frame of reference differs
from that of an inertial observer. The additional term Fi = mat , arising from the accelerated motion
of the non-inertial frame, is called an inertial force. It should be noted that the inertial forces are
fictitious. They are not result of the interaction of bodies, but are due to the accelerated motion of
the frame of reference. In contrast to real forces, which obey the third Newton’s law and always
arise in action-reaction pairs, inertial forces have not identifiable source or agent of interaction and
can be seen only by an accelerating observer.

3.10 FUNDAMENTAL INTERACTIONS AND FORCES


The forces, which act between the bodies surrounding us, are manifestations of the microscopic
interactions between very large number of particles. All forces observed in nature may be reduced
to several fundamental forces or interactions from which the secondary forces are derived. Until
recently gravitational, electromagnetic, strong and weak interactions were considered as
fundamental. The gravitational interaction holds stars in systems, called galaxies, and in turn it
holds the galaxies in clusters of galaxies.
The electromagnetic interaction is underlying for forces between atoms and molecules and for
many forces between macroscopic objects. The weak interaction is responsible for the decay of
elementary particles. For example, emission of electron by radioactive nucleus is due to the weak
forces. The strong interaction binds together the components of atomic nucleus.
The origins of fundamental interactions are called sources. According to the law of action and
reaction, the source is both the origin of forces and the subject on which the forces act. All sources
may be considered as systems composed of point-like sources. That is why the general laws of
forces are usually formulated in terms of different point-like sources associated with different
particle properties. The gravitational, electromagnetic, strong and weak interaction has their sources
in the mass, electric charge, colour charge and weak charge, respectively.
Many of forces among many sources may be reduced to a superposition of forces between two
sources, called two-body forces. The magnitude of the two-body forces depends on the distance
between the two particles. Some of two-body forces are central. A force whose direction passes
through a fixed point is called a central force and the fixed point is called the centre of force. For
example, the gravitational and electromagnetic forces are central with centres in the sources of
forces.
The operation of fundamental forces may be explained in terms of so-called exchange model of
forces. According to this model, the sources emit and absorb particles, called carrier bosons, which
carry momentum between interacting particles. The exchange of moment between sources generates
forces between them. The carrier bosons, which are responsible for gravitational, electromagnetic,
strong and weak forces, are called photon, gluon, weak boson and graviton, respectively. The
PHYSICS 1 2002 S. Nitsolov 22

Table 3.1 Some data about the fundamental forces.

FORCE PROPERTY CARRIER RANGE, RELATIVE


BOSON M STRENGTH

gravitational mass graviton ∞ 10-38


electromagnetic electric charge photon ∞ 10-2
strong colour charge gluon 10-15 1
weak weak charge weak boson 10-18 10-14

fundamental forces can be compared using the following table.


There exists a continuous trend in physics toward the unification of the fundamental interactions.
Recently, Glashow, Weinberg and Salam predicted theoretically (Nobel prize for 1979) and Rubia
Van de Meer confirmed experimentally (Nobel prize for 1984) that the electromagnetic and weak
interactions are just different aspects of the same interaction, called electroweak interaction. Thus,
the number of fundamental interactions was reduced to three: gravitational, electroweak and strong.
Maybe in the near future the electroweak interaction will be unified with the strong interaction.
Many physicists believe that the three fundamental forces will be unified into a single superforce.
PHYSICS 1 2002 S. Nitsolov 23

4 APPLICATION OF NEWTON'S LAWS

4.1 INTRODUCTION
Newton’s second law relates the acceleration and the force and is the base in solving problems in
mechanics. The problems in dynamics can be divided into two basic kinds: in the first kind of
problems we have to determine the motion of a body under the action of known force and in the
second one we find the characteristics of the interaction from measurements of body’s motion. In
this chapter we consider several forces that occur frequently in everyday life. They are gravity,
contact and restoring forces. The gravity is a fundamental force since it is a direct manifestation of
the gravitational interaction, which is one of the three fundamental interactions. The rest forces are
derived. They result from the operation of underlying electromagnetic forces between a large
number of molecules.

4.2 APPLICATIONS OF THE LAWS OF MOTION


The basic problem of dynamics is the determination of the motion for given forces. That means to
determine the position of the body at any instant of time, if the force law is known. For that purpose
it is possible to use Newton's second law in the form F = ma.

4.2.1 MOTION UNDER CONSTANT FORCE


In the case of a constant force the acceleration is also constant. Substituting a = F/m into the
equations for the velocity and displacement of uniformly accelerated motion, 2.33 and 2.33,
respectively, we obtain
F
v = v o − (t − to ) (4.1)
m
F (t − to ) 2
r − ro = v o (t − to ) + (4.2)
m 2
When the directions of the initial velocity and the force are parallel, the motion is rectilinear. When
they are not parallel, the motion will be curvilinear and its trajectory will lie in the plane determined
by these vectors.

4.2.2 RESULTANT FORCE


In writing the Newton's second law we have considered a system of only two interacting particles.
Let us consider a particle m which is subjected to the forces F1, F2, F3,... due to interactions with its
environment. According to the second Newton’s law, the rate of change of momentum produced by
each force Fi when it acts alone is given by
dp i
= Fi (4.3)
dt
When all forces act simultaneously on m, they produce a total rate of change of momentum p
characterised by the resultant force F = dp/dt. Assuming that there is no interference between
different interactions of the particle with other objects (that is, the individual forces on m are not
altered), we may write
F = ∑ Fi (4.4)
i
That is,
when several forces act on a particle simultaneously, their resultant can be found by vector
addition of the individual forces.

This statement constitutes the superposition principle for forces. In terms of rectangular
PHYSICS 1 2002 S. Nitsolov 24
components of the forces the superposition principle is expressed by
Fx = ∑ Fix , Fy = ∑ Fiy , Fz = ∑ Fiz (4.5)
i i i

4.2.3 EQUILIBRIUM OF A PARTICLE


The particle is in equilibrium when its acceleration is zero. Since F = ma, if a = 0, then the net
force must also be zero, F = 0. That is,
a particle is in equilibrium, if the resultant of all forces acting on the particle is zero
or
F = ∑ Fi = 0 (4.6)
i
This equation expresses the condition for equilibrium of a particle.

4.2.4 TANGENTIAL AND NORMAL FORCE


Let us consider a particle moving along a curve. Using the decomposition a = aT u T + aN u N
(Eq.2.30), we can resolve the resultant force acting on particle into two components directed along
the tangent to the trajectory and the principal normal of motion.
F = ma = FT + FN (4.7)
The components of the force FT and FN, called tangential and normal or centripetal force, are
given by
dv mv 2
FT = m u T FN = uN (4.8)
dt R
where uT and uN are unit vectors tangent and normal to the path. The direction of the normal force
always passes through the centre of curvature. If the normal force is zero, then the radius of
curvature is infinite and the trajectory is a straight line. If the tangential force is zero, then the
magnitude of the velocity is constant which results in uniform curvilinear motion.

4.3 GRAVITATION

4.3.1 NEWTON'S LAW OF UNIVERSAL GRAVITATION


The force of gravitation F12 on a particle of mass
m1 exerted by the particle m2 is given by
mm
F12 = − G 1 2 2 u12 (4.9)
r
Figure 4.1 Gravitational forces between two where r is the distance between the particles, u12 is
particles. a unit vector directed from the particle m2 to the
particle m1 and G is the gravitational constant.
This is the statement of Newton's law of universal gravitation for particles. The minus sign
indicates that the force is attractive. According to the third Newton's law, the force F21, acting on
the second particle has the same magnitude and the opposite direction (Figure 4.1).
The value of G obtained from the most recent experiments is 6.67259 × 10-11 N m2 kg-2. It is easy
to see that the numerical value of G is equal to the force of gravitational attraction between two
particles of mass 1 kg separated by distance of 1 m. This force is exceedingly small.
The law of universal gravitation can be directly applied also for uniform spheres. The
gravitational force on a body outside such sphere is equal to that exerted by a particle of the same
mass placed at the centre of the sphere.

4.3.2 THE GRAVITATIONAL FIELD


From the force law 4.9, it follows that if we place a mass m at different positions around another
masses, it experiences a force F. To characterise the effect of gravitation we introduce the quantity
of gravitational field. It is defined as
PHYSICS 1 2002 S. Nitsolov 25
F
g= (4.10)
m
If we compare the definition of gravitational field with the Newton’s second law, F = ma, we see
that the gravitational field equals the acceleration of a particle due to the force of gravity. If the field
is produced by a particle of mass M, the gravitational field is given by
F M
g= =G 2 u (4.11)
m r
where u is a unit vector directed to the particle of mass M.

4.3.3 GRAVITY NEAR THE EARTH'S SURFACE


Consider a body of mass m falling freely at altitude h above the earth's surface The gravitational
forces exerting on the body by other objects are much smaller than the force of gravity due to the
earth and they usually can be neglected. If we neglect also the air resistance, the only force acting
on the body is the gravity. The earth is spherical and can be treated as particle in the gravitational
interaction. Since the dimensions of bodies on the earth surface are small compared with the
distance from the surface to the centre of the earth (where the equivalent point-like mass is placed),
not very large objects can also be treated as particles. Then, the force of gravity on the body of mass
m at altitude h above the earth surface is given by the low 4.9 or
m m
F =G E
(4.12)
( r + h)
E
where mE = 5.97× 10-24 kg and rE = 3.38 × 106 m are the Earth's mass and radius, respectively.
Applying the second Newton's law, F = ma , we obtain
G m G m 1  2h 
F= E
= E
=g ≈ g 1 -  (4.13)
(r + h) 2 r 2 (1 + h/r )2 (1 + h/r ) 2  r 
E E E E
 E 

G mE
where g = is the magnitude of the acceleration due to gravity on the earth's surface. It is
rE2
directed to the Earth’s centre. Since the earth is not exactly spherical, the value of g varies slightly
from place to place. It is seen from Equation 4.13 that g decreases slightly with the altitude. For
example, at altitude of 10 km above the earth’s surface, the acceleration due to gravity is
g = 9.78 m/s2. This value is about 0.3% smaller than it is at see level, g = 9.81 m/s2.

4.4 CONTACT FORCES


Contact forces appear any time when two bodies contact each other. These are the forces that
prevent the bodies from interpenetrating. According to the third Newton’s law these forces are
equal in magnitude and opposite in direction. The contact forces result from atomic repulsion.
When atoms get nearer their electrons repel each other and this repulsion increases strongly as the
distance between atoms decreases. The microscopic mechanism of the contact forces is complex
and incompletely understood.
It is convenient to analyse the contact forces between solid bodies resolving them into two
components, which are parallel and normal to the surface of contact. The normal component is
called the normal force and the component parallel to the surface is called the frictional force.
PHYSICS 1 2002 S. Nitsolov 26
4.4.1 THE NORMAL FORCE
Consider a block of mass m, which is at rest on a
horizontal surface. (Figure 4.2). The block is
subjected only to the force of gravity mg and the
contact force FN exerted by the supporting surface.
Since the block is at rest the net force on it is zero.
Therefore, contact force exerted by the surface on
the block is equal and opposite to its weight.
FN = - mg (4.14)
The contact force is normal because it is directed
perpendicular to the contact surface. Its magnitude
depends on the weight. That is, the normal contact
Figure 4.2 The normal force.
force adjusts itself to oppose the block’s
penetrating into the support.

4.4.2 THE FRICTIONAL FORCE


In Figure 4.3 we show situation in which the block
is pulling by an additional force F directed parallel
to the surface. If the force does not exceed a
certain limit, the block remains at rest. Therefore,
the contact force neutralises not only the weight
but also opposes the sliding motion in direction
parallel to the surface. The horizontal component
of the contact force between static objects is called
the static frictional force
FF = - F (4.15)
If the pulling force reaches a certain critical value
Figure 4.3 The normal and frictional force. FF,S the block suddenly begins to slide. Then to
keep the block moving a smaller pulling force is
sufficient. The force of friction FF,K between objects moving relative to each other is called the
kinetic frictional force.
Experiment shows that the frictional force is proportional to the normal force FN pushing
together the two surfaces.
FF,S = µS FN FF,K = µK FN (4.16)
where the proportionality constants µS and µK are called the static and kinetic coefficients of
friction, respectively.

4.5 RESTORING FORCE


Many objects like springs, rods and strings have the property of elasticity. This is the property of a
body to resume its original size and shape after being subjected to deforming stresses. If an initially
relaxed spring is compressed or extended, it exerts a force on the body, which causes the
deformation so as to restore its relaxed state. This is an example of a restoring force.
For not large deformations the restoring force is proportional to the displacement x of the free
end of the body from its initial position in the relaxed state (x = 0).
F = - kx (4.17)
where the coefficient is called force (or spring) constant. The minus sign indicates that the
restoring force is always opposite to the direction of the displacement. The relationship 4.17 is
known as Hook’s law.
PHYSICS 1 2002 S. Nitsolov 27

5 WORK AND ENERGY

5.1 INTRODUCTION
It is well known, that, when we move a body, we apply a force both to accelerate the body and to
keep its motion overcoming the resistance of the environment. In general, the force changes its
magnitude and direction. In order to describe the result of its action in a finite time interval is useful
to introduce another quantity. The total action of the force on the body during its displacement from
one place to another may be characterised by notion of work. The work is related with another
important quantity, called energy, which obeys a low of conservation analogous to that of
conservation of momentum.

5.2 WORK

5.2.1 RECTILINEAR MOTION UNDER THE ACTION OF THE CONSTANT FORCE

Suppose that a body moves along a straight line


under the action a constant force F (Figure 5.1).
The work done by the force in the displacement
∆r = r - r' from point P to point P’ is defined by
the scalar product
W = F.r (5.1)
Figure 5.1 Rectilinear motion under the action or
of the constant force W = FT s (5.2)
where FT = Fcosθ is the component of the force along the trajectory and s is the displacement along
the path. That is, the work is equal to the product of the displacement and the component of the
force along the displacement.
π
If the angle θ is less than , then cosθ is positive and the work is positive. Also, if θ is larger
2
π
than , then the work is negative. When the force is perpendicular to the displacement, then the
2
work is zero.

5.2.2 CURVILINEAR MOTION UNDER THE ACTION OF A VARIABLE FORCE


In general, the force F is variable and the trajectory C of the particle is a curve line (Figure 5.2). In
this case, the definition of the work may be used for an infinitesimal displacement dr, in which the
change of the force is negligible. Thus, the small amount of work dW done by the force F during the
displacement dr is
dW = F.dr (5.3)
or
dW = FT ds (5.4)
where FT = Fcosθ is the tangent component of the
force and ds is the infinitesimal displacement along
the path.
In order to determine the work done by the force
Figure 5.2 Work in infinitesimal displacement in a finite displacement from A to B, we must divide
it into small displacements and calculate the sum of the correspondent amounts of work.
W = F1 .dr1 + F2 .dr2 + … = FT 1 .ds1 + FT 2 .ds2 + … (5.5)
Since the displacements are infinitely small, the summation may be replaced by the line integral
PHYSICS 1 2002 S. Nitsolov 28
B B
W = ∫ F.dr = ∫ FT .ds (5.6)
A A
This integral must be calculated along the trajectory of motion. For that purpose, we must know the
equation of the trajectory and the function F = F(r). The SI unit of work is called a joule,
J = Nm = kg m s.
In the special case of constant force, the work in the displacement ∆r from point A to point B is
W = F.dr1 + F.dr2 + … = F. ( dr1 + dr2 + …) = F.∆r (5.7)
That is, the work of a constant force is equal to the scalar product of the displacement and the force
and, therefore, does not depend on the trajectory between the beginning and the end of the path.
The work done by restoring force F = - kx (Eq. 4.17) in displacement of the free end of a
deformed body from position x1 to position x2 is expressed by the integral
x2 x2 x2
 x 2  kx 2 kx 2
W = ∫ Fdx = − ∫ kxdx = − k ∫ d   = 1 − 2 (5.8)
x1 x1 x1  2 2 2
We see that the work of the restoring force that obeys Hook's law also depends only on the initial
end final position of the displacement.

5.3 POWER
To express how fast the work is done we use the concept of power. The average power is defined
as the average time rate of the work done W during a time interval ∆t, that is,
W
Pave = (5.9)
∆t
The instantaneous power is the limiting value of the average power when the time interval
approaches zero.
dW
P = lim Pave = (5.10)
∆t → 0 dt
Using the definitions of the work and velocity (Eq. 2.12 and 5.3, respectively) we obtain
dW F.dr dr
P= = = F. = F.v (5.11)
dt dt dt
The SI unit of power is called a watt, W = Js-1 = kg⋅m2⋅s-3. Another unit of work based on watt is
the kilowatt-hour, kWh, equal to the work done by an engine of power of one kilowatt during one
hour. 1 kWh = (1000 W)(3600 s) = 3.6 × 106 J.

5.4 KINETIC ENERGY


Consider a particle, which moves under the action of a force F. The equation, which expresses the
definition of the work, may be transformed as follows
dv ds m  mv 2 
dW = FT ds = maT ds = m ds = m dv = mvdv = d (v ) = d  2
 (5.12)
dt dt 2  2 
The work done by the force during a finite displacement along the trajectory from point A to point
B may be expressed by the equation
B B
 mv 2  1 1
W = ∫ dW = ∫ d 
2 2
 = mvB − mvA (5.13)
A A 
2  2 2
where vA and vB are the magnitudes of the particle's velocity at A and B, respectively. Eq. 5.13 is
known as work-energy theorem. The quantity
1
Ek = mv 2 (5.14)
2
is called the kinetic energy of the particle. The kinetic energy 5.14 can also be expressed in terms
of momentum
PHYSICS 1 2002 S. Nitsolov 29
p2
Ek = (5.15)
2m
By means of Ek the work may be expressed as a difference of the values of the kinetic energy at the
end and beginning of the displacement.
W = Ek(B) − Ek(A) (5.16)
or
W = ∆Ek (5.17)
That is,
the work done by the net force in moving a particle equals the increment in its kinetic energy.
The increment of kinetic energy of a particle depends only on its initial and final state of motion
while the work depends on the values of the force along the path followed. It should be noted that
the value of kinetic energy depends also on the choice of the inertial frame of reference.

5.5 POTENTIAL ENERGY

5.5.1 CONSERVATIVE FORCES


There exists an important class of forces, called
conservative forces. A force is conservative, if the
work done by the force in moving a particle
depends only on the initial and final position of the
particle, regardless of the path followed (Figure
5.3). That is, the amounts of work WAB(1) and
WAB(2) done by a conservative force in moving a
particle from point A to point B along the
Figure 5.3 The work along two different paths. trajectories 1 and(1)2 are equal.
WAB = WAB(2) = WAB (5.18)
The total work W performed passing clockwise the closed path ABA is
W = WAB + WBA (5.19)
Since reversing of the direction of motion is equivalent to reversing the sign of the line integral then
WBA = - WAB and therefore, W = 0. Thus, for conservative forces
W = ∫ F.dr = 0 (5.20)
where ∫ F.dr is line integral, called circulation, which is calculated along a closed path. That is,
the work of a conservative force along any closed path is zero. It can be proved that if the force
satisfies this condition then its work does not depend on the path followed and therefore the two
conditions are equivalent.
Since the work done by a constant force (Eq. 5.7) and restoring force (Eq. 5.8) depends only on
coordinates of the initial and final position we may conclude that these forces are conservative. The
gravitational and electric forces of static charges are also conservative.

5.5.2 POTENTIAL ENERGY


If F is a conservative force, there exists a function of coordinates Ep(r), called potential energy of
the particle, such that
W = Ep(A) - Ep(B) (5.21)
where Ep(A) and Ep(B) are the values of the potential energy at the initial and final points of the path
followed. In order to prove this statement we shall define the potential energy in an arbitrary point A
as
Ep(A) = WAO + c (5.22)
where O is the origin of the reference frame and c is an arbitrary constant. Then the difference
Ep(A) - Ep(A) may be expressed as
Ep(A) - Ep(B) = (WAO + c) - (WBO + c) = WAO - WBO = WAO + WOB (5.23)
The sum WAO + WOB in the right-hand side of equation gives the work WAB done when the particle
PHYSICS 1 2002 S. Nitsolov 30
moves from A to B passing through point O. Therefore,
Ep(B) - Ep(A) = WAB (5.24)
As a result, we obtain the equation, which defines the potential energy.
The change in potential energy is ∆Ep(AB) = Ep(B) - Ep(A). Omitting the subscripts we may write
W = - ∆ Ep (5.25)
That is,
the work done by conservative forces equals the decrement of the potential energy of the
particle.
Comparing this relation with the equation for the work done by a constant force we can define its
potential energy as
Ep(r) = - F.r (5.26)
Potential energy is defined within an arbitrary constant, because if we add a constant to a quantity,
its increment does not change. That is why, the zero or reference level of potential energy may be
chosen whenever it best suits us. For example, the most convenient reference level of potential
energy of falling bodies is zero at the Earth's surface. The zero of the potential energy of a charged
particle is chosen at an infinite distance from the other charges.

5.6 RELATION BETWEEN FORCE AND POTENTIAL ENERGY


The work done by the force F in a small displacement dr of a particle is
dW = F.dr = Fsds = - ∆Ep (5.27)
where Fs is the component of the force in direction of the displacement dr with magnitude ds.
Therefore,
dEp
Fs = − (5.28)
ds
Such derivative with respect to the displacement in given direction is called directional derivative. It
gives the space rate of change of the potential energy in that direction. Similar relations are valid for
each components of the force along the three coordinate axes. Thus,
∂Ep ∂Ep ∂Ep
Fx = − Fy = − Fz = − (5.29)
∂x ∂y ∂z
where the symbol of partial derivative is used, because each derivative is calculated on the
assumption that other two coordinates are constant.
If f is a scalar function of the coordinates x, y, z, a vector which rectangular components are
equal to the directional derivatives along the corresponding coordinate axis is called gradient of the
function f and is designated by the symbol gradf.
∂f ∂f ∂f
gradf = i + j + k (5.30)
∂x ∂y ∂z
Using the definition of gradient we may write
F = −gradEp (5.31)
That is, the potential energy is the negative of the gradient of the potential energy.
Consider a special case when the potential energy depends only on the distance from a fixed
point. The potential energy at an arbitrary point A with a radius vector r with respect to the centre O
may be written as
Ep = Ep(r) (5.32)
In order to calculate the gradient of the potential energy at A we chose the reference frame with
origin at the centre O and X-axes coincident with r. The directional derivatives along Y- and Z-axes
are zero because the potential energy varies only in the direction of r. Therefore the direction of the
force is along X-axes, that is, along the direction of the radius vector, so that the force is central.
Replacing s with r in 5.28, we have
∂Ep
F = Fr = − (5.33)
∂r
PHYSICS 1 2002 S. Nitsolov 31
5.7 LAW OF CONSERVATION OF ENERGY
Assume that only conservative forces act on a particle. Excluding the work from Eq. 5.17 and 5.25,
we obtain
∆Ek + ∆Ep = ∆ ( Ek + Ep ) = 0 (5.34)
the quantity
1
E = Ek + Ep = mv 2 + Ep (5.35)
2
is called the total energy of the particle. Therefore,
the total energy of a particle on which only conservative forces act remains constant with
respect to an inertial frame of reference.
E = Ek + E p = const. (5.36)
This statement is called law of conservation of energy of a particle. In other words, during motion
under conservative forces, both the kinetic and potential energies may vary, but always in such way
that, if one form of energy increases, another one decreases by the same amount. We say that there
is a continuous exchange of kinetic and potential energy. The law of conservation of energy of a
particle is, on one hand, a consequence of the laws of motion; on the other, it is a special case of
more general law of nature - the law of conservation of energy.

5.8 WORK OF NON-CONSERVATIVE FORCES


We know many cases when we can not call the forces acting on a body conservative. For example,
the force of friction depends on the path followed and therefore, it is non-conservative. Usually, a
body is subject to conservative and to non-conservative forces at the same time. The total work
done during the motion under the joint action of the two kinds of forces is
W = −∆Ep + W ' (5.37)
where EP is the potential energy of the conservative forces and W' is the work done by the
non-conservative forces. Combining this equation with the work-energy theorem 5.17, we obtain
∆Ek + ∆Ep = W ' (5.38)
or
∆( Ek + Ep ) = W ' (5.39)
This result gives us a possibility to determine the gain or loss of energy due to the non-conservative
forces.

5.9 POTENTIAL ENERGY CURVES. EQUILIBRIUM


Ep Consider the motion of a particle, whose potential
Ep(x)
energy depends only on its coordinate x. In order
to analyse the motion, we shall use the graph
A
3 representing the potential energy E as a function of
x (Figure 5.4). Such graphs, called potential
2 energy curves, describe clearly the motion of a
B M2 C
body under action of a conservative force.
1 According to Eq. 5.31, the force on the particle is
D G I
M2 given by
M1
x dEp
O F =− (5.40)
Figure 5.4 The curve of potential energy. dx
The derivative dEp/dx is the slope of the curve
Ep(x). The slope is positive whenever the curve is increasing, and negative whenever the curve is
decreasing. Therefore, the force is directed to the left whenever the curve is increasing and directed
to the right whenever the curve is decreasing. At the points M1, M2, M3 where the potential energy is
minimum or maximum the derivative is zero and therefore, the force is also zero, that is, the particle
is in equilibrium. A small displacement from the equilibrium position, where the potential energy is
PHYSICS 1 2002 S. Nitsolov 32
maximum, gives rise to a force, which tends to restore the particle to its initial place. Therefore, the
points (M1 and M3) where the potential energy is minimum are of stable equilibrium. The point
(M2) where the potential energy is maximum is of unstable equilibrium since a small displacement
from the equilibrium position gives rise to a force, which tends to remove the particle from the
initial position.
The total energy E (represented in Figure 5.4 by horizontal lines 1, 2 and 3 for different values of
E) at any position of the particle is a sum of its potential energy given by the ordinate of the curve
and of its kinetic energy Ek = E - Ep given by the distance from curve to the E - line. Since the
kinetic energy is necessarily positive, the particle can not occupy positions in which the total energy
is less than the potential energy. For example, if a particle at energy level 1 is moving to the left, it
could not cross the potential curve at point A since its kinetic energy would be negative in the
region in the left side of A. Therefore, the particle reverses its direction of motion at point A, called
turning point. The region where the potential energy is higher than the total energy of a particle is
called potential barrier.
At lower energy level (horizontal line 2), the motion of the particle is confined to the interval by
the two potential barriers and the particle oscillates between the turning points B and C. Thus, the
motion of the particle is limited to the region BC, called potential well. When the particle has an
energy corresponding to line 3 the motion is allowed in the potential wells DG and HI. The
transition from one potential well to another is forbidden because it would require passing through
the potential barrier GH. The motion in the potential wells is oscillatory between points D and G or
between H and I.
The same analysis may be applied to the more general case of curvilinear motion when the potential
energy depends only on one variable. For example, the motion under the action of central force may
be treated in this way replacing x by r.
PHYSICS 1 2002 S. Nitsolov 33

6 SYSTEMS OF PARTICLES

6.1 INTRODUCTION
In the previous chapters we discussed the dynamics of a single particle representing its interaction
with the rest of universe by the force acting on the particle or by the potential energy. We shall now
investigate the mechanics of many-particle systems. The basic notion for study of systems of
particles is the centre of mass. We shall also discuss and generalise linear and angular momentum
and its conservation in the case of systems of interacting particles. Many processes in the nature
involve exchanges of energy between different bodies by various mechanisms and can be explained
in terms of exchanges of energy between a system of particles and its surroundings. We shall
discuss the work and energy for systems of particles and generalise the principle of conservation of
energy.

6.2 CENTRE OF MASS


Consider a system composed of N particles of
masses m1, m2, …, mN located at points with
position vectors r1, r2, …, rN (Figure 6.1). A
centre of mass (CM) is a point determined by the
position vector
m r + m2 r2 + … + mN rN 1 N
rCM = 1 1
m1 + m2 + … + mN
= ∑ mi ri (6.1)
M i =1
N
where M = ∑ mi is the total mass of the system.
i =1
The rectangular coordinates of the centre of mass
Figure 6.1 Centre of mass of a system of
are
particles
N
1 1
xCM =
M
( m1 x1 + m2 x2 + … + mN xN ) =
M
∑m x
i =1
i i

N
1 1
yCM =
M
( m1 y1 + m2 y2 + … + mN yN ) =
M
∑m y
i =1
i i (6.2)

1 1 N
zCM =
M
( 11 2 2
m z + m z + … + mN N )
z = ∑ mi zi
M i =1
Any ordinary object contains so many particles (atoms and molecules) that we can assume that its
matter has continuous distribution. Then we may replace any particle in Eq. 6.1 with the
correspondent infinitesimal element of mass dm at point r, so that the sums become integrals.
1
M V∫
rCM = rdm M = ∫ dm (6.3)
V

where M is the total mass of the body and integration is taken over the volume V occupied by mass.
For symmetrical uniform body the centre of mass lies at a point, a line or a plane of symmetry. For
example, the centre of mass of a body with rotational symmetry (sphere, cone, cylinder) lies at the
axis of rotation, which is the line of symmetry.

6.3 LINEAR MOMENTUM


The total momentum P of a system of N particles of masses m1, m2, …, mN is defined as
PHYSICS 1 2002 S. Nitsolov 34
N N
P = p1 + p 2 + … + p N = ∑ p i = ∑ mi v i (6.4)
i =1 i =1
The velocity of the centre of mass can be calculated taking the time derivative of its position vector
dr 1 N dr 1 N 1 N
v CM = CM =
dt

M i =1
mi i1 =
dt

M i =1
mi v i = ∑ pi
M i =1
(6.5)

Using the definition of total momentum this result can be rewritten in the form
P = Mv CM (6.6)
That is,
the total momentum of the system of particles is the same as if it were a particle carrying the
whole mass of the system and moving with the velocity of the centre of mass.
For that reason, when we speak of the velocity of a body composed of many particles we
understand the velocity of its centre of mass. We also assume that the measuring instruments are
attached to an inertial frame, called laboratory or L-frame. Sometimes the motion of a system may
be analysed more simply in a reference frame attached to the centre of mass, called a centre-of-
mass frame or C-frame. In C-frame the position vector of centre of mass is always zero and
therefore the total momentum of the system of particles is also zero.

6.3.1 RELATION BETWEEN THE EXTERNAL FORCES AND THE MOMENTUM


The particles in a system may interact among themselves as well with bodies, which do not belong
to the system. The forces exerted on the particles of a system by other particles within the system
and by bodies outside the system are called internal and external forces, respectively. Thus, the
equations of motion for any of the N particles composing the system may be written in the form
dp1
= F1 + F12 + F13 … F1N
dt
dp 2
= F2 + F21 + F23 … F2N
dt (6.7)

dp N
= FN + FN1 + FN2 … FN , N −1
dt
where Fi is the resultant external force acting on mi and Fij is the internal force exerted on particle i
by particle j. Adding the equations of motion for all particles of the system, we obtain
d p 1 dp 2 dp
+ + … N = (F1 + F2 … + FN ) + (F12 + F13 … F1N +
dt dt dt
+ F21 + F23 … F2N + (6.8)

+ FN1 + FN2 … FN , N −1 )
The sum on the left-hand side of Eq. 6.8 can be rewritten in the form
d p 1 dp 2 dp d dP
+ + … N = (p1 + p 2 … + p N ) = (6.9)
dt dt dt dt dt
where P is the total momentum of the system. The first sum in parentheses on the right hand side is
the resultant of the external forces Fext acting on the system
F ext = F1 + F2 … + FN (6.10)
The second sum in parentheses on the right-hand side of the equation is composed of the all internal
forces of the system. For each pair of particles with numbers i and j the corresponding internal
forces Fij and Fji in the sum may be arranged in pairs of the kind Fij + Fji. According to the law of
action and reaction, Fij = - Fji, or Fij + Fji = 0. Therefore, the sum of the internal forces is also zero.
Using this result and Eq. 6.9 and 6.10, Eq. 6.8 can be written in the form
dP
= F ext (6.11)
dt
PHYSICS 1 2002 S. Nitsolov 35
That is,
the time rate of change of the total momentum of a system of particles is equal to the resultant
of the external forces acting on the system.
Combining Eq. 6.6 and 6.11 we obtain
dP d ( Mv CM ) dv
= = M CM = MaCM = F ext (6.12)
dt dt dt
where aCM is the acceleration of the centre of mass. This result is similar to Newton's second law. It
indicates that
the centre of mass moves as if it were a particle of mass equal to the total mass of the system
subjected to the resultant external force.

6.3.2 PRINCIPLE OF CONSERVATION OF MOMENTUM


The relation between the external forces and system’s total momentum, expressed in Eq. 6.11,
dP
shows that if Fext = 0, then = 0 and the total momentum does not change in time. In particular,
dt
the external force is zero for a system of particles upon which no external forces act, or the net
external force on any particle is zero, called isolated system. We can express this as
the total linear momentum of an isolated system is constant.
N
P = ∑ pi = const. (6.13)
i =1
This statement constitutes the principle of conservation of linear momentum. According to this
principle, the interaction between the parts of the isolated system produces only an exchange of
momentum between the bodies within the system but it does not change the total momentum. Since
the total momentum is constant, according to Eq. 6.6, the centre of mass of the closed system moves
with constant velocity relative to any inertial frame.
The principle of conservation of linear momentum is one of the most universal principles of
physics. No exceptions to this principle are known. If there is any kind of interaction, no matter how
complicated, between the particles of a closed system, the total momentum before and after the
interactions act, is the same. The principle of conservation of momentum allows us to solve many
problems without knowing the details of the interactions.

6.4 ANGULAR MOMENTUM


By analogy with the angular momentum of a particle, expressed in Eq. 3.16, the total angular
momentum L of a system of N particles of masses m1, m2, m3, …, mN relative to the reference point
O, fixed to an inertial frame of reference is defined as
N N
L = L1 + L 2 + … + L N = ∑ L i = ∑ ri × mi v i (6.14)
i =1 i =1
The angular momentum L' relative to another reference point O' fixed to the same frame with
position vector R relative to O is given by
N N N
L ′ = ∑ (R + ri ) × mi v i = R × ∑ mi v i + ∑ ri × mi v i = R × P + L (6.15)
i =1 i =1 i =1
If the total linear momentum is zero, L' = L, and we see that the angular momentum is independent
on the choice of the reference point.

6.4.1 RELATION BETWEEN ANGULAR MOMENTUM AND TORQUE


The angular momentum of a single particle L = r × p is related to the torque of the applied force
dL
τ = r × F by the equation = τ (Eq. 3.14). In the case of the system of particles we may apply
dt
this equation to each particle. The result is
PHYSICS 1 2002 S. Nitsolov 36
dL1
= τ1 + τ12 + τ13 … τ1N
dt
dL 2
= τ 2 + τ 21 + τ 23 … τ 2 N
dt (6.16)

dL N
= τ N + τ N 1 + τ N 3 … τ N , N −1
dt
where Li = ri × pi is the angular momentum of the i-th particle, τi = ri × Fi is the torque due to the
external forces on mi and τij = ri × Fij is the torque of the force on mi exerted by mj. All momenta
and torques are computed relative to one and the same arbitrary chosen point O fixed to an inertial
reference frame. Adding all the equations, we obtain
dL1 dL 2 dL N
+ + + = τ1 + τ 2 + τ N + (τ12 + τ13 … + τ1N
dt dt dt
+ τ 21 + τ 23 … + τ 2 N (6.17)

+ τ N 1 + τ N 3 … + τ N , N −1 )
The sum on the right side of the equation is composed of the torques of all internal forces of the
system. For each pair of particles i and j the corresponding torques τij and τji in the sum may be
arranged in pairs τij + τji. Since Fij = - Fji we may write
τij + τji = ri × Fij + rj × Fji = (ri - rj) × Fij = rji × Fij (6.18)
where rji = ri - rj is the vector connecting particles j and i. If we assume that the internal forces act
along the line joining the two particles, then the vectors rji and Fij are parallel and their vector
product is equal to zero. Therefore, the sum of the torques of internal forces is also zero. Denoting
N N
by L = ∑ Li the total angular momentum of the system and τ = ∑ τi the total torque exerted by
i=1 i=1
the external forces we have
dL
= τ ext (6.19)
dt
That is,
the time rate of change of the total angular momentum of a system of particles is equal to the
sum of the torques of the external forces acting on the system.
This is the fundamental law of the dynamics of rotation. Eq. 6.19 has meaning only if the vectors
L and τext are calculated relative to one and the same reference point fixed to an inertial frame. It
can be shown that the equation remains valid if the reference point is at the centre of mass of the
system of particles even when the centre of mass accelerates with respect to an inertial frame.
By the same reasoning as in Eq. 6.15, we obtain that the torque of the external forces τ' relative
to another reference point O' fixed to the same frame with position vector R' = OO' relative to O is
given by
N N N
τ ′ = ∑ (ri − R ) × Fi = ∑ ri × Fi − R × ∑ Fi = τ − R × F (6.20)
i =1 i =1 i =1

If the resultant force is zero, τ' = τ, that is, the resultant torque is independent on the choice of the
reference point. An important example is the case in which the external forces can be reduced to
two forces F1 and F2 of equal magnitude and opposite directions, F1 = - F2, having lines of action a
distance l apart. Such a pair of forces is called a couple. We choose the reference point on the line
of action of F1. Then the torque of F1 is zero and the torque of the couple is equal to the torque of
the second force, or τ = l × F2. The vector of the torque is directed perpendicular to the plane
formed by the lines of the forces and its direction is determined by the right-hand rule. The
magnitude of the torque is equal to τ = lF.
PHYSICS 1 2002 S. Nitsolov 37
6.4.2 LAW OF CONSERVATION OF ANGULAR MOMENTUM
In the special case when the torque of the external forces is zero, Eq. 6.19 gives
L = const. (6.21)
Since the external forces on the particles of an isolated system are zero we can state that,
the total angular momentum of an isolated system of particles is constant.
This statement is called the law of conservation of angular momentum for a system of particles.
It is a profound law of nature like the principle of conservation of linear momentum. According to
the principle, if one part of a closed system changes its angular momentum, the rest part must
experience the opposite change in its angular momentum in order to conserve the total angular
momentum. For example, the orientation of a spacecraft can be changed by the use of rigidly
mounted flywheel. The spacecraft is an isolated system and its total angular momentum is always
zero. If the flywheel is set in rotational motion, the spacecraft will start to rotate in the opposite
sense to maintain the zero value of the total angular momentum of the system. When the flywheel
stops, the spacecraft will also stop, but it will have different orientation. In the same way, a cat can
control its orientation in space rotating rapidly its tail instead of flywheel.

6.4.3 ORBITAL AND INTERNAL ANGULAR MOMENTUM


We consider the angular moment of the system relative to a point fixed to L-frame. Using the
decomposition
ri = rCM + ri '
(6.22)
v i = v CM + v i '
where ri and vi and ri' and vi' are the position vector and velocity of mi relative to L- frame and to
C-frame, respectively, the angular moment can be written in the form
N N
L = ∑ ri × mi v i =∑ (rCM + ri ') × mi ( v CM + v i ') =
i =1 i =1
N N
 N  N
= rCM × v CM ∑ mi + rCM × ∑ mi v i ' +  ∑ mi ri '  × v CM + ∑ ri '× mi v i ' = (6.23)
i =1 i =1  i =1  i =1
N
= rCM × Mv CM + rCM × P '+ r 'CM × Mv CM + ∑ ri '× mi v i '
i =1
Since the total momentum of a system of particles and the position vector of centre of mass are
always zero in C-frame, the second and the third terms at the right-hand side are zero. Then we
obtain the decomposition
L = Lorb + Lint (6.24)
where
Lorb = rCM × MvCM (6.25)
N
L int = ∑ ri '×mi v i '
i =1 (6.26)
The first term, Lorb, called orbital angular momentum is the angular momentum of a particle of
mass M (equal to the total mass of the system) located at the centre of mass. The second term, Lint,
called internal angular momentum, or spin (for elementary particles) is the total angular
momentum computed relative to the centre of mass.
Similarly, the total torque is the sum of two terms:
N N N N N
τ ext = ∑ ri × Fi =∑ (rCM + ri ') × Fi = rCM × ∑ Fi + ∑ ri '× Fi = rCM × F + ∑ ri '× Fi (6.27)
i =1 i =1 i =1 i =1 i =1
The first term is the torque of the resultant external force applied to the centre of mass. The second
N
term τ CM = ∑ ri ' × Fi is the resultant external torque relative to the centre of mass.
i=1

dL
In addition to Eq. 6.19, = τ ext we can find similar equations between the correspondent terms
dt
PHYSICS 1 2002 S. Nitsolov 38
of total angular momentum and torque. Taking the time derivative of the orbital angular momentum
we have
dL orb drCM dv
= × Mv CM + rCM × M CM = v CM × Mv CM + rCM × F = rCM × F (6.28)
dt dt dt
Substracting Eq. 6.19 and Eq. 6.28 we obtain
dLint
= τ CM (6.29)
dt
This equation is valid even if the centre of mass accelerates with respect to an inertial frame of
reference.
These results show that the motion of the system is a superposition of the motion of the centre of
mass itself and the motion of its components around the centre of mass. For example, the angular
momentum due to the rotation of Earth about its North-South axis is given by the internal angular
momentum. The Earth’s angular momentum due to its orbiting is its orbital angular momentum.
The total Earth’s angular momentum relative to the Sun is equal to the sum of these two terms.

6.5 KINETIC ENERGY


We consider a system composed of N particles of masses m1, m2, … mN moving with velocities v1,
v 2, … v N. The system has total kinetic energy Wk, which is defined to be the sum of the individual
particle kinetic energies. Thus,
N N
m v2
Ek = ∑ Ek ,i =∑ i i (6.30)
i =1 i =1 2
The kinetic energy of the system also depends on the choice of the frame of reference. In order to
calculate the kinetic energy relative to L-frame we use the relation vi = vCM + vi' (Eq. 6.20) between
velocities vi and vi' of the particle i referred to L - frame and C - frame. We have
N
m v2 N
m ( v + v 'i ) 2
Ek = ∑ i i = ∑ i CM =
i =1 2 i =1 2
2
N
mi vCM N N
m v '2
=∑ + ∑ mi ( v CM .v 'i ) + ∑ i i = (6.31)
i =1 2 i =1 i =1 2
2
MvCM N N
m v '2
= + ( vCM .∑ mi v 'i ) + ∑ i i
2 i =1 i =1 2
N
Since the total momentum ∑ m v ' of the system relative to C - frame is zero, we may write
i =1
i i

Ek = Ek,orb + Ek,int (6.32)


where
2
MvCM
Ek,orb = (6.33)
2
and
mi v 'i2N
Ek,int = ∑ (6.34)
i =1 2
are called orbital and internal kinetic energy of the system. The first term is the kinetic energy of
a particle having the total mass of the system and travelling with the velocity of the centre of mass
relative to an inertial reference frame. The second term is the kinetic energy of the particle’s motion
relative to the centre of mass, which is independent on the choice of reference frame. That is,
the kinetic energy of the system is a sum of two terms associated with the motion of the centre
of mass referred to an inertial reference frame and the motion of the particles relative to the
centre of mass.
For example, the kinetic energy of the system Earth-Moon is composed of the kinetic energy
associated with its orbiting around the sun and the kinetic energy of the motion of the two planets
relative to their common centre of mass. Similarly, the kinetic energy of a gas molecule is a sum of
PHYSICS 1 2002 S. Nitsolov 39
the kinetic energy due to the motion of its centre of mass and the kinetic energy of the internal
rotational and vibrational motion of its atoms referred to the centre of mass.

6.6 WORK- ENERGY THEOREM


The particles of a system may interact among themselves as well with bodies, which do not belong
to the system. The work Wi done by the forces on i–th particle can be separated into two parts: the
work Wi,int done by the internal forces and the work Wi,ext done by the external forces, that is,
Wi = Wi,int + Wi,ext (6.35)
Using the work-energy theorem for the particle i, we have
∆Ek,i = Wi,int + Wi,ext (6.36)
Adding ∆Ek,i for each particle in the system gives
N N N N

∑ ∆Ek ,i = ∆∑ Ek ,i = ∆Ek = ∑ Wint,i + ∑ Wext ,i


i =1 i =1 i =1 i =1
(6.37)

Using Eq. 6.32 we have


∆Ek,int + ∆Ek,orb = Wint + Wext (6.38)
N N
where Wint = ∑ Wint,i and Wext = ∑ Wext ,i are the total internal and external work, respectively. Eq.
i =1 i =1
6.38 is called the work-energy theorem for a system of particles.
If the internal forces applied on the particles of a system are conservative, we may define the
internal potential energy of the system as a sum of the potential energies of all the particles
associated with the internal forces.
N N
∆Ep,int = ∑ ∆Ep,int,i = −∑ Wint,i = − Wint (6.39)
i =1 i =1
Substituting Wint into Eq.6.38 gives
∆( Ep, int + Ek,int + Ek,orb ) = Wext (6.40)
The quantity
Eint = Ep,int + Ek,int (6.41)
is called internal energy of the system. The internal kinetic energy is independent on the frame of
reference. In general, the potential energy depends on the positions of the particles of the system
relative to other bodies. If the internal forces are central, the potential energy of the interaction
between a pair of particles i and j depends only on the distance rij separating the particles. The total
internal potential energy of the system consists of the potential energies of all pairs of particles and
therefore depends only on the distances between particles. Thus we see that the internal energy does
not depend on the choice of reference frame. In terms of internal energy Eq. 6.40 may be written as
U ≡ Eint (6.42)
or
the change of the sum of internal energy and orbital kinetic energy of the system of particles is
equal to the work of the external forces.
If the resultant of the external forces is zero, then vCM = const and ∆Eint = Wext. In this case the
external forces can change only the internal energy of the system. If the internal energy is constant,
 Mv 2 
then ∆  CM  = Wext . That is, the work is used only to change the orbital kinetic energy.
 2 

6.7 PRINCIPLE OF CONSERVATION OF ENERGY


If both internal and external forces are conservative, Eq. 6.40 may be written in the form
∆Ek + ∆Ep,int + ∆Ep,ext = 0 (6.43)
or
Ek + Ep = const. (6.44)
where Ep = Ep,int + Ep,ext is the potential energy of the system. The quantity E = Ek + Ep is called the
PHYSICS 1 2002 S. Nitsolov 40
total mechanical energy of the system. We see that
in the absence of nonconservative forces, the total mechanical energy of a system is constant
with respect to an inertial reference frame.
This statement is called principle of conservation of mechanical energy. In the special case
when the system of particles is isolated we can state that,
if the internal forces are conservative, the total mechanical energy of an isolated system is
constant with respect to an inertial reference frame.
This is another more restrictive form of the principle of conservation of mechanical energy. It
was derived as a theorem from Newton's laws of motion by assuming that all internal forces are
conservative. Nevertheless, the principle of conservation of energy has general validity for all the
phenomena in the nature, if terms represented all possible kinds of energy of the system like
chemical, electromagnetic, nuclear etc. are included into the total energy. Then it can be stated as
follows:
the total energy of a closed system is constant with respect to an inertial reference frame.
That is, the energy can be changed from one form to another, but cannot be created or destroyed.
The conservation of energy is one of the most fundamental principles of nature. To our knowledge,
no violation of this principle has been ever observed.

6.8 COLLISIONS
Any encounter between two or more bodies in which the interaction is relatively strong and acts for
short time may be considered to be a collision. The objects that collide might be elementary
particles, atoms and molecules in a gas, billiard balls, cars etc. The forces of such short interactions
are so great that the role of the rest of the forces is negligible. Thus, we may consider the colliding
bodies as an isolated system, which obeys the principles of conservation of momentum, angular
momentum and energy. The law of conservation of energy requires that the sum of the potential and
kinetic energy of the system remains constant. If the kinetic energy of the system is also conserved,
the collision is said to be elastic. Otherwise, if the kinetic energy changes, the collision is inelastic.
Collisions of macroscopic bodies always involve certain transfer of kinetic energy into other forms
of energy. However, in many collisions the lost kinetic energy is negligibly small compared to the
initial kinetic energy so that they can be treated as elastic. For example, collisions between steel
balls can be considered as elastic. If the two bodies stick together after the collision, the change of
kinetic energy is maximum and the collision is said to be completely inelastic. This is the case of
head-on collision of two cars.

6.8.1 ELASTIC COLLISION


Consider the collision of two elastic balls A and B
(Figure 5.2). Suppose we are given the masses mA
and mB and the initial velocities vA1 and vB1 of the
colliding bodies and we wish to compute their
velocities vA2 and vB2 after the collision. We shall
Figure 6.2 Central elastic collision of the balls neglect the rotational motion of the balls and shall
A and B assume that the collision is head-on or central,
that is, before the collision the velocities of the
balls are directed along the line joining their centres. The contact forces during the impact and the
changes in momenta they produce are directed along the same line. Thus, we see that the velocities
after a central elastic collision are also directed along the line joining the centres of the balls.
It is convenient to perform the computations in the frame associated with the centre of mass of
the system of colliding bodies. The velocities in the C-frame are v'A1 = vA1 - vCM; v'B1 = vB1 - vCM;
m v + mB v B1
v'A2 = vA2 - vCM; v'B2 = vB2 - vCM where v CM = A A1 (Eq. 6.5). Then the laws of
mA + mB
conservation of momentum and energy can be written in the form
mA v ′A1 + mB v ′B1 = mA v ′A 2 + mB v ′B 2 (6.45)
PHYSICS 1 2002 S. Nitsolov 41
mA v ′A21 mB vB′21 mA vA′22 mB vB′22
+ = + (6.46)
2 2 2 2
Taking into account that in the C-frame the total momentum is always zero and all the velocities are
directed along the line joining the centres of the balls, the conservation of momentum can be
expressed in scalar form as
mA v A′ 1 + mB vB′ 1 = mA vA′ 2 + mB vB′ 2 = 0 (6.47)
It is easy to see that the system of Eq. 6.46 and 6.47 has two solutions: v'A2 = v'A1; v'B2 = v'B1 and
v'A2 = - v'A1; v'B2 = - v'B1. Only the second solution has physical sense because the first solution does
not obey the conditions of the problem (both velocities remain unchanged after the collision). Thus
we see that in the C-frame the velocities after the collision change their signs but remain unaltered
in magnitude. The same is valid for the relative velocities of the balls, v'A2 - v'B2 = -( v'A1 - v'B1). In
the L-frame, the velocities after the collision are
(m − mB ) v A1 + 2mB v B1 )
v A2 = v ′A2 + v CM = − v ′A1 + v CM = − v A1 + 2 v CM = A
mA + mB
(6.48)
(mB − mA ) v B1 + 2mA v A1 )
v B 2 = v ′B 2 + v CM = − v ′B1 + v CM = − v B1 + 2 v CM =
mA + mB
If the balls have equal mass, then Eq. 6.48 reduces to vA2 = vB1 and vB2 = vA1, that is, the balls
exchange their velocities in the collision. In particular, if one of the balls was initially at rest, it
moves with the velocity of the second ball and the moving initially ball comes to rest. In the case of
collision of a ball with massive object like wall, mA << mB, and we have vA2 = - vA1 + 2vB1 and
vB2 = vB1. Therefore
∆vA = 2(vB1 - vA1) ∆vB = 0 (6.49)
which indicated that the change in velocity of the ball is equal to the its reversed doubled initial
relative velocity while the velocity of the massive object remains unchanged.
In the case of non-central elastic collision one has to consider both normal and tangential
components of the velocities. The normal components of the velocities obey Eq. 6.46 and 6.47 and
their solution 6.48. The tangential components of the velocities can be changed only by friction
during the impact of the balls. But if there were friction, the mechanical energy would change.
Therefore, the tangential components of the velocities remain unaltered.

6.8.2 COMPLETELY INELASTIC COLLISION


For the case of completely inelastic collision of two bodies A and B, the calculation of the final
velocities is simplified by the fact that by definition the bodies move as a unit after the collision and
therefore their common velocity is equal to the velocity of the centre of mass of the system.
m v + mB v B1
v A2 = v B 2 = v CM = A A1 (6.50)
mA + mB
The change in kinetic energy in the collision is equal to
(mA + mB )vCM2
mA v A21 mB vB21 mA mB
∆Ek = Ek 2 − Ek 1 = − − =− ( v A1 − v B1 ) 2 (6.51)
2 2 2 2(mA + mB )
The right side is necessarily negative, so the total kinetic energy decreases in the collision.
PHYSICS 1 2002 S. Nitsolov 42

7 MOTION OF RIGID BODY

7.1 INTRODUCTION
To give an example of the application of the mechanics of many-particle systems presented in the
previous chapter, we consider the motion of rigid body. We shall focus our attention mostly on
rotational motion and we shall consider simple situations, which illustrate the most important
properties of rotation of rigid bodies. We shall also demonstrate the similarities between rotational
and translational motion.

7.2 TRANSLATIONAL AND ROTATIONAL MOTION


A system of many particles is called rigid body when the distances between the particles of the
system do not change. Therefore, a rigid body conserves its shape under the action of external
forces. Absolutely rigid body does not exist since any real body deforms when a force acts on it.
However, the deformations of a body are often negligible, so that the concept of a rigid body is a
good approximation in many cases.
Any motion of a rigid body can be resolved into translation and rotation. Translational motion
is defined as a motion in which all the particles have the same velocity and therefore, they describe
parallel paths so that any straight line of the body remains parallel to its initial position. In rotation,
all the particles describe circular paths whose centres are located on a straight line called axis of
rotation. In translation, all the particles of a rigid body have the same velocity and acceleration. It is
therefore, sufficient to determine the motion of an arbitrary point of a body to describe the motion
of the entire body. For example, the motion of the centre of mass may be determined using Eq. 6.6.

7.3 ROTATION OF A RIGID BODY ABOUT A FIXED AXIS


We next discuss the special case of rotation about a fixed axis. Figure 7.1 represents a rigid body of
arbitrary shape mounted on an axle supported in fixed bearings O' and O''. As the body rotates its
particles move in circles centreed on the Z-axis. The angular moment Li of a small element of the
body of mass mi and position vector ri relative to the origin O is
Li = ri × mivi (7.1)
The axial component of Li is given by Eq. 3.20:
Liz = miRi2ω (7.2)
where ω is the angular velocity of rotation and Ri
is the perpendicular distance from the axis of
rotation to the mass element i. The axial
component of the total angular momentum
 2
Lz = ∑ Liz =∑ mi Ri ω =  ∑ mi Ri  ω = I ω (7.3)
2

i i  i 
where the quantity
N
Figure 7.1 Rotation of rigid body about a fixed I = m1 R12 + m2 R22 + … + mN RN2 = ∑ mi Ri2 (7.4)
axis i =1
is called moment of inertia or rotational inertia
of the body relative to the axis of rotation Z. For a continuous distribution of matter the summation
is replaced by the integration over the volume V occupied by mass.
I = ∫ R 2 dm (7.5)
V
PHYSICS 1 2002 S. Nitsolov 43
7.3.1 MOMENT OF INERTIA
The moment of inertia is proportional to the total mass of the object and depends strongly upon the
distribution of the mass perpendicular to the axis of rotation. The larger moment of inertia, the more
difficult it will be to start the object rotating or stop it from rotation. The moment of inertia is a
measure of inertia of the object like the mass. However, I depends on the location of the axis of
rotation while the mass is an intrinsic property of the body. Moments of inertia relative to axes of
symmetry for various common bodies are given in Table 7.1.
Table 7.1 Moments of inertia of some symmetric bodies

Hollow sphere Solid parallelepiped


2  2 R12 R22  1

I = M  R1 + R2 − 2
2
 I = M (a 2 + b 2 )
5  R1 + R1R2 + R22  12

Hollow cylinder Long thin rod


1 1
I = M ( R12 + R22 ) I = ML2
2 12

All of the moments of inertia in the table are calculated for an axis passing through the centre of
mass of the object. Calculating the moment of inertia about any other axis can be performed by use
of Steiner’s theorem, known also as parallel axis theorem.
An object with moment of inertia I about an axis
Z' a Z of rotation passing trough the centre of mass is
shown in Figure 7.2. We choose the reference
frame with origin at the centre of mass and Z-axis
CM coincident with the axis of rotation. Consider the
moment of inertia I’ for the body about the axis Z’
parallel to Z and a distance a away. By definition
I' I (Eq. 7.4),
N N
I ' = ∑ mi Ri2 = ∑ mi ( a + R i ' )
2
Figure 7.2 Moments of inertia about two (7.6)
parallel axes i =1 i =1
where Ri' is the position vector of the i-th mass
element relative the centre of mass. Expanding and rearranging, we have
N
 N  N   1 N 
I ' = a 2 ∑ mi + 2a.  ∑ mi R i '  + ∑ mi R 'i2 = Ma 2 + 2M a.  ∑ mi R i '   + I ' (7.7)
i =1  i =1  i =1   M i =1 
The term in parenthesis is equal to the projection of the position vector of the centre of mass on XY
- plane and is zero since the centre of mass is coincident with the origin. Thus, we have
I’ = Ma2 + I (7.8)
This equation is called Steiner’s theorem. The equations in Table 7.1 can be extended by use of the
Steiner’s theorem.

7.3.2 EQUATION OF ROTATIONAL MOTION


In general, the angular momentum of a rigid body has not the same direction as the angular
velocity. However, it may be shown that for each body, there exist three mutually perpendicular
directions, called principal axes of inertia, for which the angular moment is parallel to the axis of
rotation. That is, if the rotation is about one of the principal axes
L = Iω (7.9)
The corresponding values of the inertial moment I1, I2 and I3 are called principal moments of
inertia. For symmetrical uniform body the principal axes coincide with axes of symmetry.
Consider a rotation of a rigid body about its principal axis. Substituting L = Iω into Eq. 6.19, we
PHYSICS 1 2002 S. Nitsolov 44
obtain
dL d dω
= ( I ω) = I = τ ext (7.10)
dt dt dt
Omitting the subscript, we have
Iα = τ (7.11)
That is, for a rigid body rotating about a principal axis, the time rate of the change of its angular
velocity is proportional to the resultant torque acting on the body. This equation is analogous to
Newton's second law. We see that the moment of inertia plays the same role for rotational motion
that mass does for translational motion.
When the torque is zero, it follows from the equation that the angular velocity is also constant.
This statement is analogous to the law of the inertia. It should be noted that if the rotation is about
axis, which does not coincide, with any principal axis and the torque is zero, the angular velocity is
not constant since Eq. 7.9 is not valid in this case.

7.3.3 EQUILIBRIUM
Since the motion of a rigid body is a combination of a translation and rotation, there are two
equilibrium conditions, F = 0 and τ = 0. The condition for translational equilibrium is
∑i Fi = 0
(7.12)
∑ ix
i
F = 0 ∑ iy
F
i
= 0 ∑ iz
F
i
= 0

that is, the sum of all the forces must be zero and therefore, the acceleration of the centre of mass is
also zero.
In the case of a rotational motion the sum of all the torques relative to any point must be zero.
∑i τi = 0
(7.13)
∑ ix
τ
i
= 0 ∑ iy
τ
i
= 0 ∑ iz
τ
i
= 0

This is the condition for rotational equilibrium. It was shown in the previous chapter (Eq. 6.20)
that, if the net force acting on the system is zero, the choice of the origin taken for the computation
of torques is of no importance.

7.4 ENERGY OF A RIGID BODY


Consider a rigid body, which translates as well as rotates about an axis passing through the centre of
mass of the body. The kinetic energy of the body obeys the general relation 6.32. According to the
definition of a rigid body, its particles do not move relative to each other and therefore, the rotation
is the only possible motion of the particles of a rigid body relative to the centre of mass. Thus, Eq.
6.32 can be rewritten in the form
2
MvCM
Ek = + Ek,rot (7.14)
2
Since the velocity vi' of the i-th particle with respect to the centre of mass is given by vi' = ωRi
where is the angular velocity and Ri is the distance from the axis of rotation, its rotational energy
is
 N 
N
m v′ 2 N
m Rω 2  ∑ mi Ri2  ω 2
Iω 2
Ek,rot = ∑ i i = ∑ i i i =  i =1  = (7.15)
i =1 2 i =1 2 2 2

where I is the moment of inertia of the body. Thus, we obtain


MvCM2
Iω 2
Ek = + (7.16)
2 2
that is, the kinetic energy of a rigid body is a sum of kinetic energies of its translational and
PHYSICS 1 2002 S. Nitsolov 45
rotational motion. If the axis of rotation coincide with any principal axis, then L = Iω, and
I ω 2 L2
Ek,rot = = (7.17)
2 2I
The potential energy of the internal forces does not change during the motion since the distances
between the particles of a rigid body are constant. Since the potential energy is determined within a
constant, we may define it to be zero, Ep,int = 0. If the external forces are conservative, the total
energy of a body is expressed by the following equation:
2
MvCM Iω 2
E = Ek + Ep,int + Ep,ext = + + Ep,ext (7.18)
2 2
For example, the total energy of a bullet shot by a rifled gun is composed of the kinetic energy of
its translational motion and its rotation along its axis and its potential energy in earth’s gravitational
field.

7.5 ANALOGY BETWEEN THE ROTATIONAL AND


TRANSLATIONAL MOTION
To describe rotational motion, we introduced quantities like angular momentum and torque and
stated relations between them that corresponded to similar quantities and equations for linear
motion. The similarities between translational and rotational motion are summarized in Table 7.2.

Table 7.2 Analogous quantities and relations.

Linear motion Rotational motion


infinitesimal displacement, dr infinitesimal angle of rotation, dθ
dr dθ
velocity, v = angular velocity, ω =
dt dt
dv dω
acceleration, a = angular acceleration, α =
dt dt
mass, m moment of inertia, I
linear momentum, p angular momentum, L
force, F torque, τ
p = mv Lz = I ω
dp dL
=F =τ
dt dt
ma = F Iα = τ
mv 2 Iω2
Ek = Ek =
2 2
PHYSICS 1 2002 S. Nitsolov 46

8 IDEAL GAS

8.1 INTRODUCTION
It is well known that any body is composed of an enormous number of particles (atoms and
molecules). The kinetic theory of matter states that these particles are in continuos random
motion. Generally speaking, the laws of mechanics are also valid for such many-particle systems.
However, when the number of particles increases, the mathematical difficulties in determination of
motion of individual particles sharply increase and for large numbers of particles the calculations
become unrealisable. In addition, usually we are interested in the behaviour of the system as a
whole rather then in the motion of the each individual molecule. Statistical physics relates the
collective or macroscopic properties of a system to its microscopic structure without considering the
detailed motion of each particle. It is interested not in the behaviour of the individual molecule, but
only in average quantities characterising the motion of a very large number of particles.
Thermodynamics concerns macroscopic processes in which energy is transferred as heat or work
without explicitly considering the microscopic structure of matter. Thermodinamical and statistical
approaches mutually supplement each other. To demonstrate this we consider the simplest kinetic
theory of dilute gases.

8.2 THERMAL EQUILIBRIUM AND TEMPERATURE

8.2.1 THERMAL EQUILIBRIUM


In thermodynamics a system is a particular amount of matter (composed of enough large number of
atoms and molecules) under consideration. The state of a particular system is identified by
characteristic macroscopic properties called variables of state. For example, the state of a gas in
cylinder with piston is determined by its pressure p, volume V, temperature T and internal energy U
(in thermodynamics the internal energy is usually denoted by U instead of Eint). A system can be
isolated from the rest of the world by a thermally nonconducting boundary. It is a fact of
experience, that an isolated system after a sufficiently long time approaches state, in which all
variables of state are constant, called thermal equilibrium. In a non-equilibrium state, at least one
parameter has not a definite value. For example, when a gas is rapidly compressed in cylinder with
piston, the pressure of the gas near the piston will be higher than elsewhere and the value of the gas
pressure in the cylinder is not determined. If after the deviation from equilibrium the system is left
again to itself, the equilibrium will be restored. The transition from a non-equilibrium state to
equilibrium is called relaxation and the time required for such a transition is called relaxation
time. If system A in thermal equilibrium is put into intimate contact with a second system (system
B) also in equilibrium, the both systems achieve a new equilibrium state and we say that the two
systems are in thermal equilibrium with each other. Two systems can be in state of thermal
equilibrium even if they are thermally isolated from each other. To determine if the systems are in
equilibrium with each other, the so-called zeroth law of thermodynamics can be used. It states that
if two bodies are separately in thermal equilibrium with another body, then they are in
thermal equilibrium with each other.
Thermodynamics studies systems at (or near) equilibrium. At equilibrium all parameters of the
system have definite values and the number of independent parameters of state is reduced. For
example, a given amount of uniform fluid in equilibrium has only two independent parameters. If
we choose two parameters, the remaining parameters of state are related to the values of the two
parameters. Such a relation is called an equation of state. Any change of state is called a process.
The process is identified by the sequence of states through which the system passes. If all these
states are equilibrium, the process is called quasi-static or equilibrium process. In general, each
process is associated with some disturbance of equilibrium. However, the condition of an
PHYSICS 1 2002 S. Nitsolov 47
equilibrium process can be satisfied, if the series of states are be traversed enough slow (the
duration of the process is much more than the relaxation time), allowing the system to come to
equilibrium after each small change. Thus, during the equilibrium process, the parameters of state
are well defined and the deviations from their equilibrium values are negligible.

8.2.2 TEMPERATURE
In everyday life, temperature refers to how hot or cold a body is. It can be defined more precisely
by means of the notion of thermal equilibrium. Temperature is a characteristic of the state of
thermal equilibrium. All systems in thermal equilibrium with each other have the same temperature.
Systems with different temperatures are not in equilibrium. The temperature of a system increases,
if it receives energy at constant external conditions. The instruments used to measure the
temperature are called thermometers. They determine the temperature from some property of the
thermometric substance that depends on temperature. For example, the temperature of a fixed
amount of gas in equilibrium can be determined measuring its volume and pressure. To measure the
temperature of a body we place the thermometer in contact with the body until the whole system of
two bodies in contact approaches thermal equilibrium and determine the temperature by measuring
the parameters of the working substance of the thermometer.

8.2.3 TEMPERATURE SCALES


The Celsius temperature scale uses the unit of temperature, called degree Celsius, C°. The scale
is based on the assumption that the relation between the property X used in the measurement, and
the temperature t is linear.
t = aX + b (8.1)
To determine the constants a and b we have to measure X for two fixed points on the temperature
scale. The temperature of melting ice is chosen to be the first point with numerical value 0 C°. The
second point (value 100 C°) is the temperature of boiling water.
The thermodynamic temperature scale, called also absolute or Kelvin, is the standard SI
temperature scale. The unit of thermodynamic temperature, called kelvin, K, represents exactly the
same temperature interval as the degree Celsius. The thermodinamical scale is also based on the
linear relation between the property X used in measurement and the temperature T, where b = 0,
T = aX (8.2)
To eliminate constant a we need only one fixed point. It is the unique temperature, called tripled
point, at which saturated vapour, pure water and melting ice are in equilibrium (at a pressure of 610
Pa). Its value is defined to be 273.16 K. The values of temperature when measured in the Kelvin
and in the Celsius scales are related by
T = t + 273.15 (8.3)
Some English-speaking countries use the Fahrenheit scale. The Fahrenheit and Celsius
temperatures are related by
9
TF = 32 + t (8.4)
5
Since the thermometers uses different substances they agree only at the fixed points and need to be
calibrated against a standard thermometer.

8.3 THE IDEAL GAS EQUATION

8.3.1 ISOTHERMAL PROCESS


The behaviour of dilute gases at equilibrium is governed by relatively simple empirical laws. The
first of the empirical gas laws was discovered by the Anglo-Irish scientist Robert Boyle in 1662 and
by the French physicist Edme Mariotte in 1676 independently. They recognised that the volume of a
given quantity m of gas is inversely proportional to the pressure p at constant temperature T.
pV = const. (m, T fixed) (8.5)
This relation is known as Boyle's law. An equilibrium process during which the temperature does
PHYSICS 1 2002 S. Nitsolov 48
p not vary is called isothermal process. It can be
represented on a p - V diagram. (Figure 8.1). A set
of states with the same temperature forms a curve
on the diagram called isotherm.

T2 T1<T2 8.3.2 ISOBARIC PROCESS


T1 French scientist Jacques Charles demonstrated
experimentally in 1787 that the volume of fixed
0 V amount m of gas is directly proportional to the
Figure 8.1 Isothermal process. thermodynamic temperature T when the pressure p
is kept constant.
V
= const. (m, p fixed) (8.6)
T
This equation is known as Charles's law. A process during which the pressure is kept constant is
called isobaric process.

8.3.3 ISOCHORIC PROCESS


The third empirical gas law is Gay-Lussac's law. It is named in honour of the French scientist
Joseph-Louis Gay-Lussac. In 1802, he found experimentally that for fixed amount m of gas at
constant volume, the pressure p of the gas is directly proportional to the thermodynamic
temperature T, that is
p
= const. (m, V fixed) (8.7)
T
A process during which the volume does not change is called isochoric process.

8.3.4 THE IDEAL GAS EQUATION


The three experimental gas laws can be summarised in one universal relation between the pressure,
volume, temperature and amount of gas. To do it let us consider a transition of the gas from an
initial state in which values of variables of state are p1, V1 and T1 to final state, determined by the set
of values p2, V2 and T2. The transition may be realised in two steps. The first step is an isochoric
process from the initial state to an intermediate state, determined by p2, V1 and T’. The second step
is an isobaric process from this intermediate state to the final state. According to the Gay-Lussac's
law 8.7, we may write for the isochoric transition to the intermediate state
p1 p2
= (8.8)
T1 T '
Applying the Charles's law 8.6 for the transition to the final state, we have
V1 V2
= (8.9)
T ' T2
Combining the two equations we obtain
p1V1 p2V2
= (8.10)
T1 T2
or
pV
= const. (m fixed) (8.11)
T
where the constant depends on the amount of gas. For gases it is useful to use the quantity of
amount of substance the unit of which is mole (abbreviated mol) instead of mass. According to the
International System (SI) definition, one mole is the amount of substance which contains as many
atoms or molecules as exactly 12 grams of carbon-12. The numbers of atoms in one mole is called
Avogadro's number, denoted by NA. Its value is
NA = 6.0221367×1023 mol-1 (8.12)
In the special case when the amount of gas is one mole the constant in the equation 8.11 is
PHYSICS 1 2002 S. Nitsolov 49
designated by R.
pV = RT (8.13)
The constant R, called gas constant, has the same value for all gases.
R = 8.3144 J/mol⋅K (8.14)
In the general case, the gas contents an arbitrary number nm of moles. Since for a given
temperature and pressure the volume is proportional to the amount of gas, we may write
pV = nmRT (8.15)
This equation is called the ideal gas low or equation of state of an ideal gas. We use the term
"ideal" since all real gases show small deviations from it. For the dilute gases the ideal gas low is a
good approximation, which becomes more accurate as the gas density decreases.
The equation of an ideal gas may be written in another form. If N is number of molecules of the
gas, the number of moles is
N
nm = (8.16)
NA
Substituting for nm into 8.15, we obtain another form of ideal gas equation.
N
pV = RT = NkT (8.17)
NA
where
R
k= = 1.380657×10-25 J/K (8.18)
NA
is called Boltzmann's constant.

8.3.5 DALTON'S LAW


A mixture of several gases obeys laws very similar to those, which govern chemically
homogeneous gases. The pressure of a mixture is given by Dalton's law:
the pressure of a mixture of gases is equal to the sum of the partial pressures of all the
individual components.
p = ∑ pi (8.19)
i
where pi is the partial pressure of the i-th component, that is, the pressure, which this component
would exert if it alone occupied the volume of the mixture at the same temperature.

8.4 KINETIC THEORY OF THE IDEAL GAS

8.4.1 THE MICROSCOPIC INTERPRETATION OF PRESSURE


Our goal in this section is to demonstrate the connection between the properties of the gas
molecules and the macroscopic variables of state as pressure and temperature. From microscopic
point of view a gas may be considered as a system composed of a large number of molecules
moving freely in all directions and with different velocities. The molecules interact between
themselves only during the collisions. The pressure exerted on the walls of container is due to the
collisions of the gas molecules with the wall. During each collision there is an exchange of
momentum, that is, there are forces exerted by the colliding molecules. Since the number of
molecules striking the wall per unit of time is very large, the total normal force F exerted by the
molecules on a given area A of the wall does not vary. Then the pressure p of gas is defined by
F
p= (8.20)
A

8.4.2 CALCULATION OF PRESSURE OF AN IDEAL GAS


Let nm moles of an ideal gas in thermal equilibrium at temperature T and pressure p be confined in a
container of volume V. To calculate the pressure on the walls of the container we can neglect
intermolecular collisions since the molecules of an ideal gas are far enough apart. We assume for
PHYSICS 1 2002 S. Nitsolov 50
simplicity the molecules to interact with the wall as elastic balls that collide with a massive object.
We consider first a single molecule of mass m hitting the wall of the container with velocity v and
rebounding with velocity v' (Figure 8.2). From the theory of elastic collisions we have (Eq.6.49).
∆vN = vN ∆vT = 0 (8.21)
or
∆v = - 2vN (8.22)
The molecule strikes the container and rebounds at
an angle equal to the angle of incidence θ. The
change in momentum after the impact is normal to
the wall. Its magnitude is
∆p = m∆v = 2mvN = 2mvcosθ (8.23)
Figure 8.2 Elastic collision of a molecule Let us assume that the gas is enclosed in a
with the wall of the container 4
spherical container of radius R ( V = π R 3 )
3
shown in Figure 8.3. Since the molecule interacts only with the wall it moves along a straight line
with constant velocity between two successive collisions. It can be seen from Figure 8.3 that the
trajectory of the molecule consists of chords of equal length
∆l = 2Rcosθ (8.24)
The time between any two successive impacts is
∆l 2 R cos θ
∆t = = (8.25)
v v
so that the number of impacts per unit time is
1 v
= (8.26)
∆t 2 R cos θ
The rate with which the momentum is transferred to
the wall is equal to the product of the number of
Figure 8.3 The motion of a molecule in a impacts per unit time and the change of momentum
spherical container for one impact (Eq. 8.23).
∆p v mv 2
= 2mv cos θ = (8.27)
∆t 2 R cos θ R
To find the force of pressure we have to summarise the contributions of all the gas molecules
N
∆p N
mv 2 m N mN 2
F = ∑ i =∑ i = ∑ vi2 = vrms (8.28)
i ∆ti i R R i R
where N = nmNA is the total number of molecules in the container and
N
1
vrms =
N
∑v
i
2
i (8.29)

F
is called the root-mean-square (rms) velocity of the molecules. Since the pressure is p = and
A
the area of the spherical container is A = 4π R 2 we have
F 1 mN 2 1 mN 2 1 mN 2
p= = vrms = vrms = vrms (8.30)
4π R 2
4π R R
2
3 4 3 V
πR 3

3
or
1
pV = mNvrms 2
(8.31)
3
This result relates macroscopic quantities as pressure and volume to the root-mean-square velocity,
which is a typical microscopic characteristic.
PHYSICS 1 2002 S. Nitsolov 51
8.4.3 THE MICROSCOPIC INTERPRETATION OF TEMPERATURE
Comparing Eq. 8.31 with the ideal gas equation in the form pV = NkT (Eq. 8.17), we have
2
mvrms
T= (8.32)
3k
or
3kT
vrms = (8.33)
m
We see that the temperature of the ideal gas depends only on the mass of the molecules and their
root-mean-square velocity. The root-mean-square velocity can be related to the average kinetic
energy Ek,ave of the translational molecular motion. By definition
1 N mv 2
Ek,ave = ∑ i (8.34)
N i 2
so that
1 N mv 2 m  1 N  m 2
Ek,ave = ∑ i =  ∑ vi2  = vrms (8.35)
N i 2 2N i  2
Using the relation between Ek,ave and vrms, Eq.8.32 and 8.33 become
2 Ek,ave
T= (8.36)
3k
and
3
Ek,ave = kT (8.37)
2
This result shows that the temperature of the ideal gas is measure of the average translational kinetic
energy of the gas molecules.

8.4.4 THE INTERNAL ENERGY OF AN IDEAL GAS


By definition (Eq. 6.41), the internal energy U = Ep,int + Ek,int of an ideal gas in container is a sum of
the potential energy of the intermolecular forces and the kinetic energy associated with the motion
of the gas molecules relative to the centre of mass of the system. Since the intermolecular forces of
an ideal gas are negligible, the internal energy consists only of the molecular kinetic energy of the
"disordered" motion of molecules (the "ordered" motion of the gas as a whole is associated with
orbital kinetic energy). It is not difficult to calculate the internal kinetic energy for a monatomic gas
whose molecules can be treated as point-like particles. The translational kinetic energy of the
"disordered" motion is the only internal energy the monatomic molecules have or
N
mv 2
U =∑ i (8.38)
i 2
By use of Eq. 8.34 and 8.37 the internal energy can be written as
3 3
U = NkT = nm RT (8.39)
2 2
Therefore, the internal energy of an ideal monatomic gas is a linear function of the temperature and
number of molecules and is independent on other variables. This statement is also valid for any
ideal poliatomic gas but the coefficient of proportionality depends on the degrees of freedom of the
molecule. According to the so-called principle of equipartition of energy, the average energy of
f
the molecule is equally distributed along its f degrees of freedom and is equal to kT . For
2
example, a diatomic molecule has three translational, two rotational and two vibrational degrees of
freedom, totally f = 7, and its internal energy should be

7
U= NkT (8.40)
2
PHYSICS 1 2002 S. Nitsolov 52

9 FIRST AND SECOND PRINCIPLE OF


THERMODYNAMICS

9.1 INTRODUCTION
Thermodynamics is a branch of physics which studies the thermal properties of macroscopic
systems without explicitly considering the microscopic structure of matter. The behaviour of the
systems is described by macroscopic concepts such as those of temperature, pressure and heat.
Thermodynamics is based on several principles, which are a generalisation of numerous
observations and experiments. Although thermodynamics was developed before the microscopic
nature of matter was well understood, the principles of thermodynamics are ultimately explained by
a statistical treatment of the random motion of atoms and molecules.

9.2 INTERNAL ENERGY AND WORK OF MANY-PARTICLE SYSTEM


Any thermal system is composed of large number of particles. The energy transfer between the
 2
MvCM 
system and its surroundings is given by the work-energy relation ∆  Eint +  = Wext (Eq. ). If
 2 
we choose C-frame of reference, the orbital kinetic energy becomes zero and we have
∆U = Wext (9.1)
where U ≡ Eint is the internal energy and Wext is the work of external forces on the system. That is,
the increment of the internal energy is equal to the work done by the external forces on the system.
There are various mechanisms of energy exchange between the system and its surroundings. In this
connection, it was found convenient to decompose Wext into two parts, called mechanical work or
simply work and heat.

9.3 WORK.
The energy exchange between a system of many particles with its surroundings is called work when
it is associated with a collective displacement under the action of a macroscopic force. By
convention, the work done by the system on its surroundings is usually taken as positive and
denoted by W. For example, when a gas in a cylinder expands against the piston (Figure 9.1), it does
work. The force on the piston of area A is
F = pA (9.2)
The infinitesimal work dW done by the gas to
move the piston a distance dx is
dW = Fdx = pAdx = pdV (9.3)
where dV = Adx is the infinitesimal change in
volume of the gas. For a finite change from V1 to
Figure 9.1 Work done by a gas when it V2 the work done by the gas is
changes its volume. V2

W = ∫ pdV (9.4)
V1

This equation is valid for the work in any change of


the volume of gas. To compute the integral we need
the relation between p and V (shown in the p - V
diagram on Figure 9.2). The work done by the gas is
Figure 9.2 Work done by a gas during a given by the area under the curve AB in the p - V
transformation. diagram. This result shows that the work depends also
PHYSICS 1 2002 S. Nitsolov 53
on the type of process. We shall calculate W for some special transformations.

9.3.1 CYCLE
The process in which the system returns to its
initial state is called cycle (Figure 9.3). The work
in a cycle is determined by
W = ∫ pdV (9.5)
In the p - V diagram the work is given by the area
Figure 9.3 Work done by a gas during a cyclic enclosed within the curve representing the cycle.
process The sign of W depends on the direction of the
cycle. If the cycle is realised in the clockwise
direction, the work done by the gas is positive and vice versa.

9.3.2 ISOBARIC PROCESS


In the case of an isobaric process, p = const,
V2 V2

W= ∫ pdV = p ∫ dV = p (V
V1 V1
2 − V1 ) = p∆V (9.6)

From the ideal gas law we have p∆V = nmR∆T so that


W = nm R∆T (9.7)
The work in isobaric process is proportional to the change in temperature of the ideal gas.

9.3.3 ISOTHERMAL PROCESS


For an isothermal transformation of an ideal gas we have
n RT
p= m (9.8)
V
and substituting this result for p in Eq.9.4, we find
V2 V2
dV V
W = ∫ pdV = nm RT ∫ = nm RT ( ln V2 − ln V1 ) = nm RT ln 2 (9.9)
V1 V1
V V1

9.4 HEAT
There exist some energy exchanges between the system and its surroundings which are not
associated with a collective displacement and are, therefore, not associated with the mechanical
work. This is the case of energy transfer between the system and another body which is hotter or
colder than the system. In this case, the change of the internal energy is a result of interactions
between the particles of systems put into contact. The molecules, which are in random motion,
collide each other. At each collision, a small amount of energy is exchanged. Energy is transferred
also at distance by emission and absorption of electromagnetic radiation. The concept of heat is
referred to the energy transfer due to the microscopic exchanges of energy in the random collisions
between the molecules of the system and its surroundings, without any collective displacement. The
amount of heat Q is defined as the net energy transferred in this way to the system. By convention,
it is positive when the heat is absorbed by the system and it is negative when the heat is given off by
the system. When heat is neither absorbed nor given off, the transformation is called adiabatic.
When the temperature of two systems is the same (that is, they are in thermal equilibrium), there
is no heat transfer between them. That means that, on the average, the same amount of energy is
transferred in one direction as in the other. When the temperature is different, heat is transferred
from the system at higher temperature to the system at lower temperature.

9.5 FIRST LAW OF THERMODYNAMICS


Using the notions of work and heat, we may rewrite the work-energy relation (Eq. (9.1)) in the form
PHYSICS 1 2002 S. Nitsolov 54
∆U = Q − W (9.10)
That is,
the increment of internal energy of a system is equal to the heat absorbed minus the work
done by the system on its surroundings.
This statement, called the first law of thermodynamics, is the law of conservation of energy
applied to many-particle systems.
In order to account the energy exchange by a radiation transfer we may write
∆U = Q − W + R (9.11)
where R is the radiation energy absorbed by the system. This is a more general form of the first law
of thermodynamics.
The internal energy depends only on the state of the system and it does not depend on the type of
the process. We have seen that the work depends on the process. Now, writing the first law of
thermodynamics in the form
Q = ∆U + W (9.12)
we see that the heat also depends on the type of the process because it is related to W. We shall
apply the first law of thermodynamics to various transformations.

9.5.1 CYCLE
When the transformation is cyclic, the initial and final states are the same and the internal energy
does not vary, ∆U = 0. Then Eq. 9.10 becomes
Q =W (9.13)
Thus, the work done by the system in a cyclic process is equal to the heat absorbed by the system.

9.5.2 ADIABATIC PROCESS


For an adiabatic transformation, by definition, Q = 0, and the first law of thermodynamics takes the
form
∆U = −W (9.14)
That is, the internal energy decreases by an amount equal to the work done by the system. Since the
internal energy of an ideal gas depends only on the temperature, when the gas expands (the work is
positive), its temperature decreases and vice versa. It can be shown that in a quaei-static adiabatic
process the pressure and volume of ideal gas are related by the equation
pV γ = const (m fixed) (9.15)
where γ = Cp/CV. The molar heat capacities Cp and CV are considered in the next section.

9.5.3 ISOCHORIC PROCESS


In the case of isochoric transformation ∆V = 0, and, therefore W = 0. Then the first law becomes
∆U = Q (9.16)
When no work is done, the change in internal energy is equal to the heat added to the system.

9.5.4 RADIATIVE PROCESS


For a radiative process, when energy is transferred only by emission and absorption of radiation,
∆U = R (9.17)

9.6 HEAT CAPACITY


It was found experimentally that the heat required to produce a given rise of temperature of a given
body is proportional to the mass of a body and to the change of its temperature. If the body contains
nm moles of substance, we may write
dQ = Cnm dT (9.18)
where C is the molar heat capacity of the substance. It is defined as the amount of heat required to
rise the temperature of one mole of the substance by one degree. That is,
PHYSICS 1 2002 S. Nitsolov 55
1 dQ
C= (9.19)
nm dT
Since the heat depends on the process, the heat capacity also depends on the process of heating. The
most important heat capacities are the heat capacity at constant pressure, Cp, and the heat capacity at
constant volume, CV, defined by
1  dQ 
Cp =   (9.20)
nm  dT  p = const
1  dQ 
CV =   (9.21)
nm  dT V =const
In the case of isochoric process Q = ∆U so that
1  dU 
CV =   (9.22)
nm  dT V =const
In the case of isobaric process
1  dQ  1  dU  1  dW 
Cp =   =   +   (9.23)
nm  dT  p = const nm  dT  p =const nm  dT  p = const
3
For ideal monatomic gas we use the expression for the internal energy U = nm RT (Eq. 8.39)
2
and the expression for the work in isobaric process W = nm R∆T (Eq. 9.7 ). Thus we obtain
3
CV = R (9.24)
2
3 5
Cp = R + R = R (9.25)
2 2

9.7 ENTROPY

9.7.1 IRREVERSIBLE AND REVERSIBLE PROCESSES


Consider a cycle composed of compression and
expansion of a gas in a cylinder with piston (Figure
9.4). In the initial state (a) the gas is at equilibrium. If
we move the weight to the top of the piston, the gas is
rapidly compressed as shown in (b). We may expand
the gas to restore the initial state. If the weight is
removed the external pressure drops and the gas
expands up to the initial position on the piston. The
(a) (b) (c) gas approaching equilibrium returns to its initial state,
Figure 9.4 Irreversible process as shown in (c). The cyclic process is completed.
However, the weight is now at the bottom. That is, in
a cycle consisted of non-equilibrium processes, the
system returns to its initial state, but a permanent
change is produced in the surroundings. The system
cannot be restored to its original state without also
making some change in the surroundings. We say that
such process is irreversible.
We consider now a cycle composed of equilibrium
compression and expansion of the same system
(a) (b) (c) (Figure 9.5). Firstly we move one small weight to the
Figure 9.5 Reversible process top of the piston shown in (a). The external pressure
increases slightly and the equilibrium of the gas is slightly disturbed. The piston compresses the gas
until equilibrium is restored. Then we may repeat the process a number of times by moving the next
PHYSICS 1 2002 S. Nitsolov 56
weight until the piston approaches the lowest position, shown in (b). To restore the gas to the initial
state, we have to place back, in the reverse order, the same weights until the cycle is completed
leaving negligibly small change in the surroundings (c). That is, it is possible to restore the system
to its original state making some infinitesimally small change in the surroundings. Such process is
called reversible.

9.7.2 DEFINITION OF ENTROPY


If a system, at temperature T, absorbs heat dQ during an infinitesimal reversible transformation, the
entropy of the system is defined as a quantity whose change is determined by
dQ
dS = (9.26)
T
For a finite reversible transformation
2 2
dQ
∆S2 − ∆S1 = ∫ dS = ∫ (9.27)
1 1
T
When the transformation is irreversible, it may be replaced by a reversible transformation, which
connects the initial and the final state. It is important, that the change in entropy is independent on
the type of the reversible process. In the case of a cycle, (Figure 9.3) we may decompose a cyclic
transformation into two transformations between points 1 and 2 and write
2 1 2 2
dQ dQ dQ dQ dQ
∆S = ∫ =∫ +∫ =∫ −∫ =0 (9.28)
T 1
T 2
T 1
T 1
T
That is, in a cycle, the change in entropy of a system is zero.
dQ = TdS (9.29)
and therefore
2
Q = ∫ TdS (9.30)
1
For a cycle ∆U = Q - W = 0 and
W =Q= ∫ TdS (9.31)
where the integral depends on the particular reversible transformation.

9.7.3 THE ENTROPY OF AN IDEAL GAS


To calculate the entropy of ideal gas we start from the definition and apply the first law of
thermodynamics
dQ dU + pdV
dS = = (9.32)
T T
3
Using the formulae for the internal energy U = nm RT (Eq. 8.39) and the ideal gas law
2
pV = nm RT (Eq.8.15), we obtain
3 dT dV  3 dT dV 
dS = nm R + nm R = nm R  +  (9.33)
2 T V 2 T V 
Integrating, we have
 3 dT dV  3 
S = nm R  ∫ +∫  = nm R  ln T + ln V  + S0 (9.34)
2 T V  2 
or
3 T V 
∆S = S2 − S1 = nm R  ln 2 + ln 2  (9.35)
 2 T1 V1 

9.7.4 SECOND LAW OF THERMODYNAMICS


There exist some processes that are unidirectional. The irreversible processes are always
PHYSICS 1 2002 S. Nitsolov 57
unidirectional: heat is transferred from a hot body to a colder one, gases expand into vacuum etc.
The reverse of any of these processes is not observed although that it would not violate any of the
known conservation laws. In the case of a heat transfer the change in entropy of the system
consisted of two bodies is
∆Q1 ∆Q2
∆S = ∆S1 + ∆S2 = + (9.36)
T1 T2
If T1 < T2, then Q1 = - Q2 > 0 and
1 1 T −T 
∆S = ∆Q1  −  = ∆Q1  2 1  > 0 (9.37)
 T1 T2   T1T2 
In the general case of processes in isolated system, it was found that
the total entropy of the system remains constant if the processes are reversible or increases if
the processes are irreversible,
or
∆S = ∑ ∆Si > 0 (9.38)
i
This is the second law of thermodynamics. Processes which occur in an isolated system are those
in which energy is conserved and entropy increases or remain constant.

9.8 CARNOT CYCLE

(a) p - V diagram (b) T - S diagram


Figure 9.6 Carnot cycle

A thermal engine is a device that operating cyclically, transforms heat into work. During every
cycle, energy is transferred as heat from a heater to the working substance of the engine. A portion
of this energy is transformed into useful work and the rest is given (lost) to a cooler. To analyse a
thermal engine we shall consider a reversible cycle called Carnot cycle. A particular Carnot cycle
is shown in p - V diagram and T - S diagram in Figure 9.6. It is composed of two isothermal
transformations A→B and C→D and two adiabatic transformations B→C and D→A. During the
isothermal process A→B at a temperature T1, the working substance absorbs heat Q1 from the
heater and expands performing work. The change of entropy is
Q
∆S AB = 1 > 0 (9.39)
T1
During the adiabatic process B→C, the working substance continues to expand and perform work.
Its temperature decreases from T1 to T2. There is no heat transfer and therefore, the entropy is
constant.
∆S BC = 0 (9.40)
During the isothermal process C→D at a temperature T2 < T1, the working substance is compressed
PHYSICS 1 2002 S. Nitsolov 58
and it gives off heat Q2 to the cooler. The change in entropy is negative,
Q
∆SCD = 2 < 0 (9.41)
T2
During the adiabatic compression D→A, the temperature increases from T2 to T1 and the entropy
does not vary.
∆S DA = 0 (9.42)
At the end of the cycle the net changes in entropy and internal energy must be zero
Q Q
∆S = ∆S AB + ∆S BC + ∆SCD + ∆S DA = 1 + 2 = 0 (9.43)
T1 T2
∆U = Q − W = 0 (9.44)
The work W done in the cycle can be determined using the T - S diagram. It is equal to the area
of the rectangle ABCD.
W = ( T1 − T2 )( S 2 − S1 ) (9.45)
The efficiency η of a thermal engine is determined as the ratio of the net work done by the engine to
the heat Q1 transferred from the heater to the working substance, that is,
W
η= (9.46)
Q1
Since Q1 = T∆SAB = T1(S2 - S1),
(T − T )( S − S ) (T − T ) T
η = 1 2 2 1 = 1 2 = 1− 2 (9.47)
T1 ( S2 − S1 ) T1 T1
PHYSICS 1 2002 S. Nitsolov 59

10 ELECTRIC CHARGE AND ELECTRIC FIELD

10.1 INTRODUCTION
The science of electricity originates centuries ago in the observation known to Thales of Milethus
(600 - 548 B.C.). A rubbed with silk piece of amber attracts bits of straw. Similar phenomena may
be demonstrated by rubbing a glass rod with silk. As a result of rubbing, these bodies acquire a new
property called electricity. The word "electricity" comes from the Greek word electron, which
means amber. The electricity plays an important role in the modern technology, motors, computers,
communications, etc. But electricity is important not only because it has given rise to the different
branches of technology. Generally speaking, the electrical interaction determines the properties of
matter in bulk since the forces that act between atoms and molecules to hold them together to form
liquids and solids are electrical forces. Our purpose in this chapter is to introduce the basic concepts
of electric interaction of static electric charges in free space.

10.2 ELECTRIC INTERACTION AND ELECTRIC CHARGE

10.2.1 FUNDAMENTAL LAW OF ELECTRIC INTERACTION


Suppose that we place a glass rod rubbed with a
cloth near another electrified glass rod hanging
from a string (Figure 10.1). The electrified rods
will repel each other. The same result occurs if we
repeat the experiment with electrified rods of
ebonite. On the other hand, a rubbed rod of ebonite
will attract a rubbed glass rod. Thus we conclude
Figure 10.1 Interaction between electrified that there are two kinds of electrification. We may
bodies then state the fundamental law of electric
interaction as follows:
The interaction between two electrified bodies is repulsion when they have the same state of
electrification and attraction when they have different kinds of electrification.
Benjamin Franklin (1706 - 1790), American scientist and statesman named the kind of
electrification of rubbed amber positive and the kind of electrification of a rubbed glass negative.
We see that electric interaction is similar to the gravitation but they differ from one another since
there are two kinds of electric interaction instead of only one kind in the case of gravitation.

10.2.2 ELECTRIC CHARGE


In order to characterise quantitatively the state of electrification, we attach to each body a
coefficient, called electric charge, or simply charge. The charge determines the strength of electric
interaction of the body with other bodies. Since there are two kinds of electrification, the charge
may be positive or negative depending on the kind of electrification. If a body is electrically neutral,
its charge is zero. The net charge of a body is the algebraic sum of its negative and positive charges.
The electric charge may be defined operationally using the following procedure. In order to
compare two charges q and q' we place them consecutively in the same point near an arbitrary
chosen charged body and measure the electric forces F and F'. Defining the values of the charges q
and q' as proportional to F and F', respectively, we may write
q F
= (10.1)
q' F '
Thus, if we know the charge q', we can determine the charge q. In defining of the charge, we have
assumed implicitly that the charges are at rest or are moving with very small velocities relative to an
PHYSICS 1 2002 S. Nitsolov 60
observer. The motion of the electric charges gives rise to additional forces, which will be discussed
later. In this chapter we shall discuss only the electric interaction between stationary charges, which
is the subject of electrostatics.

10.2.3 PRINCIPLE OF CONSERVATION OF ELECTRIC CHARGE


Atoms, molecules and most macroscopic bodies seem to be neutral. The electrification of neutral
bodies by friction may be explained by transfer of charge from one body to another. A glass rod
rubbed with a cloth acquires a positive charge. The measurement shows that a negative charge of
equal magnitude appears on the cloth. That is, the net charge of the system of the two bodies
remains zero. This is an example of the principle of conservation of electric charge, which states
that
the total charge of an isolated system does not change.
There are not any known exceptions to this law.

10.2.4 THE QUANTIZATION OF ELECTRIC CHARGE


As we discussed earlier, the atom is composed of positively charged nucleus and several negative
electrons moving about the nucleus. The nucleus consists of uncharged neutrons and positive
protons. The charge of an electron is equal in magnitude to the charge of a proton. The absolute
value of the charge of electron
e = 1,602177×10-19 C (10.2)
is called elementary charge. The atom has the same number of electrons and protons in its normal
state, and hence it is neutral. Since each body consists of a multiple number of protons and
electrons, we may conclude that the electric charge is quantized. That is,
any amount of electric charge is a multiple of the elementary charge e.

10.3 COULOMB'S LAW


Charles A. de Coulomb (1736 - 1806) measured the electric forces between charged bodies and
established that
the force of electric interaction between two charged point-like particles at rest is
proportional to their charges and to the inverse of the square of the distance between them,
and its direction is along the line joining the two charges.
This statement, called Coulomb's law, may be expressed by
qq
F12 = 1 2 2 u12 (10.3)
4πε 0 r12
where F12 is the force exerted on a charge q1 by a
charge q2 (Figure 10.2), r12 is the distance between
the charges,
ε0 = 8,85418781762 ×10-19 C2/N⋅m2 (10.4)
Figure 10.2 Electric interaction between two is called the vacuum permitivity and u12 is a unit
point-like charges. vector directed from q2 to q1 The SI unit of charge
is the coulomb, C, defined as the amount of
charge that passes through a section of a conductor when the current is one ampere. Notice that in
accord with Newton's third law
qq qq
F21 = 1 2 2 u 21 = − 1 2 2 u12 = −F12 (10.5)
4πε 0 r12 4πε 0 r12
When the number of charged particles is greater than two, Coulomb's law must be supplemented by
the principle of superposition of forces. According to this principle, the resultant force F on a
charge q is a vector sum of the forces F1, F2, F3,... due to each of the charges q1, q2, q3,... . That is,
qqi q qi
F = ∑ Fi =∑ u1 = ∑ u1 (10.6)
i i 4πε 0 ri
2
4πε 0 i ri 2
where r is the distance from q to qi and ui is the unit vector directed from qi to q. In the case of
PHYSICS 1 2002 S. Nitsolov 61
continuous distribution of charges the sums should be replaced by integrals.
q u q ρu
F= ∫
4πε 0 V r 2
dq = ∫
4πε 0 V r 2
dV (10.7)

dq
where V is the entire volume of charge distribution, ρ = is the volume charge density.
dV

10.4 ELECTRIC FIELD


Suppose that we place a charge q at different positions around other electric charges. Due to its
electric interaction with them, the charge q experiences a force F given by Coulomb's law. The
region, in which an electric charge experiences a force of electric origin, is called electric field. The
strength of the electric field at a point may be characterised quantitatively by the intensity of the
electric field, or simply electric field, E, defined as the force per unit charge placed in the point.
F
E= (10.8)
q
It is assumed that the test charge q used to measure the force F is very small. Then the field created
by q does not change the distribution of charges that are responsible for the field that is to be
examined. The SI unit of the intensity of the electric field following from the definition is N/C, but
later it will be seen that V/m is an equivalent unit, where V is the abbreviation of volt. If the field is
produced by a single charge q', its intensity is equal to
F qq ' q'
E= = u= u (10.9)
q ' 4πε 0 r q
2
4πε 0 r 2
where r is the distance from the charge to the point and u is the unit vector directed from the charge
to the point.
In the case of several point charges, the electric field may be calculated using the principle of
superposition (Eq. 10.6).
F F qi
E = = ∑ i = ∑ Ei = ∑ ui (10.10)
i 4πε 0 ri
2
q i q i

qi
where Ei = u i is the electric field due to the charge qi. We see that the principle of
4πε 0 ri 2
superposition is valid also for the intensity of the electric field.

10.5 ELECTRIC POTENTIAL

10.5.1 ELECTRIC POTENTIAL ENERGY


When n electric field moves a charge from one place
to another it does work. Consider the work done by
the electric force when a test charge q is moved from
point A to point B in an electric field. The
infinitesimal work dW done by the field on the charge
q in an infinitesimally small displacement dr is (Eq.
5.3)
dW = F.dr = qE.dr = qE.dr (10.11)
Figure 10.3 Work done by the electric field Suppose firstly that the field is produced by a point
produced by a point charge. charge q' (Figure 10.3), we can rewrite this
expression as
q′ qq ′ 1
dW = qE.dr = q u.dr = u.dr (10.12)
4πε 0 r 2
4πε 0 r 2
where r is the distance between q' and q and u is the unit vector directed from q' to q. The scalar
product may be written as
PHYSICS 1 2002 S. Nitsolov 62
u.dr = drcosθ = dr (10.13)
where dr is the infinitesimal change of the distance r between the charges. Then the work WAB in
finite displacement from point A to point B is given by
qq ′ 1 qq ′ qq ′  1 1 
B B B
1
WAB = ∫ dW = ∫ dr = − ∫ d = −  −  (10.14)
A
4πε 0 A r 2
4πε 0 A  r  4πε 0  rB rB 
This result indicates that the work done by the electric field in moving a charge from one position to
another depends only on the initial and final position of the charge but not on the path taken.
Therefore, the electric force is conservative force. This statement is valid also in the general case,
when the electric field is produced by several electric charges. In this case, the net work WAB is the
sum of the works WAB1, WAB2, WAB3,... done by the fields of the individual charges q1, q2, q3,... .
That is,
q  1 1 
WAB = q ∑
i
− i  − 
4πε 0  riB riB 
(10.15)

We see that in the general case the electric force is also conservative and it is possible to define
potential energy Ep of the electric field. If we compare this result with the general relation between
work and potential energy, WAB = Ep(B) - Ep(A), (Eq. 5.24), the potential energy Ep of the electric
force in moving the test charge q can be written in the form


qi
Ep = q (10.16)
i
4πε 0 ri
The potential energy is set to be zero at infinity distance from the charges, as is the usual practice.
This choice is not essential, however, since the potential energy is determined within an arbitrary
chosen constant.

10.5.2 ELECTRIC POTENTIAL


The electric potential V at a point is defined as a potential energy per unit positive charge placed at
this point, or
Ep
V= (10.17)
q
The SI unit of electric potential is volt (abbreviated V) equal to J/C. The potential as well as the
potential energy is determined within an arbitrary constant. By means of potential the relation
between work and potential energy can be written in the form
A
W = q ∫ E.dr = −∆E p = − q∆V = q ( VA − VB ) (10.18)
B
Hence,
A
W
q ∫B
= E.dr = −∆V = VA − VB (10.19)

or
A
∆V = − ∫ E.dr (10.20)
B
The potential difference between two points is equal to the work done by the field to move a test
unit charge from one point to another. For an infinitesimally small displacement dr the potental
difference is given by
dV = - E.dr (10.21)
If the integral is taken over a closed path L, the potential difference is zero and
∫ E.dr = 0
L
(10.22)

that is,
the circulation of the static electric field along any closed path is zero.
When the electric field is due to a group of point charges, substituting the electric potential
PHYSICS 1 2002 S. Nitsolov 63
energy from Eq. 10.16 into Eq. 10.17, we find

∑V
qi
V= Vi = (10.23)
4πε 0 ri
i
i
which gives the potential as a sum of potentials of the individual charges.

10.5.3 RELATION BETWEEN ELECTRIC POTENTIAL AND ELECTRIC FIELD


The potential V at a given point A can be found from Eq. 10.19 choosing point P at infinity. The
result is
A
V = − ∫ E.dr (10.24)

If the field is known, then V can be determined evaluating the integral. The general relation between
potential energy and force F = - gradEp (Eq. 5.31) can be used to perform the reverse operation.
Substituting F = qE and Ep = qV, we have
qE = - grad(qV) (10.25)
or
E = - gradV (10.26)
That is, the electric field is the negative of the gradient of the electric potential. The minus sign
shows that the electric field points in direction in which the potential decreases.

10.6 MOTION OF CHARGED PARTICLES IN ELECTRIC FIELD

10.6.1 A POINT CHARGE


Since the intensity of the electric field is equal to the force per unit charge, the force acting on a
charged particle is
F = qE (10.27)
If the charge is positive, the force has the direction of E. But if the charge is negative, the force has
opposite direction to E. Therefore, the positive and negative charges placed in an electric field tends
to move in opposite directions. Thus, the field may cause a charge separation, called polarisation.
In the case of uniform electric field, the force is constant and the acceleration is also constant.
qE
a= (10.28)
m
We see that the motion of the charged particle in an electric field is similar to the motion of a
particle under the force of gravity.

10.6.2 AN ELECTRIC DIPOLE


A system of two equal and opposite charges + q
and - q, separated by a very small distance a, is
called electric dipole (Figure 10.4). The electric
dipole moment or dipole moment is defined as
Figure 10.4 Electric dipole p = qa (10.29)
where a is the vector displacement from the negative to the positive charge. Many molecules have
dipole moments. For example, in the HCl molecule the H electron is displaced closer to the Cl atom
and the molecule, which is neutral as a whole has dipole moment directed from the Cl to the H
atom. Usually, the first step in description of more complicated configuration of charge is to
determine its dipole moment.
PHYSICS 1 2002 S. Nitsolov 64
Suppose we place a dipole in an electric field
(Figure 10.5). The forces on the charges are
F1 = qE1 and F2 = qE2 and their resultant is
F = F1 + F2 = qE1 - qE2 = q(E1 - E2) (10.30)
If the field is uniform, E1 = E2 = E, the resultant
force on the dipole is zero. Then the forces on the
charges form a couple whose torque τ relative to
the centre of the dipole is
Figure 10.5 Electric dipole in an electric field.
1 1
τ = ( a) × F1 + (− a) × F2 = a × (qE) = (qa) × E = p × E (10.31)
2 2
This result shows that an electric dipole in uniform field experiences a torque that tends to orient the
dipole parallel to the field.
The electric field due to an e;lectric dipole can
be found calculating its potential. (Figure 10.6).
The symmetry of the problem shows that the
potential at a point P depends oily on the distance r
from the centre O of the dipole to P and on the
angle θ between r and p. According to the
principle of superposition, the potential is the sum
of the potentials due to each of the two point
charges, so that
−q q q  r2 − r1 
V= + =   (10.32)
Figure 10.6 Electric field due to a dipole. 4πε 0 r1 4πε 0 r1 4πε 0  r2 r 
We shall limit our considerations to points which
are distant from the dipole, that is, a << r. Then we have r2 -r1 = acosθ and r1r2 = r2 resulting in
qa cos θ p cos θ
V = = (10.33)
4πε 0 r 2 4πε 0 r 2
y
Substituting cos θ = and r = x 2 + y 2 in the previous equation we obtain
r
p y
V= (10.34)
4πε 0 ( x + y 2 )3 2
2

Using Eq. 10.26, E = - gradV, we find


∂V p 3 xy ∂V p 2 y2 − x2 ∂V
Ex = − = ; E = − = ; Ez = − = 0 (10.35)
∂x 4πε 0 ( x 2 + y 2 )5 2 ∂y 4πε 0 ( x 2 + y 2 )5 2
y
∂z
p
2
Along the axis of the dipole, x = 0 and Ex = 0; E y = ; Ez = 0. We see that at large distances
4πε 0 y 3
the field of the dipole decreases with the distance as r-3 while the field of a single point charge is
proportional to r-2.
PHYSICS 1 2002 S. Nitsolov 65

11 GAUSS' LAW

11.1 INTRODUCTION
In this chapter we continue the discussion of static electric charges in free space. Once we have
introduced the electric field, we discuss the use of electric lines of force and equipotential surfaces.
Gauss' law is a new formulation of Coulomb's law. It is more general relationship between the
charges and the electric field and provides deeper understanding of the nature of electric field.
Gauss' law sometimes simplifies calculations of the field when there is symmetry in the charge
distribution. Finally, we consider examples of use of the law for situations of high symmetry.

11.2 LINES OF FORCE AND EQUIPOTENTIAL SURFACES

11.2.1 LINES OF FORCE


In order to visualise the electric field we use the
lines of force (Figure 11.1). The tangent at each
point of such a line coincides with the direction of
electric field. The density of lines (number of field
lines drawn per unit cross-sectional area
perpendicular to the lines) is proportional to the
magnitude of the intensity. Force lines originate on
positive charges and terminate on negative
Figure 11.1 A line of force charges. The number of lines originating or ending
on a charge is proportional to its magnitude. They never intersect because if two lines of force
intersect they would have two tangents at the point of intersection and therefore the field would also
have two different directions at one and the same point.
The lines of force of the electric field of a point
charge are straight lines passing through the
charge (Figure 11.2). When the intensity has the
same magnitude and direction everywhere, the
electric field is called uniform. In this case the
lines of force are parallel and equally spaced. A
uniform field may be produced by two parallel
(a) positive charge (b) negative charge metal plates charged with equal but opposite
Figure 11.2 Lines of force (solid) and charges.
equipotential surfaces (dashed) of point charge
11.2.2 EQUIPOTENTIAL SURFACES
The electric potential may be represented
graphically by drawing equipotential surfaces.
Equpotential surfaces are surfaces on which the
potential is the same at all points, V = const. When
a charge moves over an equipotential surface the
potential does not vary. Since the potential does
not change the work in any infinitesimal
displacement over the surface is zero and,
according to Eq. 10.21, dV = E.dr, the tangential
component of the electric field must also be zero.
Figure 11.3 Lines of force (solid) and
Therefore, the electric field at any point is
equipotential (dashed) surfaces
perpendicular to the equipotential surface, as
shown in Figure 11.3. The equipotential surfaces never intersect because if two surfaces intersect
PHYSICS 1 2002 S. Nitsolov 66
they would have two different normals at the point of intersection and therefore the field would also
have two different directions at one and the same point. The equipotential surfaces for point charges
are shown in Figure 11.2.

11.3 FLUX OF THE ELECTRIC FIELD


Consider a plane surface S immersed in a uniform
electric field E. The normal uN to the surface
makes an angle θ with the direction of E (Figure
11.4). The quantity
ΦE = EScosθ (11.1)
is called the flux of the electric field or electric
flux through the surface S. The flux may be
written in the form
Figure 11.4 Electric flux ΦE = EScosθ = ES⊥ (11.2)
where S⊥ = Scosθ is the area of the projection of S
π π
on a surface perpendicular to E. Note that S⊥ < 0 if θ < and S⊥ > 0 if θ > . It follows from this
2 2
result and the properties of lines of force that the electric flux is proportional to the number of lines
of force passing through S. Since
Ecosθ = E.uN (11.3)
the flux may be written as a scalar product
ΦE = EScosθ = E.uNS =E.S (11.4)
π π
where S = uNS is called area vector. The electric flux is positive if θ < and negative if θ >
.
2 2
The flux can be associated with the number of lines of force cutting the surface S. If we choose the
coefficient of proportionality between their density and the magnitude of the field to be 1, we see
that the total number of lines of force crossing S is ES⊥, or equal to ΦE.
In the general case when the field is non-uniform or the surface is non-planar, the surface may be
divided into elementary surfaces. If the surfaces are infinitesimally small we can consider them to
be planar and electric field to be uniform on a given elementary surface. Then, the definition of flux
is applicable for each of the surfaces and we can write
dΦE = E.dS = EdS⊥ (11.5)
The total flux through S is given by the surface integral
Φ E = ∫ dΦ E = ∫ E.dS = ∫ E.dS⊥ (11.6)
S S
If the surface is closed, the flux is given by
ΦE = ∫ d ΦE = ∫ E.dS = ∫ E.dS ⊥ (11.7)
In this case, the direction of dS is defined to point outward from the enclosed volume. Thus, for a
closed surface the value of the flux over a specified area is positive if the lines of force point
outwards, and negative if they point inwards.

11.3.1 CALCULATING A FLUX FROM A POINT CHARGE


Gauss's law is a relation between electric charges and their flux through an imaginary closed surface
of arbitrary shape called a gaussian surface. Firstly, we shall derive the Gauss's law in the special
case of one point-like charge. After that, we shall generalise the law for an arbitrary number of
charges. Consider a point charge q inside a gaussian surface S (Figure 11.5). The flux through the
surface is given by
q q dS⊥
Φ E = ∫ EdS⊥ = ∫ dS⊥ = ∫ (11.8)
4πε 0 r 2
4πε 0 r2
PHYSICS 1 2002 S. Nitsolov 67
Taking into account that
dS ⊥
dΩ = (11.9)
r2
is the solid angle corresponding to the elementary
surface dS and that the total solid angle around the
charge is 4π, we may replace the integration over dS
by integration over dΩ. Then we may write
Figure 11.5 A single charge inside a gaussian q q q
surface ΦE = ∫ dS ⊥ = ∫ dΩ = (11.10)
4πε 0 r 2
4πε 0 ε0
If the charge is placed outside the closed surface
(Figure 11.6), we see that the incoming flux is
balanced by the outgoing flux and the total flux is
zero. For example, the elementary flux dΦE' = E'.dS'
through dS' is equal in magnitude to the flux dΦE'' =
E''.dS'' through dS'' since the solid angle dΩ is the
same for both elementary surfaces dS' and dS''.
However, the two fluxes have opposite signs since
Figure 11.6 A single charge outside a gaussian dΦE' is incoming and dΦE'' is outgoing, and therefore
surface their sum is zero. Since the flux for any such pair is
zero the contribution of the charge outside the closed
surface is zero.

11.3.2 CALCULATING A FLUX FROM SEVERAL CHARGES


In the case of several electric charges inside a gaussian surface S, we may calculate the total flux
using the principle of superposition for the electric fields produced by the charges and the previous
result for the flux of one charge. Thus, we have
q  1

 i

 i
( )
Φ E = ∫ E.dS = ∫  ∑ Ei  .dS = ∑ ∫ Ei .dS =∑  i  = ∑ qi
i  ε0  ε0 i
(11.11)

or
ε 0 ∫ E.dS = q (11.12)
where q = ∑ qi is the net charge inside the closed surface.This relation is called Gauss' law. It
i
says that
the electric flux through a closed surface is proportional to the net electric charge enclosed by
the surface.
The electric flux is independent on the shape of the surface and the distribution of the charges inside
the surface.
PHYSICS 1 2002 S. Nitsolov 68
11.4 FIELD PATTERNS

11.4.1 UNIFORMLY CHARGED INFINITE PLANE


We shall determine the electric field due to an
infinite uniformly charged plane (Figure 11.7). If a
charge is distributed over a surface then the charge
per unit area can be expressed as
∆q
σ = lim (11.13)
∆S → 0 ∆S

where σ is called the surface charge density. If


the charge is uniformly distributed then σ = const.
The symmetry of the problem suggests that the
electric field depends only on the distance from the
Figure 11.7 A gaussian cylinder extended plane. The field is perpendicular to the plane and
through an uniformly charged plane. opposite in direction on both sides of the plane.
We choose a gaussian surface as a closed right
cylinder with bases situated symmetrically relative to the plane. The bases and the cross-section of
the cylinder with the plane have the same area S. The Gauss' law for the cylinder is
ε 0 ∫ E.dS = ∫ σ dS = σ S (11.14)
cross −sec tion
To evaluate the closed integral we split it into three parts
∫ E.dS = ∫ E.dS + ∫ E.dS +
S right left

lateral
E.dS (11.15)
base base surface

Since the field is directed along the axis of the cylinder the flux through the lateral surface is zero.
For both sides the field is parallel to the normal so that the fluxes through the bases are equal to ES.
Thus, Gauss' law may be written in the form
ε 0 ES1 + ε 0 ES2 = 2ε 0 ES = σ S (11.16)
or
σ
E= (11.17)
2ε 0
That is, the electric field is uniform. This result is a good approximation for finite planes if the
distance to the plane is much smaller than the distance to the edge of the plane.

11.4.2 TWO UNIFORMLY CHARGED INFINITELY LARGE PLANES WITH EQUAL BUT OPPOSITE
CHARGES
According to the principle of superposition, the
electric field of a pair of plane parallel surfaces
carrying equal but opposite charges (Figure 11.8)
is the sum of the fields E+ and E- due to the
positive and negative surface, respectively.
Outside the surfaces the individual electric fields
are equal in magnitude but opposite in direction so
that their resultant is zero. In the region between
the surfaces E+ and E- are equal in magnitude and
direction. The resultant field E = E+ + E- has the
same direction and magnitude
Figure 11.8 Two uniformly charged planes σ σ σ
E = E+ + E− = + = (11.18)
with equal but opposite charges 2ε 0 2ε 0 ε 0
We see that the electric field outside the plates is zero and the field between the plates is uniform
PHYSICS 1 2002 S. Nitsolov 69
σ
with magnitude E = .
ε0
To find the potential difference between the planes we substitute F = qE and W = q∆V into the
relation between the work done by a constant force, W = F.∆r (Eq. 5.7). Thus, we obtain
σd
∆V = V+ − V− = (11.19)
ε0

11.4.3 CYLINDRICAL CHARGE DISTRIBUTION


We consider firstly an electric charge distributed
uniformly along a straight infinite line. To
characterise such charge distribution we use linear
charge density λ defined as the charge per unit
length. It can be expressed as
∆q
λ = lim (11.20)
∆S → 0 ∆l

The symmetry of the problem indicates that the


electric field at each point is radial and depends
only on the distance from the axes of the cylinder to
the point. We choose gaussian surface as a cylinder
Figure 11.9 A gaussian surface enclosing of radius r and height L coaxial with the charge
straight line charge distribution (Figure 11.9). The Gauss' law for the
cylinder is
L
ε 0 ∫ E.dS = ∫ λdl = λ L (11.21)
0
Splitting the closed integral into three parts

S
∫ E.dS = ∫ E.dS +
upper

lower
E.dS + ∫
lateral
E.dS (11.22)
base base surface

we see that only the third integral is not zero because the normals to the bases are perpendicular to
the field. Since the field on the lateral surface has constant magnitude and is normal to the surface
we can write

S
∫ E.dS = ∫ E.dS = E ∫ dS = 2π rLE
lateral lateral
(11.23)
surface surface

Substituting this result into Gauss' law we obtain


λ
E= (11.24)
2πε 0 r
In the case of a uniformly charged infinite cylinder with radius a, the field outside the cylinder
can be found following the same way as in the previous problem. If the radius of the gaussian
surface is greater than a the charge enclosed by it is also λL so that the final result remains the
same. That is, the electric field outside the
cylinder is the same as if all the charge were
concentrated along the axis. For a gaussian
surface within the cylinder, r < a, multiplying the
λL λ
volume charge density ρ = = and the
π a L π a2
2

volume V = πr2L, we find that the total charge


Figure 11.10 The variation of electric field enclosed by the surface becomes
with distance from the axis of an infinitely λ Lr 2
long uniformly charged cylinder q = ρ V = (11.25)
a2
PHYSICS 1 2002 S. Nitsolov 70
Then, by use of Gauss' law we obtain
λr
E= (11.26)
2πε 0 a 2
Figure 11.10 shows this result graphically.
If the charge is concentrated only on the surface of the cylinder, the field distribution outside the
cylinder remains the same. However, any gaussian surface within the cylinder does not contain
charge and therefore, the field inside the cylindrical shell of charge is zero.

11.4.4 SPHERICAL CHARGE DISTRIBUTION


Consider a spherically distributed charge Q enclosed by a sphere of radius a. By the spherical
symmetry of the problem, the field is radially directed and its magnitude at a point depends only on
the distance between the centre of the sphere and the point. A useful gaussian surface is a sphere of
radius r concentric with the charge distribution. Using Gauss' law and taking into account the
spherical symmetry, we have
ε 0 ∫ E.dS = ε 0 E ∫ dS = ε 0 E 4π r 3 = q (11.27)
where q is the net charge enclosed by the surface, or
q
E= (11.28)
4πε 0 r 2
Outside the sphere, r > a, q = Q and the field is given by
Q
E= (11.29)
4πε 0 r 2
We see that the electric field outside the spherically distributed charge is the same as if all the
charge were concentrated at the centre of the charge.
When r < a we shall consider the special cases
of spherical shell of uniform charge and sphere of
uniform charge. If all the charge is on the surface
of the sphere, then q = 0 and therefore the field
inside a shell of uniform charge is zero. In the case
of uniform sphere of charge the volume charge
Q
Figure 11.11 The variation of electric field density is ρ = so that the net charge
4π a 3
with distance from the centre of an uniformly enclosed by the gaussian surface is
charged sphere
Qr 3
q = ρV = 3 (11.30)
a
and the field is given by
Qr
E= (11.31)
4πε 0 a 3
The electric field of the uniformly charged sphere is illustrated in Figure 11.11.
PHYSICS 1 2002 S. Nitsolov 71

12 CONDUCTORS AND DIELECTRICS IN AN


ELECTRIC FIELD

12.1 INTRODUCTION
In the preceding chapter we discussed only electrostatic fields in vacuum without interesting in the
influence of the field on the matter. However, when a body is placed in an external electric field, the
electric charges of the body are subject of additional forces, which tend to move the charges from
their initial positions giving rise to a new charge distribution. This results in changes of the
properties of the body as well as of the net field. The response of the substance to an applied electric
field depends on the ability of its charged particles to move under the action of the field. Conductors
are materials, which contain many charged particles, capable to migrate within the boundaries of the
material, and which are therefore called free charges or mobile charge-carriers. For example,
metals, electrolytes and ionised gases are conductors. Insulators or dielectrics have no free charges.
Their charges are bound to specific atoms or molecules and can not move freely through the
medium. Semiconductors are intermediate between conductors and insulators.

12.2 CONDUCTOR IN AN ELECTRIC FIELD


Consider an isolated conductor in an electrostatic equilibrium, that is, there is no systematic drift of
its free charges. If an external electric field is applied on the conductor, its free charges are set in
motion in direction depending on the kind of their charge. For example, in metals, the only mobile
charge carriers are negative charged electrons moving in the direction opposite to the field. The free
charges move under action of the field until they reach the boundary of the conductor. Since the
charges can not leave the conductor they accumulate on its outer surface until the field they produce
neutralises the external field inside the conductor. Then the collective motion of charges stops and
the conductor reaches an electrostatic equilibrium again (Figure 12.1).
If we put an excess charge on initially
uncharged conductor, electrical forces of repulsion
will act on the free charges producing a
redistribution of the added charge over the outer
surface of the conductor. The charge will spread
over the surface until the internal electric field
vanishes and a steady state is reached. Thus, we
may conclude that, at electrostatic equilibrium, the
electric field is zero everywhere inside the
conductor. If this were not so, the forces due to the
Figure 12.1Metal plate immersed in a uniform field would produce a collective motion of the free
electric field charges. Likewise, at electrostatic equilibrium, the
electric field at the outer surface of the conductor
is normal to the surface. Since the electric field is zero inside the conductor, all points of the
conductor at electrical equilibrium have the same potential. Therefore, the surface of the conductor
is an equipotential surface. Since the field inside the conductor is zero, the flux through any closed
surface contained in the conductor is also zero. Then, according to Gauss' law, the net charge inside
any closed surface must also be zero. That is,
the excess electric charge placed on a conductor in equilibrium resides entirely on its outer
surface.
PHYSICS 1 2002 S. Nitsolov 72
Gauss' law allows to obtain a relation between the
magnitude E of the electric field and the surface
charge density σ at the surface of a conductor
(Figure 12.2 ). We choose a gaussian surface as a
small cylinder of base dS with a lateral surface
perpendicular to the surface of the conductor. One
side of the cylinder is just above the conductor and
the other is just bellow the surface. Since the
cylinder is small we may assume that the field and
surface charge distribution are uniform in the
cylinder. Taking into account that the field is
parallel to the lateral surface and its magnitude is
Figure 12.2 A cylindrical gaussian surface zero inside the conductor, we may write
embedded in the conductor ε 0 ∫ E.dS = ε 0 E.dS = ε 0 EdS = σ dS (12.1)
which gives
σ
E= (12.2)
ε0
The charge distributes itself unevenly on a
conductor surface of irregular shape. It may be
shown that the surface charge density of a
conductor at equilibrium increases as the radius of
curvature of the surface decreases. Therefore, the
electric field tends to be relatively high at sharp
points and edges and relatively low on plane
Figure 12.3 Electric lines of force at the regions (Figure 12.3).
conductor surface of irregular shape

12.3 DIELECTRIC IN AN ELECTRIC FIELD

12.3.1 POLAR AND NONPOLAR MOLECULES


The behaviour of a dielectric in an electric field may be explained in atomic terms. The molecules
of the dielectric materials can be divided into two groups, depending on their symmetry. In a polar
molecule the centre of negative charge does not coincide with the centre of the positive charge, and
therefore, the molecule has a permanent electric dipole moment. The simplest polar molecule
consists of two different atoms. For example, in the molecule of HF, the electron of the H atom
moves nearer to the F atom than to the H atom because the charge of the F nucleus is greater than
that of H. The dipole moment of the molecule is directed from the F atom to the H atom.
In the absence of the external electric field the
polar molecules are randomly oriented as a result
of the chaotic thermal motion (Figure 12.4 (a)) and
the substance does not exhibit any macroscopic
dipole moment. If an external electric field is
applied on the substance, the molecules experience
a torque that tends to orient their dipole moments
in the direction of the field. The alignment is not
perfect because of the thermal agitation (Figure
(a) (b) 12.4 (b)). The degree of alignment will increase as
Figure 12.4 (a) Polar molecules in the absence the field is increased or as the temperature is
of electric field. (b) Partial alignment of the decreased. When the field is removed the thermal
polar molecules in an external electric field. motion destroys the alignment of the molecules.
PHYSICS 1 2002 S. Nitsolov 73
Atoms and centrosymmetric molecules do not
have a permanent dipole moment because their
centers of positive and negative charge coincide.
However, in such molecules (called nonpolar) the
centers of positive and negative charge may be
separated under the action of an external electric
field. Then the molecule acquires an induced
(a) (b) electric dipole moment in the direction of the
Figure 12.5 (a) Nonpolar molecules in the applied field (Figure 12.5). Such partial charge
absence of electric field. (b) Polarization of the separation under the action of an external field is
nonpolar molecules in an external electric called polarization.
field.
12.3.2 POLARIZATION OF MATTER
We have seen that in the both cases of polar and
nonpolar molecules, the particles of the dielectric,
immersed in an external electric field (Figure
12.6), become electric dipoles, oriented in the
direction of the field. The effect of polarization of
nonpolar molecules as well as orientation of
permanent electric dipoles is to separate the
centres of the positive and negative charges of a
Figure 12.6 Polarization of a piece of dielectric certain piece of dielectric. The polarization of the
dielectric produces pile up of positive charge on
one side of the piece and of negative charge on the opposite side. The piece of dielectric becomes a
large electric dipole remaining neutral as a whole. In the dielectric polarization, the displacement of
the charges from their equilibrium positions is considerably less then atomic diameters. There is no
transfer of charges over macroscopic distances such as in the case of conductor. Some crystal
substances, known as ferroelectrics, exhibit permanent polarization in the absence of the external
electric field. It is associated with the spontaneous parallel orientation of the electric dipoles in
microscopic regions, called domains. The alignment in a domain results from the strong interaction
between polar molecules of the ferroelectric.

12.3.3 VECTOR OF POLARIZATION


The vector of polarization P at a point is defined as the electric dipole moment of a small region of
a medium, divided to the volume of the region. If the region ∆V contains N molecules, its dipole
moment is equal to the sum of the dipole moments of all molecules in ∆V. Therefore, the
polarization is given by
1
P=
∆V i
∑ pi (12.3)

If the mean dipole moment of a molecule is denoted by p, we may write


N
P= p = np (12.4)
∆V
where n is the number of particles per unit volume. In the case of an isotropic medium the
directions of the mean dipole moment and the external electric field coincide. In general their
directions are different.
For not very strong fields the polarization is proportional to the intensity of the field, so that,
P = ε0 χeE (12.5)
where the coefficient χe, called electric susceptibility, describes the response of the dielectric to the
action of the external field and depends on the properties of the molecules. The electric
susceptibility due to the orientation depends on the temperature (as T-1) because the thermal motion
disarranges the alignment of the molecules. The induced polarization of the molecules is related to
the internal structure of atoms and molecules and its contribution to the susceptibility is independent
PHYSICS 1 2002 S. Nitsolov 74
on the temperature.

The polarization and the charge density on the


surface of the polarized dielectric are closely
related. In order to find this relation we consider a
dielectric slab, placed in uniform external field
(Figure 12.7). The induced polarization of matter
is perpendicular to the surface of the slab. The
charges induced on the opposite surfaces of the
slab perpendicular to the field are q and -q so that
the induced dipole moment of the slab is lq. Since
Figure 12.7 A dielectric slab in an electric
the polarization is given by the dipole moment of
field
the slab lq, divided by its volume, we have
lq q
P= = =σ (12.6)
lS S
This result is a particular case of the more general relation
PN = σ (12.7)
where PN is the component of the polarization in the direction of the normal to the surface of the
dielectric. Obviously, the polarization and the surface charge density have the same unit, C⋅m2.

12.3.4 ELECTRIC DISPLACEMENT


Consider a dielectric slab placed in a uniform
electric field produced by two metal plates
carrying equal and opposite charges (Figure 12.8).
The charges on the surface of the dielectric with
density σpol, called polarization charges, must be
distinguished from the free charges with surface
density σfree. The electric field in the slab is a
superposition of two fields. According to the
relation of the field between two parallel planes
charged with equal but opposite charges and their
surface density, the field produced by the free
Figure 12.8 Electric fields in a dielectric due to
σ
the free and polarization charges charges has magnitude free and is directed from
ε0
the left plate to the right plate. The field due to the polarization charges has the opposite direction
σ pol
and its magnitude is . It partially neutralizes the external field of the free charges. Therefore,
ε0
the resultant field is given by
ε 0 E = σ free − σ pol (12.8)
Substituting the surface density σpol with the polarization we obtain
ε 0 E + P = σ free (12.9)
The electric displacement D at a point of the electric field is defined as
D = ε0E + P (12.10)
It is a convenient concept to reformulate the laws of electrostatics for dielectric media. It is
expressed in Cm-2. The relation between the electric displacement and the free charges on the
surface of the conductor has a general validity and may be written in the form
σ free = DN = D.u N (12.11)
That is, the component of electric displacement along the normal of the surface of a conductor is
equal to the surface density of the free charges on the conductor.
PHYSICS 1 2002 S. Nitsolov 75
12.3.5 GAUSS LAW
The total charge of the conductor can be expressed as
qfree = ∫ σ free dS = ∫ D.u N dS = ∫ D.dS (12.12)
S S S
This result can be written in the form
∫ D.dS = q
S
free (12.13)

It may be shown theoretically that this equation is valid in the general case regardless of the
presence of a conductor in the closed surface. Thus the Gauss' law may be reformulated as follows.
The flux of the electric displacement through a closed surface is equal to the net free charge
inside the surface.
In the case when the relation between the polarization and the electric field is linear we have
D = ε 0 E + P = ε 0 (1 + χ e )E = ε E (12.14)
where
ε = ε 0 (1 + χ e ) (12.15)
is called permitivity. The dimensionless quantity
εr = 1 + χe (12.16)
is called relative permitivity or dielectric constant. Thus, we may write Gauss' law in the form
ε 0 ∫ ε r E.dS = qfree (12.17)
S

For homogeneous medium (εr is constant) we have


ε ∫ E.dS = qfree (12.18)
S
It can be shown by use of this equation that the electric field and the potential of a point charge are
given by
qu
E= (12.19)
4πε r 2
q
V = (12.20)
4πε r

12.4 ELECTRIC CAPACITANCE


Suppose we put a charge q on an insulated neutral conductor. It may be shown experimentally and
theoretically that the potential V of the charged conducting body is directly proportional to the
charge q, that is,
q
V = (12.21)
C
where the coefficient
q
C= (12.22)
V
is called capacitance of the body. The unit for C is farad, F = C/V. The capacitance depends upon
the size and the shape of the conductor and the environment. For example, the potential of a
conducting sphere of radius a and charge q surrounded by dielectric is given by
q
V = (r > a) (12.23)
4πε r
so that
C = 4πε r (12.24)
ε
The dielectric materials rise the capacitance, relative to vacuum, by a factor of ε r = , since the
ε0
opposite charges induced on the dielectric surface adjacent to the conductor reduce the effective
PHYSICS 1 2002 S. Nitsolov 76
charge of the conductor. In order to calculate the capacitance of the Earth we may considered it as a
conducting body. Then Earth's capacitance is equal to 0.07 F.
If we place a charged conducting body near to a neutral conductor, the opposite electric charges
induced on the surface of the conductor neutralize partially the field of the charged body. As a
result, the potential as well as the capacity of the charged body increases. Thus, for a fixed charge,
the capacity is made large by bringing a second conductor very close to the first one. If the second
conductor has an opposite charge, its potential is also affected and the potential difference between
them is reduced. Such system of two insulated conductors designed for the storage of charges is
called capacitor. If the charges on the conductors are q and -q, the capacitance of the capacitor is
defined as
q
C= (12.25)
∆V
where ∆V is the potential difference between conductors. Capacitors are very useful in electrical
and electronic engineering.
A parallel-plate capacitor (Figure 12.9) consists
of two parallel metal plates of area S separated by
a distance d. The space between the plates is filled
by a dielectric of permitivity ε. The potential
difference between the plates is
σd
V1 − V2 = ∆V = (12.26)
ε
where ε0 is replaced by ε to take into account the
q
presence of dielectric. Since σ = , we obtain
S
Figure 12.9 A plane parallel capacitor. dq
∆V = (12.27)
εS
so that
εS
C= (12.28)
d

12.5 ENERGY OF THE ELECTRIC FIELD


A charged conductor has a certain potential energy Ep, equal to the work required to transfer the
charge from infinity to the conductor overcoming the repulsion of the charges already present. In
order to calculate the potential energy of the charged capacitor we consider the work dW needed to
add an infinitesimally small charge dq' to the conductor of capacitance C having charge q'. If the
potential of the conductor is V, the increase in potential energy dEp equal to the work dW is
q'
dE p = dW = Vdq' = d q' (12.29)
C
The total change in energy when the charge is increased from zero to q is
q
1 q2
E p = ∫ q ' dq ' = (12.30)
C0 2C
In the case of a charged capacitor of capacity C we consider the work needed to transfer an
infinitesimally small charge dq' from one plate to the other. This work is expended to overcome the
electric field between the plates, already charged by the opposite charges -q' and +q'. If the potential
difference is ∆V, the infinitesimal increase of energy is
q'
dE p = dW = Vdq' = d q' (12.31)
C
Integrating, we obtain the total change in energy
q
1 q2
E p = ∫ q ' dq ' = (12.32)
C0 2C
PHYSICS 1 2002 S. Nitsolov 77
Using the relation between the field and the surface charge density on a conductor and the
expression for the capacitance of the parallel-plate capacitor, we have
−1 2
q2 1  εS  1 σ  1 2
Ep = = (σ S ) 2   = ε   ( Sd ) = ε E V (12.33)
2C 2  D 2 S 2
where V is the volume between the plates. This result is a particular case of more general relation
between the potential energy of the system of charges and the electric field. It may be shown that
for an arbitrary charge distribution
1
E p = ∫ ε E dV
2
(12.34)
2
where the integral extends over all the space. It is reasonable to suppose that the potential energy of
a system of charges resides in the electric field around the charges so that in a given volume dV has
1 2
been stored an energy dE p = ε E dV . Then the energy density (energy per unit volume) is given
2
by
dE 1 2
= ε E (12.35)
dV 2
This equation is also valid for time-dependent fields.
PHYSICS 1 2002 S. Nitsolov 78

13 ELECTRICAL CURRENTS

13.1 INTRODUCTION
At electric equilibrium, the free electrons in a metallic conductor move randomly in all directions
like gas molecules. The number of electrons moving along each direction is equal, that is, they have
no net directed motion. The presence of an electric field inside the conductor gives rise to a drift of
free charges. This phenomenon having important practical application shall be considered in this
chapter.

13.2 CURRENT AND CURRENT DENSITY

13.2.1 INTENSITY OF AN ELECTRIC CURRENT


In general an electric current consists of flow of charged particles. This includes all kinds of
charged particles, such as free electrons in metals, electrons and holes in semiconductors, electrons
and ions in ionised gases, ions in electrolytes, charged particles in vacuum etc. The average
intensity of an electric current trough a given section is defined as
∆Q
I= (13.1)
∆t
where ∆Q is the net amount of charge past trough the section during the interval ∆t. The
instantaneous intensity may be defined as the limit of the average intensity as ∆t approaches zero,
that is,
dQ
I= (13.2)
dt
Then, the total charge past trough the section is given by the integral
Q = ∫ Idt (13.3)
In the SI the unit of current, ampere (A), is a fundamental unit in electricity. According to the
definition, the unit of charge, coulomb, may be expressed as A⋅s. The direction of an electric current
is defined as that of the directed motion of positive charges. It coincides with the direction of the
applied electric field that produces the current.

13.2.2 CURRENT DENSITY


A more detailed description of the current may be
given by the concept of electrical current density.
The current density is a vector characteristic of the
current in a point. If a current dI passes trough a
small area dS⊥ perpendicular to the flow velocity
of the charges, the current density is defined as
vector j of magnitude
dI
I= (13.4)
dS ⊥
Figure 13.1 Current density and current
through a small surface pointed in direction of the current. Then the
current dI through a plane surface dS which makes
an arbitrary angle with the current density (Figure 13.1) may be written as
dI = jdS⊥ = jdS cos θ = j.u N dS = j.dS (13.5)
where dS = u N dS . Thus, the electric current trough any surface S may be expressed by the flux of j
over the surface.
PHYSICS 1 2002 S. Nitsolov 79
I = ∫ dI = ∫ j.dS (13.6)
S S

13.2.3 EQUATION OF CONTINUITY


According to the principle of conservation of the electric charge, the current I = ∫ j.dS
S
out of a

closed surface S, enclosing a volume V, must be equal to the rate at which the charge Q inside the
surface decreases. That is,
dQ
I = ∫ j.dS = − (13.7)
S
dt
Representing the charge inside the surface as
Q = ∫ ρ dV (13.8)
V
we obtain
d
∫ j.dS = − dt ∫ ρ dV
S V
(13.9)

This is the equation of continuity of the electric charge. In the stationary case the amount of
charge inside any surface is time-independent. Therefore,
∫ j.dS = 0
S
(13.10)

The lines along which the free charges move do


not begin at one point and end at another, since at
these points the amount of charge would be time
dependent. Since these lines do not begin or end
they must be closed. Therefore, the steady current
flows along the closed path. Applying the equation
of continuity to the steady currents at a junction
point where several conductors meet (Figure 13.2),
we have

S
∫ j.dS = ∫ j.dS + ∫ j.dS + … = I1 + I 2 + … = ∑ Ii = 0
S1 S2 i

Figure 13.2 Steady currents at a junction point (13.11)


of several conductors. This equation is known as junction theorem or
first Kirchhoff's rule after the German physicist
Gustav Kirchhoff. It states that
the algebraic sum of the currents at a junction is zero.
Currents leaving the junction are taken as positive and those approaching the junction as negative. It
follows from the junction theorem that if the conductor has not branches, the current entering the
surface is equal to the current leaving the surface. Therefore, the current is the same in all parts of
the conductor.

13.2.4 RELATION BETWEEN THE CURRENT DENSITY AND DRIFT VELOCITY


The current density may be related to the
microscopic characteristics of the free charges. We
consider a current consisted of free charges, each
of charge q, moving with average drift velocity v
(Figure 13.3). The charge dq that passes through
the surfase dS in time dt is equal to the amount of
charge contained in a cylinder of base dS and
height vddt. The volume of the cylinder is
Figure 13.3 Drifting charges in a conductor. dV = dSvddt. Then
PHYSICS 1 2002 S. Nitsolov 80
dq = nqdV = nqdSvd dt (13.12)
and
dI dq
j= = = nqvd (13.13)
dS ⊥ dtdS ⊥
Since the current density and drift velocity are vectors, we may finally write
j = nqv d (13.14)
The current density at any point has the direction in which the positive charge at this point would
move.

13.3 OHM'S LAW

13.3.1 MACROSCOPIC STATEMENT


Georg Ohm discovered experimentally in 1826 a relation between the current and the potential
difference between two points of a conductor, known as Ohm's law. It states that
for a metallic conductor at constant temperature, the potential difference between two points of the
conductor is proportional to the electric current,
or
∆V = RI (13.15)
where ∆V is the potential difference, also called voltage, and the constant R is called the electric
resistance of the conductor between the two points. The unit of resistance, ohm, (W) is expressed as
V/A. A device used to provide a certain resistance to control the intensity of the current in electric
circuits is called resistor. It should be noted that many materials and devices do not obey Ohm's
law.

13.3.2 DIFFERENTIAL FORM OF OHM'S LAW


In order to find a relation between the microscopic
and macroscopic properties of a conductor we
consider a cylindrical conductor, shown in Figure
13.4. The current I through the cross-section S of
the conductor may be written as I = jS, where j is
the current density. Using Ohm's law we obtain
∆V
j= (13.16)
RS
Since the potential difference is related to the
Figure 13.4 Drifting charges in a conductor. electric field we have ∆V = El, where l is the
lenght of the cylinder, so that
l
j= E =σE (13.17)
RS
where
l
σ = (13.18)
RS
is called the electrical conductivity of the substance. The reciprocal of conductivity ρ = 1/σ is called
the resistivity. The result obtained above may be written in vector form as
j=σE (13.19)
This equation is called a differential form of Ohm's law. Substituting Eq. 13.19 into Eq. 13.14, we
obtain
σ
vd = E (13.20)
nq
This equation shows that if given material obeys Ohm's law, its free charges move with constant
drift velocity under the action of the applied electric field.
PHYSICS 1 2002 S. Nitsolov 81
13.4 CONDUCTION OF ELECTRICITY IN METALS
In order to relate the conductivity with the microscopic properties of the metal we shall adopt
so-called free electron model. According to this model, the free electrons behave in a manner
similar to the molecules of a gas. In the absence of an external field free electrons move randomly
with average thermal velocity vth. During the thermal motion electrons undergo continiously
inelastic collisions with the crystal lattice ions. The average distance l between two collisions is
called mean free path. Then, the mean time between two collisions is
λ
τ= (13.21)
vth
In the presense of an external electric field the electrons begin to drift slowly in direction opposite
to E. The velocity distribution of electrons is affected very slightly since their average drift velocity
vd is very much less then their thermal velocity by a factor of 1010 . Under the action of the electric
force an electron of mass m and charge e asquires an acceleration
F eE
a= = (13.22)
m m
In the mean time τ between two collisions the electron changes its velocity by aτ. In the collision it
loses its drift velocity and begins to accelerate again. The average drift velocity during each free
path is
aτ eEτ eE λ
vd = = = (13.23)
2 2m 2mvth
Combining this result with Eq. 13.20, we obtain
ne2 λ
σ = (13.24)
2mvth
This expression is called Drude's formula.

13.5 ELECTRIC POWER


A steady potential difference ∆V, applied between two points of a conductor causes a steady electric
current I. Maintaining the current requires energy because the electric field continiously does work
accelerating the free charges in the conductor. The work for transferring a charge Q through the
potential difference is
W = Q∆V (13.25)
Then the power required to maintain the current is
dW dQ
P= = ∆V (13.26)
dt dt
This is a general relation which gives the instantaneous power of the current passing through any
device. In particular, for conductors and devices that obey Ohn's law we have
(∆V )2
P = I ∆V = I ( RI ) = RI 2 = (13.27)
R
This equation is known as Joule's law. The SI unit of power is watt = J/s .
From microscopical point of view the transfer of energy is realized by drifting charge carriers.
For example in metals, the collisions between the free electrons and the crystal lattice increase the
vibrational energy of the lattice. In this way the electric potential energy transforms into internal
thermal energy of the material.
The power of an electric force F = qE acting on a free charge q drifting with average drift velocity
vd is
σ  1
P = F.v d = qE.v d = qE.  E  = σ E 2 (13.28)
 nq  n
Since an unit volume of the conductor contains n charges, the power p per unit volume may be
written as
PHYSICS 1 2002 S. Nitsolov 82
p = σ E2 (13.29)
This equation expresses Joule's law in a differential form.

13.6 ELECTROMOTIVE FORCE


In the previous sections we considered the relation
between a steady current and a potential difference
between two points of a conductor. The existence of
steady current implies that the conductor is
incorporated in a closed circuit, that is, a system of
conductors forming a continious closed path. Suppose
that a closed circuit (Figure 13.5) consists of N
components, connected in series so that the same
current I passes through each in turn. A unit of charge
Figure 13.5 A closed circuit. passing through the j - th component of resistance R
does work equal to the potential drop ∆Vj = RjI. The
total potential difference is
∆V = ∑ ∆Vi = ∑ Ri I = RI (13.30)
i i

where R = ∑ Ri is the total resistance of the circuit. In each passing through the circuit a unit of
i
charge does work RI, which is transferred into heat. Since this transfer is an irreversible process the
same amount of energy must be given to the charge to maintain the current. The total work E done
in moving a unit of charge around the circuit is called the electromotive force (abbr. emf). The
device which can supply energy to an electric current is called the seat of electromotive force.
In the circuit external to the seat of emf E (Figure
13.6) the current flows from point A of higher
potential to point B of lower potential and
transforms the electric energy into heat. However
inside the seat the flow of current is from point B
to A in direction of the increasing of the electric
potential. Obviously, the forces driving the free
charges against the electrostatic force in the seat of
Figure 13.6 A circuit with an electromotive emf are non-electrostatic. Furthermore, these
force. forces are non-conservative since their work along
the closed circuit is not zero. Thus, the seat
converts non-electrical energy (chemical,
electromagnetic etc.) to electrical energy and maintain a steady potential difference (and of course
an electric field) between its terminals. Note that the emf exists also if the circuit is open and the
current does not flow.
The total resistance R of a simple circuit, shown in
Figure 13.7 may be separated into the resistance of
the seat, Ri , called internal resistance, and the
resistance Re of the rest part of the circuit, called
external resistance, or R = Ri + Re . Thus, for a
single-loop closed circuit the Ohn's law may be
expressed as
E = ( Ri + Re ) I (13.31)
In general a closed electric circuit may consists of
Figure 13.7 A symbolic representation of a several junctions and branches connecting these
simple single-loop circuit. junctions (Figure 13.8).
PHYSICS 1 2002 S. Nitsolov 83
Let Rj be the resistance, Ej - the emf and Ij - the
current in the j - th branch. Then equalizing the net
emf around the loop to the total work done by the
field to transfer a unite charge we obtain
∑ Ej = ∑ R j I j
j j
(13.32)

This equation is called second Kircchoff's rule or


loop theorem. It states that
round any closed circuit the algebraic sum of
the emf is equal to the algebraic sum of
products of resistance and current.
Figure 13.8 A single-loop circuit, containing
several resistors and sources of emf.
PHYSICS 1 2002 S. Nitsolov 84

14 MAGNETIC INTERACTION

14.1 INTRODUCTION
The science of magnetism has its roots in the
observation that certain stones have the property of
attracting bits of iron. The name magnetism comes
from the district of Magnesia in Asia Minor where,
according to ancient Greeks, this phenomenon was
recognised firstly. A body having magnetic
properties is called magnet. There are two points
in each magnet, called poles, where the magnetism
appears to be concentrated. The Earth itself
behaves as an enormous magnet with poles
Figure 14.1 Interaction between magnets situated at the geographic poles. The Earth acts on
each magnet tending to orient it in such a way that
the same pole of the magnet points toward the North Pole. Each magnet has one north-seeking
(north or positive pole) and one south-seeking pole (south or negative pole). The experiment shows
that like poles repel and unlike poles attract each other (Figure 14.1).
Initially, the magnetic interaction was considered as an interaction between magnetic poles or
charges. However, the experience show that it is impossible to isolate a magnetic charge, that is, to
find a particle having only one kind of magnetism. Furthermore, numerous experiments have
demonstrated that the magnetic interaction is due to moving electric charges or electric currents.
The electric and magnetic interactions produced by the same set of charges are related and must be
considered as a manifestation of the more general electromagnetic interaction.

14.2 MAGNETIC FIELD

14.2.1 VECTOR OF MAGNETIC FIELD


Since magnets and currents exert magnetic forces
on each other we may say, in analogy with the
electrical interaction, that the magnets and current
carrying conductors give rise to a magnetic field in
the space around themselves. To define the
magnetic field quantitatively we consider a charge
q moving with velocity v (Figure 14.2). In the
absence of an electric field the charge does not
experience an electric force. However, the magnetic
field gives rise to another force F, called magnetic
Figure 14.2 Magnetic force on a moving force. The relation between the force, the charge
and its velocity may be found experimentally by measuring the force in different conditions. This
relation is expressed by the vector equation
F = qv × B (14.1)
where vector B is characteristic of the magnetic field at the point which may vary from point to
point but does not depend on the charge and its velocity. Thus the magnetic field can be represented
by the vector B, called the magnetic flux density, also known as magnetic induction, magnetic field
strength or simply magnetic field. The value of B may be found by measuring the values of F, q and
v. The three vectors are shown in Figure 14.2. The SI unit of magnetic field is tesla (abbr. T)
expressed as
PHYSICS 1 2002 S. Nitsolov 85
T = N C m s = N(ms ) s C = kgs C (14.2)
Note that the magnetic force is perpendicular to the plane containing vectors v and B. Therefore its
work is zero and it does not change the magnitude of the velocity and kinetic energy of the charged
particle. To find the direction of the force we may apply the right hand rule. It should be noted that
the principle of superposition is valid for a magnetic field.
In the presence of both electric and magnetic field the force on an electric charge is a vector sum
of the electric force q E and the magnetic force q v x B or
F = q (E + v × B) (14.3)
This expression is called the Lorentz force.

14.2.2 LINES OF FORCE AND FLUX OF THE MAGNETIC FIELD


The lines of force of the magnetic field may be introduced in the same way as in the case of electric
field. The tangent to a line of force at any point gives the direction of magnetic field B at that point.
The number of the lines of force passing through a perpendicular cross section of unit area is
proportional to the magnitude of B. As in the case of the electric field the magnetic flux dF passing
through a small area dS is defined by the equation
dΦ = B.dS (14.4)
For a surface S of a finite area the flux is given by
Φ = ∫ B.dS (14.5)
S

14.3 MOTION OF CHARGES IN A UNIFORM MAGNETIC FIELD


Consider a charged particle moving perpendicular to a uniform magnetic field, shown in Figure
14.3 where the field lines are perpendicular to the
page and are directed up, out of the page. The
expression for the magnetic force F = q ( E + v × B )
shows that the force has magnitude q v B and
direction in the plane of the figure, which means
that the particle can not leave this plane. Since the
force is perpendicular to the velocity it does not
change its magnitude. It may be shown that the
resultant motion is uniform along a circular path.
From Newton's second law we obtain
v2
qvB = ma = m (14.6)
r
Figure 14.3 A charged particle moving mv
perpendicular to a uniform magnetic field r= (14.7)
qB
which gives the radius of the path. The angular velocity ω of the circular motion is
v q
ω= = B (14.8)
r m
Since the motion of a positive charge is in clockwise direction, the direction of the angular velocity
is opposite to the field, that is,
q
ω=− B (14.9)
m
q
The angular velocity depends only on the relative charge of the particle and the magnetic field.
m
The frequency of the circular motion
qB
ν =− (14.10)
2π m
PHYSICS 1 2002 S. Nitsolov 86
is called cyclotron frequency.
If the velocity is not perpendicular to the field it
can be separated into component parallel to the field
and perpendicular to the field. The parallel
component does not give rise to a magnetic force and
does not change. Therefore the motion is
superposition of a uniform motion parallel to the field
and uniform circular motion in a plane perpendicular
to the field. The trajectory of the resultant motion is a
helix (Figure 14.4). If the magnetic field is not
uniform, the radius of the circular motion increases as
the field decreases and vice versa.
Figure 14.4 A charged particle moving in a
uniform magnetic field 14.4 THE HALL EFFECT
In 1879, the american physicist Edwin Herbert
Hall discovered an effect, called after him Hall
effect. This effect consists in the appearence of a
transversal electric field EH, called Hall field, in a
current carrying solid conductor when a
perpendicular magnetic field is applied (Figure
14.5). The direction of EH depends on the sign of
the charge carriers in the conductor. The Hall
Figure 14.5 Hall effect in a metal slab. effect is due to the magnetic force F = qv × B
acting on the charge carriers that constitute the
electric current.
To explain the Hall effect, consider a slab of
metal through which an electric current flows
placed in a magnetic field which is perpendicular
to the current. In metals, the electric current
generally is composed of electrons so that the flow
of negative charge carriers is in the opposite
direction to the conventional current (Figure 14.6
(a) (a)). Since the magnetic field is directed up, out of
the page, the average magnetic force F = qvd × B
is directed downward according to the right hand
rule. The electrons collect along the lower side of
the slab, which thus becomes negatively charged.
The concentration of electrons at the opposite side
decreases and it becomes positively charged. The
charge separation gives rise to an electric field EH,
which exerts a force qEH on the electrons in the
(b) opposite direction to the magnetic force. At
Figure 14.6 Hall effect due to (a) positive and equilibrium the two forces are equal in magnitude,
(b) negative charge carriers.
that is, qEH = qvd B or
EH = vd B (14.11)
The potential difference ∆VH, called Hall voltage, between the opposite sides of the slab due to EH
is given by ∆VH = aEH or
∆VH = avd B (14.12)
I
Taking into account that j = nqv d (Eq. 13.14) and j = where S = ab is the cross-section of the
S
conductor we may write
PHYSICS 1 2002 S. Nitsolov 87
I
vd = (14.13)
qnab
Substituting this result into the expression for ∆VH we obtain
IB
∆VH = (14.14)
qnb
The result is also valid for positive charge carriers (Figure 14.6 (b)). Then the potential difference is
the reverse of that in the case of negative carriers. If the both types of charge carriers are present,
the direction of Hall voltage indicates which type predominates. Thus, the sign of charge carriers
for solids can be determined experimentally by the measurement of the Hall voltage.

14.5 MAGNETIC FORCE ON AN ELECTRIC CURRENT


A current in a conductor is a directed motion of electric charges. Since a magnetic field exerts a
forces on moving charges, the conductor in a magnetic field experiences a force equal to the
resultant of the magnetic forces on the free charges. Figure 14.7 shows a current-carrying conductor
in a magnetic field. Since the average force on each
charge q drifting with velocity vd is qvd × B the total
force dF acting on a small volume dV of the
conductor is
dF = nqv d × BdV (14.15)
where n is the number of free charges per unit
volume. Taking into account that j = nqv d we
obtain
dF = j × BdV (14.16)
Integrating this result we obtain the total force F on
Figure 14.7 The magnertic force on a unit a finite volume V of a conductor,
volume of a conductor F = ∫ fdV = ∫ j × BdV (14.17)
V V
To find an expression for the force acting on a small segment of current-carrying filament of length
dl and cross section S we rewrite the product jdV in the form
jdV = jSdl = Iu l dl = Idl (14.18)
where dl = uldl and ul is a unit vector which points along the wire in direction of the current. Then
the force on the segment can be written as

dF = j × BdV = Idl × B (14.19)

and the total force on a wire L is expressed by


F = ∫ dF = I ∫ dl × B (14.20)
L L

If the magnetic field is uniform, the force on a


filament of arbitrary shape (Figure 14.8) is
 
F = I  ∫ dl  × B = IR × B (14.21)
L 
where R = ∫ dl is a sum of all vector elements of
L
the curve L which is equal to the vector joining
the ends of the filament. Therefore the force on
Figure 14.8 A magnetic force on a current the current-carrying filament does not depend on
carrying filament in a uniform magnetic field. its shape.
PHYSICS 1 2002 S. Nitsolov 88
14.6 MAGNETIC TORQUE ON A CURRENT LOOP.
Consider a current carrying rectangular loop of
length a and width b placed in a uniform magnetic
field. (Figure 14.9). Let R1, R2, R3, R4 be the sides
of the rectangle oriented in the direction of the
current I (R1 = R3 = a and R2 = R4 = b), F1, F2, F3,
F4 the forces acting on the corresponding sides and
uN the normal to the plane of the loop oriented
according to the right hand rule. The loop is
oriented relative to the field so that the sides R1
and R3 are perpendicular to the field and the
normal uN makes an angle θ with the field. Using
the the relation F = IR × B (Eq. 14.21) we may
(a) calculate the forces acting on the four sides of the
loop. Since R3 = - R1 and R4 = - R2 the forces on
the of opposite sides are equal in magnitude and
opposite in direction. The resultant of all the forces
on the loop is zero. The net torque of F2 and F4 is
zero because they have the same line of action.
The forces F1 and F3 constitute a couple whose
torque is
1  1 
(b) top view τ = R 2 × F3 +  − R 2  × F1 = − R 2 × F1 (14.22)
Figure 14.9 A rectangular current loop in 2  2 
magnetic field. Taking into account that F = IR × B we have
τ = − R 2 × ( IR1 × B ) = − I ( R 2 .B ) R1 (14.23)
Since the angle between R2 and B is θ + π/2 we obtain
τ = IR1 R2 B sin θ (14.24)
Denoting by S = R1R2 the area of the rectangular enclosed by the loop we define a vector
M = ISu N (14.25)
called the magnetic moment of the current. Then the torque may be written as
τ = MB sin θ (14.26)
or, in vector form
τ = M×B (14.27)
The torque tends to orient the loop perpendicular to the magnetic field. It can be shown that this
expression for the torque is valid for a plane current loop of arbitrary shape. The behaviour of the
loop in the magnetic field is similar to that of the electric dipole in an electric field. Because of that,
we may calculate the potential energy of the loop, replacing p with M in the expression for the
potential energy of the electric dipole. The result is
Eπ = −M.B (14.28)
If the vectors B and M are parallel, the potential energy is minimum and the magnetic dipole is at
state of stable equilibrium.
PHYSICS 1 2002 S. Nitsolov 89

15 SOURCES OF MAGNETIC FIELDS

15.1 INTRODUCTION
In the previous chapter we discussed the effect of a magnetic field on moving charges and currents
and defined the vector B. This chapter deals with the generation of a magnetic field by moving
charges or, equivalently, currents. We consider the Biot-Savart law, Ampere's law and Gauss law
for magnetism which describe the magnetic field set up by moving charges and current-carrying
conductors. If the current distributions are known, these laws can be used to compute the
correspondent magnetic fields.

15.2 MAGNETIC FIELD OF A MOVING CHARGE


It was found experimentally that when a charged
particle moves relative to an observer it produces a
magnetic field in addition to its electric field
(Figure 15.1). The magnetic field B at a point at a
distance r from a charge q moving with a velocity
v may be expressed in the form
qv × u
B∝ (15.1)
r2
Figure 15.1 r
where u = is a unit vector in direction of r. In
r
µ0
the SI the constant of proportionality is , where

µ 0 = 4π × 10 −7 T −1 m ⋅ A (15.2)
is called the vacuum permeability or simply magnetic constant. Thus,
µ qv × u
B= 0 2 (15.3)
4π r
Strictly speaking, this equation is valid only for velocity, which is small, compared with the
velocity of light. The lines of force of B are concentric circles around the tangent to the path of the
qu
charged particle. Taking into account that the electric field of the charge is given by E =
4ε 0π r 2
(Eq. 10.9), we may write
1
B = µ0ε 0 v × E = v×E (15.4)
c2
where
c = 2.99792458 × 108 m / s (15.5)
is the velocity of light. Two observers moving with different velocities relative to an electric charge
measure equal electric fields but different magnetic fields. In particular an observer at rest relative
to the charge measures only an electric field.
PHYSICS 1 2002 S. Nitsolov 90
15.3 MAGNETIC FIELD OF CURRENTS

15.3.1 THE BIOT-SAVART LAW


In 1820 the Danish scientist Hans Christian
Oersted (1777-1851) discovered that an electric
current deflects a compass needle and concluded
that the current produces a magnetic field. Since a
current in a conductor is a flow of free charges the
magnetic field produced by the current is equal to
the resultant of the magnetic fields produced by
Figure 15.2
the individual moving charges. To determine the
magnetic field produced by a current element we
µ 0 qv × u
start from the expression B = (Eq. 15.3) for the magnetic field of a moving charge in a
4π r 2
conductor. The contribution dB of the elementary volume d V of the conductor to the magnetic field
can be expressed as a sum of the contributions of the moving charges in the volume. Since the
number of the charges is ndV (n is the number charge density) and they move with velocity vd
µ nqv d × u
dB = 0 dV (15.6)
4π r 2
Taking into account that j = nqv d (Eq. 13.14) we have
µ j× u
dB = 0 2 dV (15.7)
4π r
For a small segment of length dl of a filament L carrying current I (Figure 15.2) this equation may
be transformed using the expression jdV = Idl (Eq. 14.18) into
µ dl × u
dB = I 0 2 (15.8)
4π r
where dl = uldl and ul is a unit vector which points along the axes of the segment in direction of the
current. This relationship was discovered by the French physicists Jean-Baptiste Biot and Félix
Savart in 1820 and is now known as the Biot-Savart law. The lines of force are concentric circles
around the current element dl. We may find the total magnetic field B produced by the conductor L
by integrating the contributions of all current elements, that is,
µ I dl × u
B = 0 I∫ 2 (15.9)
4π L r

15.3.2 MAGNETIC FIELD OF A CIRCULAR CURRENT


Figure 15.3 shows a magnetic field produced by a
circular current I of radius a. The magnetic lines of
force are closed curves which link the current. The
magnetic field has axial symmetry. We shall
calculate the magnetic field along the axes of the
circular current.
Let P be a point at a distance R from the centre C
of the current (Figure 15.4). Consider the field dB
produced by the current element dl. Since the
Figure 15.3 Lines of force of the magnetic vector u is perpendicular to dl the magnitude of
field produced by a circular current loop the vector dB is
PHYSICS 1 2002 S. Nitsolov 91
µ 0 I dl
dB = (15.10)
4π r 2
and its direction is perpendicular to the plane
formed by dl and P. Resolving dB into two
components, dB|| and dB⊥, which are parallel and
perpendicular to the axis, respectively, we may see
that only dB|| contributes to the total magnetic field
at P. The component dB⊥ is cancelled in the
integral by the normal component of the magnetic
Figure 15.4 Components of the magnetic field field produced by the current element directly
at point on the axis of a circular cirrent loop opposed to dl. Therefore, the total field is parallel
to the axes and
B = ∫ dB|| (15.11)
L
Substituting dB|| = dBcosα into the integral we may write
µ I cos α µ Ia cos α
B = ∫ dB|| = cos α ∫ dB = 0 ∫ dl = 0 2 (15.12)
L L
4π r 2
L
2r
a
Taking into account that cos α = and r = a 2 + R 2 we have
r
µ 0 Ia 2
B= (15.13)
2 ( a2 + R2 )
32

Since the magnetic moment of the current is M = I S, we obtain finally


µ0 M
B= (15.14)
2π ( a 2 + R 2 )
32

The magnetic field at a large distance R >> a from the loop is


µ M
B= 0 3 (15.15)
2π R
This result is similar to the expression for the electric field of an electric dipole. Because of that we
may regard the current loop as magnetic dipole.

15.3.3 MAGNETIC FIELD OF A RECTILINEAR CURRENT


Consider a long rectilinear current-carrying
filament (Figure 15.5). The vector B at an arbitrary
point P is perpendicular to the plane determined by
the filament and P and it is directed downward, to
the page. Then, we can write
µ I sin θ µ I ∞ sin θ
B = ∫ dB = 0 ∫ 2 dl = 0 ∫ 2 dl (15.16)
L
4π L r 4π −∞ r
Figure 15.5 A long rectilinear current carrying From the figure we see that r = R and
filament. sin θ
R
l = − R cot gθ . Substituting r, l and dl = dθ
sin 2 θ
we obtain
µ I π sin θ sin 2 θ R µ0 I π µ I
B= 0 ∫ d θ = ∫ sin θ dθ = 0 (15.17)
4π 0 R 2
sin θ
2
4π R 0 2π R
We see that the magnetic field is inversely proportional to the distance R. The lines of force are
circles concentric with the current and perpendicular to it. Introducing a unit vector uT tangent to
the line of force we may write the magnetic field as
PHYSICS 1 2002 S. Nitsolov 92
µ0 I
B= uT (15.18)
2π R
Using this expression we can determine the forces
between two parallel currents. Consider two
parallel rectilinear conductors separated by a
distance R and carrying currents I1 and I2 (Figure
15.6). At an arbitrary point of the first conductor,
the magnetic field B2 produced by the current I2 is
given by
µ I µ I
B 2 = 0 2 uT = 0 2 u × u 21 (15.19)
Figure 15.6 2π R 2π R
where u21 is a unit vector parallel to R and directed
from the first conductor to the second one. Then the force acting on the piece of length l of the first
conductor is
µ IIl
F21 = I1lu × B 2 = 0 1 2 u × ( u × u 21 ) (15.20)
2π R
Taking into account that
u × ( u × u 21 ) = −u 21 = u12 (15.21)
we may write
µ IIl
F21 = 0 1 2 u12 (15.22)
2π R
Exchanging the subscripts 1 and 2 we obtain the force on the second conductor
µ IIl
F12 = 0 1 2 u 21 (15.23)
2π R
F F
The forces per unit length f 21 = 21 and f12 = 12 are given by
l l
µ 0 I1 I 2 µ II
f 21 = u12 f12 = 0 1 2 u 21 (15.24)
2π R 2π R

The result indicates that when the currents are in the same direction, the conductors attract each
other and when they are in opposite directions, they repel. The two forces obey the third Newton's
law.
The force of interaction between two current carrying conductors is used to define the SI unit of
current. A current of one ampere is that constant current which flowing in two infinitely long
conductors, of negligible circular cross-section, placed 1 metre apart in a vacuum, would produce a
force between them of 2×10-7 N per metre length of conductor.

15.4 AMPERE'S LAW


Applying the Bio-Savart law we calculated the
magnetic field of an infinite rectilinear current. We
shall use this result to find another relationship
equivalent to the Bio-Savart law which is more useful
in the case of current distributions of certain
symmetry. Consider a closed curve L around a long
straight conductor carrying current I (Figure 15.7).
We shall calculate a line integral ∫ B.dl around the
L

Figure 15.7 A long straight conductor carrying curve L. The product B.dl may be transformed as
curren I enclosed by a plane closed curve L.
PHYSICS 1 2002 S. Nitsolov 93
µ 0 Ir µ I
B.dl = BdlT = dω = 0 dω (15.25)
2π r 2π
Substituting the scalar product into the line integral we obtain
µ 0 I 2π
∫L B .d l =
2π ∫0
dω = µ0 I (15.26)

It can be shown that this result is valid for any shape of the current and for any shape of the curve.
If the curve L links several conductors carrying currents I1, I2, ..., I1n, using the principle of
superposition we can write

L
∫ B.dl = µ0 ∑ Ii
i
(15.27)

This means that


the circulation of the magnetic field along a closed curve is equal to the total current linked by
the curve.
This relation is called Ampere's law in honour of the French physicist André-Marie Ampère
(1775 - 1836). The direction of the currents can be specified using the right-hand rule. The positive
direction of the current is indicated by the extended thumb of the right hand if the fingers curl in the
same sense as the path of integration.

15.5 GAUSS' LAW


We shall use the expression for the magnetic field
of an infinite rectilinear current to find a relation
analogous to Gauss's law in electrostatics. We
shall calculate the magnetic flux through a closed
surface in the special case of a magnetic field
produced by an infinite rectilinear current. Let S be
the closed surface formed by a circular cylinder
coaxial to the current and by two planes normal to
the axes (Figure 15.8). The magnetic flux through
Figure 15.8 A gaussian surface enclosing a the cylinder is
rectilinear infinite current. Φ = ∫ B.dS (15.28)
S
Taking into account that the magnetic lines of force are circles concentric with the current we see
that B lies in the surface at every point. Therefore, the product B.dS = 0 everywhere over the
surface and
Φ = ∫ B.dS = 0 (15.29)
S
It may be shown that this result, known as Gauss' law of magnetism, is valid in the general case of
arbitrary magnetic field and of arbitrary closed surface. Since the flux through a surface is
proportional to the number of lines of force passing through the surface we may conclude that the
number of lines entering the closed surface is equal to the number of lines leaving the surface.
Therefore, the magnetic lines of force can never start and can never stop. The difference between
the electric lines of force and the magnetic ones may be explained by the absence of magnetic
charges on which lines of B could start and end.
PHYSICS 1 2002 S. Nitsolov 94
15.6 MAGNETIC FIELD OF A SOLENOID

15.6.1 TOROIDAL COIL


Consider a toroidal coil carrying current I (Figure
15.9). The coil consists of N turns spaced very
closely at equal distances. Because of the
symmetry we expect the magnetic lines of force to
be circles concentric with torus. According to
Ampere's law, the magnetic circulation along an
arbitrarily chosen line of force L' within the torus
is
Figure 15.9 A toroidal coil carrying current I. ∫ B.dl = ∫ Bdl = B ∫ dl = BL ' = µ0 NI (15.30)
L' L' L'
If the cross-section of the torus is small compared
with its radius, we may assume that L' is approximately equal to the length L of the thorus and write
B = µ 0 nI (15.31)
N
where n = is the number of turns per unit length. Since the total current linking with any line of
L
force outside the torus is zero the magnetic circulation must also be zero. Therefore, B is zero
everywhere outside the toroidal coil.

15.6.2 INFINITELY LONG COIL


An infinitely long coil (Figure 15.10) may be
considered as a toroidal one of very large radius.
Therefore, the magnetic field is entirely confined
to the interior of the coil and is uniform. Its
magnitude inside the coil is given by
Figure 15.10 An infinitely long solenoid.
B = µ 0 nI (15.32)
PHYSICS 1 2002 S. Nitsolov 95

16 MAGNETIZATION OF MATTER

16.1 INTRODUCTION
We have been considering the magnetic effects in the absence of matter neglecting the conductors
that produce the magnetic field. We now begin our consideration of the role of matter. The presence
of matter modifies the electric field as well as the magnetic field. Magnetic properties are exhibited
to a greater or smaller extent by all substances. To explain the magnetic properties of materials we
will start with consideration of the magnetic properties of the atoms which compose them.

16.2 MAGNETIC MOMENTS OF ELECTRONS AND ATOMS


Atom consists of positive nucleus and negative electrons, which move around the nucleus. The
magnetic properties of each individual atom are due mostly to the electrons. There are two ways in
which electrons generate magnetic field.

16.2.1 ORBITAL MAGNETIC MOMENT


The motion of an electron along a closed orbit is equivalent to a circular current loop, which
produces a magnetic dipole moment, that is, the orbiting electrons in atoms can be treated as tiny
magnetic dipoles. If the electron moves uniformly in a closed circular orbit of radius r and period T,
the equivalent current of a such motion is
q qv
I= = (16.1)
T 2π r
where v is the velocity. Then the correspondent magnetic moment of the electron is
qv 1
M = IS = π r 2 = qvr (16.2)
2π r 2
Since the angular momentum of the electron is L = mvr, we obtain
q
M = L (16.3)
2m
We see that the orbiting electron posesses a magnetic moment associated with its orbital angular
momentum.

16.2.2 SPIN MAGNETIC MOMENT


Each electron, in addition to its orbital motion around the nucleus, has a rotational or spinning
motion around its own axes. If the electron is spinning, there is a charge in motion and thus an
electric current. Because of that, the electron also possesses a magnetic moment associated with its
spin angular momentum. The spin magnetic moment is an intrinsic part of the magnetic moment of
an electron. It exists even the electron is at rest.

16.2.3 ATOMIC MAGNETIC MOMENT


Atomic nuclei also have magnetic moments. The nuclear magnetic moments are much smaller (103
times) than the electron magnetic moments. The net magnetic moment of an atom or molecule
results from contributions of its components. Since the contribution from atomic nuclei is negligibly
small, the total magnetic moment is due mostly to the electron orbital and spin magnetic moments.
The magnitudes of molecular and atomic magnetic moments are influenced by the tendency of
electrons to form pairs in which the orbital and spin moments are antiparallel and equal in
magnitude. The result is that nearly all atoms and molecules with even number of electrons have no
magnetic moment. If the number of electrons is odd. the magnetic moment is due in most cases to
the last unpaired electron.
PHYSICS 1 2002 S. Nitsolov 96
16.3 MAGNETISATION
Usually, matter in bulk does not exhibit collective magnetic moment in the absence of external
magnetic field. However, the presence of an external magnetic field as in the case of electric
polarisation gives rise to a macroscopic magnetic moment, that is, the substance exhibits magnetic
polarisation or magnetisation in the presence of a magnetic field. There exist several different
mechanisms of magnetisation. Substances can be divided into several classes depending on the
mechanism of their magnetisation by an external magnetic field.

16.3.1 DIAMAGNETISM
When matter is placed in an external magnetic field the electronic motion is distorted. As in the case
of electric polarisation of unpolar molecules the magnetic field gives rise to an additional magnetic
dipole moment, whose magnitude is proportional to the field. However, unlike the induced electric
momenta the induced magnetic momenta of atoms and molecules are antiparallel to the external
field. This effect, called diamagnetism, may be explained by the tendency of the electron to move
in a circular path in a magnetic field. The circular motion induced by the magnetic field is
superimposed on the normal motion of the electron. Such circular motion is equivalent to an
elementary current and produces an additional magnetic moment with direction antiparallel to the
external field and magnitude which does not depend on the kind and on the orientation of the atom.
Diamagnetism is a universal property of all substances inasmuch as they consist of electrons
although sometimes it is obscured by the stronger effects of magnetisation. Only those substances in
which diamagnetism is the only effect of magnetisation are referred to as diamagnetic. This
condition is satisfied by atoms and molecules which have not permanent magnetic moment.
Since diamagnetism is due to the distorsion of the motion of electrons in atoms and
molecules, it is independent on the temperature.

16.3.2 PARAMAGNETISM
If a substance whose atoms or molecules have permanent magnetic dipole moments is not subjected
to an external magnetic field, the magnetic dipoles are oriented randomly and there is no
macroscopic magnetic moment. If an external magnetic field is applied, it exerts torques on the
magnetic dipols which tend to align them in the direction of the field. The orientation of the
molecular dipoles is complete only at very low temperatures since the thermal motion tends to
disarrange the molecules. Since the extent to which the magnetic dipoles become aligned depends
strongly on the thermal motion paranagnetism is dependent on the temperature. The net magnetic
dipole moment is oriented in the direction of the external field. Such type of magnetisation is called
paramagnetism. Since the permanent dipole moments are much greater than the induced moments
the diamagnetism is masked by the paramagnetic effects in paramagnetic substances.
There also exists paramagnetism resulting from the alignment of the spin magnetic moments of
the conducting electrons in metals.

16.4 FERROMAGNETISM

Certain metals, alloys, and compounds of the transition (iron, cobalt, nickel), rare-earth
(gadolinium, dysprosium) and actinide elements can have a large permanent magnetisation. This
property, called ferromagnetism, exists below certain temperature, called Curie temperature
(named after the French physicist Pierre Curie). When the temperature of a ferromagnetic material
exceeds Curie temperature it behaves as if it were paramagnetic.

16.4.1 DOMAINS
Ferromagnetism is due to a special kind of interaction between the electronic spins of two adjacent
atoms, called exchange interaction, which may be described only in terms of quantum mechanics.
This interaction gives rise to a spontaneous parallel or antiparallel alignment of the unpaired spin
magnetic moments among adjacent atoms in the cristal lattice. As a consequence, an alignment of
PHYSICS 1 2002 S. Nitsolov 97
magnetic moments of adjacent atoms in one direction takes place in small regilons, called domains.
Each domain contains a great number of atoms (1017÷1021 ). The direction of magnetisation of a
domain is one of the crystallographic directions of easy magnetisation (the directions in a crystal in
which a given magnetic field produces maximum polarisation).

(a) B = 0 (b) B ≠ 0 (c) B ≠ 0


Figure 16.1 Domains in feromagnetic materials

If a paramagnetic is unmagnetized, the domains are randomly oriented (Figure 15.10). When an
external magnetic field is applied to the paramagnetic, the domains that are suitably oriented grow
at the expense of the others. If the external field increases, the magnetisation of the domains turns
aligning in the direction of the field. The magnetisation remains even when the magnetic field is
removed.
The thermal motion of the atoms randomizes their orientation and decreases the alignment.
When the temperature of a ferromagnetic exceeds Curie temperature, the thermal motion complitely
destroys the alignment of spins. Then the domains disappear, the substance looses its ferromagnetic
properties and becomes a paramagnetic.

16.4.2 CLASSIFICATION OF FERROMAGNETIC


MATERIALS
In ferromagnetic materials the adjacent atomic
(a) ferromagnetism moments are parallel. The exchange interaction can
produce antiparallel alignement as well. Then, If the
adjacent atomic moments are equal in magnitude, the
(b) antiferromagnetism net magnetisation is zero (Figure 16.2). In this case
the magnetisation is called antiferromagnetism.
Ferrimagnetism is a magnetisation in which
adjacent atomic magnetic moments are also
(a) ferrimagnetism antiparallel but their magnitudes are different, as
shown in Figure 16.2. The magnetisation in such
Figure 16.2 Alignment of adjacent atomic substances, called ferrits, is smaller than in the case
moments of ferromagnetic.

16.5 MAGNETIZATION VECTOR


We have seen that a piece of matter placed in a magnetic field becomes magnetised. The
magnetisation vector M is defined as a magnetic moment of the material per unit volume. If m is
the average magnetic moment of each atom or molecule and n is the number of atoms per unit
volume, the magnetisation is
M = nm (16.4)
-3 2 -1
The unit of magnetisation is m Am = Am . To describe the field of a magnetised body we shall
consider a cylindrical piece of matter of length l and cross-sectional area S magnetised uniformly in
the direction of the axis of the cylinder (Figure 16.3).
Since the microscopic magnetic dipoles are parallel to the axis of the cylinder the associated
elementary circular currents are oriented perpendicular to the axis. Inside each cross-section of the
cylinder perpendicular to the axis, the adjacent elementary currents are in opposite directions and
PHYSICS 1 2002 S. Nitsolov 98
cancel each other. The elementary currents are
l
not neutralised only on the surface of the
Imag
cylinder. Thus, we see that the magnetic field
produced by the magnetisation of the cylinder is
S
equivalent to that of a very long solenoid carrying
a current Imag per unit length. The total magnetic
mcyl
moment of the cylinder of volume Sl is
mcyl = M ( SI ) = ( Ml ) S (16.5)
Front view Since the magnetic moment of the effective
Figure 16.3 Magnetic dipole moment of a magnetisation current is
uniformly magnetised cylinder. IS = ( I mag L) S (16.6)
equalising the two expressions, we obtain
I mag = M (16.7)
It can be shown that this result is valid in the general case of an arbitrary shape of the material and
arbitrary distribution of the magnetisation. That is, the effective magnetisation current per unit
length on the surface of the magnetised body is equal to the component of the magnetisation tangent
to the surface of the body and is directed perpendicular to M. It should be noted that the effective
magnetisation current is not zero inside the magnetised body in the general case of non-uniform
magnetisation.

16.6 THE MAGNETIZING FIELD


We have seen that a magnetised body has effective magnetisation currents produced by bound
electrons, which can not move freely through the substance. A magnetic field my also be produces
by electric currents due to free charges and which are therefore called free currents. In general, the
magnetic field is a superposition of the fields produced by free currents and by currents due to
bound charges.
D C Consider a cylindrical piece of matter placed in
an uniform magnetic field B produced by a very
long solenoid of n turns per unit length carrying
A B current I (Figure 16.4). The applied magnetic field
M gives rise to a magnetisation M which is
equivalent to an effective surface current Imag = M.
B Thus, we may replace the system of the cylinder
and solenoid by a single very long solenoid
carrying total current per unit length equal to
Figure 16.4 A piece of magnetised material nI + Imag. The magnetic field B produced by this
inside a long solenoid. solenoidal current is parallel to the axis of the
solenoid. Its magnitude is given by
B = µ 0 (nI + I mag ) = µ 0 (nI + M ) (16.8)
This relation between the free current I, the magnetic field B in the medium and the magnetisation
M of the medium may be written in the form
H = nI (16.9)
where the vector
1
H= B−M (16.10)
µ0
is called the magnetising field. It is expressed in A/m.
To compute the circulation of the magnetising field ∫ H .dl around a free current, we take a
L
rectangular path ABCD, shown in the Figure 16.4. Since the sides BC, CD and DA do not
contribute the circulation, we have
PHYSICS 1 2002 S. Nitsolov 99


ABCD
H .dl = H.AB = H AB = nI AB = I free (16.11)

where Ifree is the total free current across the rectangle ABCD. Although this result was obtained for
a particular geometrical arrangement, it is of general validity,

L
∫ H.dl = I free (16.12)

Therefore,
the circulation of the magnetising field along a closed curve is equal to the total free current
linked by the curve.
In this equation all currents due to the magnetisation are excluded, while, in Eq. 15.27, the total
current linked by the closed curve, includes the free currents as well as the magnetisation currents.
The relation between H and B may be written in the form
B = µ0 (H + M ) (16.13)
For non-ferromagnetic substances the magnetisation M of a body is proportional to the magnetising
field H or
M = χmH (16.14)
where the quantity χm is called the magnetic susceptibility of the material. χm is dimensionless, and
has no unit. For diamagnetic substances χm is negative and for paramagnetic substances is positive.
For both diamagnetic and paramagnetic materials
χ m << 1 (16.15)
Substituting the equation for M into the relation between H and B, we obtain
B = µ 0 ( H + χ m H ) = µ 0 (1 + χ m ) H = µ H (16.16)
where µ = µ0 (1 + χm ) is called the permeability of the substance. It has the same units as µ0 . The
dimensionless quantity
µ
µr = = 1 + χm (16.17)
µ0
is called the relative permeability.
Using the relation B = m0 H, we transform the equation for the circulation of H into
1
∫L µ B.dl = I free (16.18)

For a homogeneous medium µ = const. so that


∫ B.dl = µ I free
L
(16.19)

We see that the effect of the matter on the magnetic field may be taken into account by replacing µ0
by µ.

16.7 ENERGY OF THE MAGNETIC FIELD


mv 2
A moving charged particle of mass m and charge q has a kinetic energy Ek = as well as
2
energy Emag of its magnetic field B. It can be shown that the energy of the magnetic field is given by
the relationship
1
Emag =
2µ V∫ B 2 dV (16.20)

where the integration is over all the volume in which the magnetic field exists. The energy density
of magnetic field (magnetic energy per unit volume) may be defined as
dEmag B 2
= (16.21)
dV 2µ

S-ar putea să vă placă și