P. G. Francis (Auth.) - Mathematics For Chemis
P. G. Francis (Auth.) - Mathematics For Chemis
P. G. Francis (Auth.) - Mathematics For Chemis
Mathematics
for Chemists
P. G. Francis
Department of Chemistry,
University of Hull
Francis, P.O.
Mathematics for chemists.
I. Chemistry- Mathematics
I. Title
510'.2454 QD42
Preface ix
2 Differential calculus 29
2.1 Significance and notation 29
2.2 The calculus limit 30
2.3 Differentiation of simple functions 34
2.4 The use of differentials; implicit differentiation 36
2.5 Logarithms and exponentials 38
2.6 The chain rule and differentiation by substitution 39
2.7 Turning points: maxima, minima and points of
inflection 41
2.8 Maxima and minima subject to constraint; Lagrange's
method of undetermined multipliers 45
vi Contents
2.9 Series 50
2.9.1 Geometric series 51
2.9.2 Power series and Taylor's theorem 51
2.9.3 Maclaurin's theorem 54
2.9.4 Inversion of a series 58
2.9.5 Empirical curve fitting by power series 60
2.10 The evaluation of limits by L'Hopital's rule 61
2.11 The principles of Newtonian mechanics 63
4 Integration 98
4.1 Significance and notation 98
4.2 Standard methods of integration 101
4.2.1 Simple functions 101
4.2.2 Reciprocal functions 101
4.2.3 Integration by parts 102
4.2.4 Integration by substitution 104
4.2.5 Expansion of algebraic functions 104
4.2.6 Integration by partial fractions 105
4.3 Standard forms of integral and numerical methods 106
4.4 Multiple integration 107
4.5 Differentiation of integrals; Leibnitz's theorem 108
4.6 The Euler-Maclaurin Theorem 109
Contents vii
5 Applications of integration 115
5.1 Plane area 115
5.2 Plane elements of area 117
5.3 Elements of volume; polar coordinates in three
dimensions 120
5.4 Line integrals 122
5.5 Curve length by integration 123
5.6 Applications of multiple integration 124
5.6.1 The pressure of a perfect gas 124
5.6.2 Interactions in a real fluid 127
5.7 The calculus of variations 129
5.8 Generalized dynamics 132
Index 190
Preface
and for a given angle, this has a numerical value that is independent of
any conventional scale of measurement. The circumference of a
complete circle of radius r being 27tr, the angle corresponding to a full
rotation is 2nr/r = 2n natural units of angle, this unit being called the
radian. We also have the conventional definition of such a full rotation
2 Algebraic and geometrical methods
as 360 degrees (360°), so that
1t radian = 180°.
The definition of angle in two dimensions as (arc length)/radius can
be extended into three dimensions; if we take a conical segment from a
sphere we can define a three-dimensional solid angle as
solid angle = (area of segment of spherical surface)/(radius)2.
(1.1 b)
For a complete sphere of surface area 41tr 2,
complete solid angle = 41tr2/r2 = 41t,
and since this expression does not contain r it is also a pure number that
is independent of the size of the sphere. The unit of solid angle is called a
steradian. It is a useful geometrical concept that (area)/(radius)2 is
constant for a given solid angle, and this is used in Section 5.7.
Our second natural number, e, arises in a different way. It is useful to
divide the mathematical methods used in science into two categories,
geometrical methods and analytical methods. The above definition of 1t
is geometrical and belongs in the field of diagrams and models. The
second category is algebraic and numerical, and does not depend on
our ability to construct models and diagrams. Analytical methods are
developed by strictly logical deductions from first principles, called
axioms. This is generally the more powerful of the two mathematical
methods and, as such, is the one preferred by mathematicians and is
often the one needed to solve the more difficult problems. Geometrical
methods, on the other hand, often seem simpler but are sometimes
deceptively so. The wise rule is to use geometrical methods with care
and to defer to analytical methods when in doubt.
This division into geometrical methods and analytical methods is
arbitrary since they are parts of the same whole and a combination of
the two is frequently used in practice. Sometimes the connection
between the two methods is not obvious, an example being provided by
the trigonometric functions sin (J, cos (J, tan (J, .•.. The geometrical
definition of sin (J is
. (J opposite side
SIn = -"'-=----- in a right-angled triangle, (1.2)
hypotenuse
whereas the analytical definition is
(J3 (J5 (J7
sin (J = (J - 3! + 5! -71 + .... (1.3)
Natural numbers 3
We can see that these two definitions are the same in a particular case
by taking the angle eto be, say, 90° = nl2 radian. Then sin e = 1 by the
geometrical definition, and if we put e = nl2 = 1.5708 into the series
(1.3) it is easily shown that successive terms become rapidly smaller and
smaller, and that the sum tends towards a limit of 1.000.
The fact that the two definitions of sin e are always the same can be
proved only by means of extensive analytical argument. One way is to
use the definition of cos e as the series
e2 e4 (J6
cos (J = 1 - 2! + 4! - 6! + .... (1.4)
This is then the derivative (Chapter 2) of sin (J. The derivative is used to
define the slope of the tangent to a curve, and the angle between two
straight lines can be obtained as the inverse of the cosine series. In
this way, geometry can be developed by the analytical approach
leading eventually to showing that (1.2) and (1.3) are equivalent to
each other.
As mentioned above, our second natural number, e, belongs in
analytical methods. It is defined by the series
x2 x3
e = 1 + x + 2! + 3! + (1.5)
so that when x = 1,
1 1
e = 1 + 1 + 2! + 3! + ... = 2.71828 . . . . . (1.6)
This arises naturally in science because e, and only eX, gives exactly the
same quantity when it is differentiated (Chapter 2). This corresponds to
the quite common situation where the rate of change of a quantity is
proportional to the quantity itself, as in first-order reaction kinetics.
More fundamentally, exponential relations arise as a result of the
simple probability that determines the Boltzmann distribution which
underlies many physical phenomena (Section 2.8).
Putting x = 0 in (1.5) gives eO = 1. This is a particular case of the
general rule that any quantity raised to the power of zero is unity, which
follows from the laws of indices:
and
so that
and when x = y,
4 Algebraic and geometrical methods
1.2 Units and dimensional analysis
The units used in the SI system are given in the appendix together with
numerical values of the fundamental physical constants and conversion
factors. Notice that the symbols for units are written only in the
singular, or they cannot be cleanly cancelled and we would also have
absurdities such as 0.99 g but 1.01 gms. This need not conflict with
colloquial usage; we may speak of a temperature difference of
10 kelvins but this is written as 10 K.
Algebraic symbols are used to denote physical quantities, such as p
for pressure; the symbol denotes the quantity, not the units. A
particular value of the quantity is then the product of a pure number
and its units, and this product follows the normal rules of algebra. Thus
2 atm is the product of the pure number 2 and the unit atm, so that if
p = 2 atm we have p/atm = 2, or a quantity divided by its units is a pure
number. An important particular case is when taking logarithms; since
the definition of a logarithm only applies to pure numbers, the value of
any quantity must first be divided by its units, such as In(p/atm). The
value obtained will then depend on the units being used since, for
example,
In (p/atm) = In(p/101.325 kPa)
= In(p/kPa) -In(101.32S)
= In(p/kPa) - 4.618.
The dimensions of a physical quantity are defined as the appropriate
combination of powers of the fundamentals mass (M), length (L) and
time (T); thus velocity has dimensions LT- 1 and force (mass
x acceleration) has dimensions MLT - 2. The dimensions are precisely
analogous to the SI units of kilogram, metre and second, so that force
has units of kg m S-2.
Dimensional analysis is a method of checking and predicting
relations between physical quantities based simply on the principle that
the dimensions must balance in any equation. A standard example of
the technique is to predict the form for Stokes's law for the drag on a
sphere moving in a viscous fluid. Since
force = mass x acceleration
it has dimensions MLT - 2. The viscosity '1 of a fluid is defined by
force per unit area
'1 = velocity gradient
force distance
= - - x --:--::--
area velocity
Units and dimensional analysis 5
with dimensions
M~ 1 ;t
T'l., x j22 x~ x L = ML -IT-I.
We then assume that the drag (resisting force) on the sphere will depend
on the size of the sphere (radius a), its velocity (v) and on the viscosity of
the fluid. Dependence on temperature will be taken care of through
change in the viscosity of the fluid. The relation that is assumed is
written as
and dimensional analysis will enable us to find the numbers oe, p and y.
The assumed relation is rewritten in terms of the dimensions (or units)
of the quantities, thus
MLT- 2 = L"(LT-I)P(ML -IT-I)y.
Finally, we equate powers of each of the primary dimensions M, Land
T to obtain
for M, y = 1,
for L, oe + p-y = 1,
for T, -p-y = -2,
so that oc = p = y = 1 and the required equation is
force = ka"v.
Physical, rather than dimensional, arguments are needed to show that
the proportionality constant k is 61t.
The same principle can be applied when using the practical SI units
of kilogram (kg), metre (m) and second (s), so that in any equation
connecting physical variables the units must cancel on the two sides of
the equation. This has two valuable and practical uses: unless the units
cancel, the equation being used is wrong; and when we wish to change
to other units, we can do so by a simple algebraic method called
quantity calculus. The latter is based on the principle that multipli-
cation by unity leaves any quantity unchanged, and we can construct
'unity brackets' by writing, as a fraction, the new units over the old
using appropriate conversion factors. As a simple example, if we write
(1 week/7 days) this has the value unity. To convert, say, 42 days into
weeks we multiply by the appropriate unity bracket and cancel the
units:
1 week
42 days = 42.eays- x 7..days" = 6 weeks.
6 Algebraic and geometrical methods
Example 1.1
The gas constant R in the equation p V = nRT has the units needed for
the equation
R=pV
nT
Example 1.2
Show that Planck's constant has the dimensions of momentum
x length.
From the appendix,
h = 6.624 X 10- 34 J s.
The units of momentum x length are those of mass x velocity x length.
The unity brackets are based on the conversion factors
11= INm,
1 N = 1 kgms- 2 ,
so that the units of momentum x length are
kg x sm x m = )g'~ )4's'i. J
- , - x ~}If x ~}Ii = Js
Units and dimensional analysis 7
Example 1.3
Suppose that we are given that the spacing between the lines in the
microwave spectrum of HeI is 20.7 cm- 1 and that we need to calculate
the moment of inertia I, but can remember only that the spacing is 2B
where B is either h 2 /8n 2 I or h/8n 2 I c. This difficulty can be resolved by
adopting the strongly recommended rule that we substitute into
formulae not just numerical values but the units as well.
We have that either
h cm
or
I = 8n 2 c x 10.35'
where h == 6.626 x 10 - 34 J sand c = 2.998 X 108 m s - 1, so that either
1=
6.626 x 10- 34,J',8'.,.Qftl$ }(m kgm
x--x--x---
.m
8n 2 x 2.998 x 10 .m x 10.35
8 .J N~ 102..ern
Example 1.4
Show that a diver at a depth of66 ft of sea water (density 1.03 gcm -3) is
under a pressure of about 3 atm.
The pressure exerted by a column ofliquid of density p and height h
is given by
(x
b )2 b2 C b2 - 4ac
+ 2a = 4a2 - ~ = -4-a-O;2;--
We now take square roots of both sides, recognizing that this
introduces ambiguity of sign, so that
X = - -b ± 1.J
- (b2 -4ac). (1.11)
2a 2a
Corresponding formulae can be written for the solution to a cubic
equation, but these are too cumbersome to be useful; it is better to find a
10 Algebraic and geometrical methods
numerical factor by trial and error and to use the formula (1.11) for the
remaining quadratic factor. Newton's method (Section 2.9.2) can be
used in the initial numerical approximation.
Real solutions of a quadratic equation will not exist when b2 < 4ac.
We can then use complex numbers to write solutions containing real
and imaginary parts as in Section 1.10.
Example 1.5
In physical applications we may expect physical conditions to restrict
the choice of roots. Thus the dissociation constant K of a solution of a
monobasic weak acid is related to its degree of dissociation tX at
concentration c by
so that
ctX 2 + KtX - K = 0,
tX= - -K ± -1 V/ (K 2 + 4Kc).
2c 2c
But K, c and tX are all positive, so that only the positive sign before the
root has physical meaning (see also Example 2.25).
I
I
I
I
I
~'-----;N :
I I
IL.. _ _ _ _ _ _ _ ---II
Fig. 1.1
B C
~------------------~
a
Fig. 1.2
1.7.3 Three-dimensional geometry
Many problems in science, and particularly in chemistry, involve more
than two variables. These correspond to diagrams in three or more
16 Algebraic and geometrical methods
dimensions. They are often treated by taking two-dimensonal sections
in planes perpendicular to each other. When lines or planes are
perpendicular to each other they are called orthogonal.
In Fig. 1.3 the point P is distance r from origin O. The coordinates of
Pare (a,b,c), these being the projections of the line OP on to the x-, y-
and z-axes. This is conveniently shown by drawing perpendicular, or
normal, lines onto the z-axis and into the (x, y) plane. The projection
ON in the (x, y) plane then gives projections OA = a and OB = b on
the x- and y-axes.
z
c
A-~~~~------~---y
B
A~ ____________ ~
x
Fig. 1.3
~+~+~ = 1, (1.13)
abc
as may be most easily checked by noting that the partial derivatives are
then constants (Chapter 3). Equation (1.13) is that of a plane through
the points A, Band C in Fig. 1.3, as can be seen by putting y = z = to
show that the intercept on the x-axis is x = a, and so on.
°
The direction of a plane in space is often required in crystallography,
when it is denoted by the Miller indices (h, k, I). These are the smallest
whole-number ratios of the reciprocals of the intercepts on the axes.
Thus the plane through A, Band C in Fig. 1.3 is denoted by
(l/a, l/b, l/c), multiplied by a common factor if necessary so as to give
the smallest whole numbers. The use of reciprocals here avoids infinite
values because a plane parallel to an axis is defined to have a
corresponding Miller index of zero.
Example 1.7
The potential energy V for simple harmonic motion (Section 6.3) is
given by V = tkx 2. Substituting V for y gives a = kj2 and b = h( = 0),
so that ab = h2 (= 0) and the curve is a parabola (curve F in Fig. 1.4).
There are standard forms of equation for each kind of curve. These
are
(a) Circle x 2 + y2 = r2, (1.15)
(b) Ellipse x 2ja. 2 + y2jp2 = 1, (1.16)
(c) Parabola y2 = 4a.x, (1.17)
(d) Hyperbola x 2ja. 2 _ y2jp2 = 1. (1.18)
Example 1.8
Show that these standard forms satisfy the general conditions stated
above.
----+-~------r-~~~_+------+_~---x
Fig. 1.4
Example 1.9
The equation
x 2 + y2 -2x+4y = 4
represents a circle because h = 0 and a = b. The presence of terms in x
and in y means that the centre of the circle is not at the origin. The
equation is reduced to standard form by moving the origin to eliminate
20 Algebraic and geometrical methods
these linear terms; the term in x can be obtained by using
(x - 1)2 = X 2 - 2x + 1
and that in y from
(y+2)2 = y2+4y+4.
We can therefore write the equation as
(x-l)2+(y+2)2 =9
so that it is a circle with centre at (1, - 2) and radius 3.
Example 1.10
Identify the curve
y = ocx(l-x)
and suggest two alternative linear forms.
This equation has the form
ocx 2 - ocx + y = 0
so that ab = h 2 and the curve is a parabola. Differentiation (Section 2.8)
shows the curve to have a maximum at x = 1/2, y = oc/4, and also y = 0
at both x = 0 and x = 1. The equation is reduced to standard form by
writing
x 2 - X + y/oc = 0,
and using
(x - 1/2)2 = x 2 - X + 1/4
we obtain
(x-l/2f = 1/4- y/oc
= -(y-oc/4)/oc.
This becomes standard form by interchanging the variables and
moving the origin to (1/2, oc/4).
A linear plot may be obtained either by plotting y against x(1 - x),
which is obvious by inspection, or from yagainst (x -1/2)2.
Such a relation is often a good approximation to the properties of
mixtures, such as the heat and volume changes on mixing liquids or on
mixing gases. Departure from symmetry about x = 1/2 can be allowed
for by using a series of terms in powers of (x - 1/2). A similar approach
Some geometrical methods 21
is used to allow for anharmonicity of molecular vibrations in
spectroscopy.
Example 1.11
Show that, neglecting air resistance and the curvature of the earth, a
projectile moves in a parabola under the action of gravity.
Given velocity of projection u at an angle (X to the horizontal, the
vertical component of the initial velocity is u sin (X. The vertical height y
at time t is given by
y = ut sin (X _!gt 2 •
gx 2
y=xtan(X-22 2'
U cos (X
Fig. 1.5
1.9 Probability
We express the probability of an event as a fraction between 0 and 1,
where zero means that the event will not happen and unity means that
the event is certain. A fundamental principle is that the probability of
an event is proportional to the number of ways in which that event can
occur. Thus in throwing two dice simultaneously we may obtain a total
of5 in four ways; by scoring any of (4, 1), (1,4), (2, 3)or (3, 2). Each of the
dice may fall in anyone of six ways, and for any particular score on one
we have six possibilities for the other; the total number of possible
results is therefore found by multiplying 6 by 6 to give 36. The
probability of scoring 5 is therefore 4/36 = 1/9.
This argument applies to independent events, which are common in
scientific applications. An important quantity is the number of ways of
arranging N objects in line. For this we may choose the first as anyone
of the objects, and so in N ways assuming them to be distinguishable.
The second is chosen from the remaining N - 1, and so in N - 1 ways.
Since each choice of the first may be followed by anyone of the choices
of the second we multiply N(N -1) to find the number of ways of
choosing the first two in the line. By repeating the argument we see that
the required number is NL
This is then extended to find the number of ways of placing N objects
into boxes that contain nl, n2, ... in successive boxes. We begin by
placing the objects in line in N! ways, and then divide the line into
groups of n 1 , n2, . .. . The objects in a group of nl may be rearranged
24 Algebraic and geometrical methods
in n 1 ! ways and these have so far been included in the total of N!. Since
these rearrangements are not to be distinguished, we must remove them
from the product by dividing by nl!, hence the required number is
N!/(nl ! n2! ... ). If the objects are placed at random into the boxes, the
probability of obtaining particular numbers in the boxes is then
proportional to this factor; this is used in Example 2.17.
Example 1.12
The equilibrium constant for the reaction
H2 +D2 = 2HD
may be calculated from simple probability theory. We suppose that a
catalyst surface, on which dissociation into atoms occurs, is exposed to
a gas mixture containing equal numbers ofR and D atoms. We assume
that molecules are formed by random pairing of atoms on desorption
from the surface.We then have equal probability of any particular atom
being H or D. We may form H2 (or D 2) only by both atoms being the
same, but HD can be formed in two ways; either one site on the catalyst
may be occupied by H and the other by D, or the other way round. We
therefore form twice as many HD molecules as H2 or D 2 , giving for the
equilibrium constant
,,
,Q Real axis
I
,
'A*(o-ib)
Fig. 1.6
Example 1.13
cos 2 y+sin 2 y = (cosy+isiny)(cosy-isiny)
= eiYe- iy = eO = 1.
28 Algebraic and geometrical methods
Addition or subtraction of (1.31) from (1.30) gives the useful
substitutions
eix+e- ix . eix -e- ix
cos x = ----:2:--- and SIn x = ---:-2-:-i- (1.33)
Example 1.14
Use equations (1.33) to show that cos 2 x - sin 2 x = cos 2x.
Differential calculus
y = '(x)
I
I
I
6x - __ ..JI
r------
I
I I
Fig. 2.1
y=t
Fig. 2.2
Fig. 2.3 is drawn to illustrate other terms that are used to describe the
behaviour offunctions. The straight line y = f(x) = x has the property
The calculus limit 33
Fig. 2.3
thatf( -x) = -f(x), meaning that changing the sign of x changes the
sign of y, and this is called an odd function of x. On the other hand, y
= x 2 isa parabola (Section 1.7.4) for whichf( -x) = f(x);changing the
sign of x does not change the sign of y and this is called an even function
ofx.
Most of the simpler functions used in physical applications are
continuous, differentiable and single-valued, and the methods of
calculus can then be used. A function such as y = l/x is 'well behaved'
except as x -+ 0, so that such functions can be used within certain
ranges, which may have to be specified.
Having performed the operation of differentiation once we can do so
again on the result of the first differentiation, assuming that also to be
differentiable. This gives the second derivative of y with respect to x,
which is written as d 2 y/dx 2 , and we can go on to higher derivatives
d 3 y/dx 3 , .••• A straight line will have the same slope at any point on
the line so that the first derivative is a constant; this means that the
second derivative, which measures the rate of change of slope with
change in x, is zero. The second derivative is thus seen to measure
curvature. When the second derivative d 2 y/dx 2 is positive, the slope is
increasing with increase in x and the curve lies above its tangent as in
34 Differential calculus
Fig. 2.1; such a curve is called concave upwards. Conversely, a curve
with decreasing slope is called concave downwards.
The notation for higher derivatives can be explained by regard-
ing dy/dx as the product of the differential operator d/dx and the
variable y. Successive differentiation is then denoted by
(d/dx)2 times y = d 2y/dx 2.
Example 2.2
Put y = x", then
(y + <5y) = (x + <5x)"
= (x + <5xHx + <5x)(x + <5x) ..•
= x" + nx"-l<5x + terms in (<5X)2, etc.
Hence
dy r (y + <5y) - Y
dx = ,,~~o <5x
= lim (nx"-l + terms in <5x, etc.).
",,-+0
When we take the limit all terms except the first contain <5x and so tend
Differentiation of simple functions 35
to zero, giving
dy d
_ = _(x") = nx"-l. (2.5)
dx dx
The formulae for differentiating a product or quotient may be
derived from these first principles. If y = uv where u and v are both
functions of x
y + l5y = (u + l5u)(v + I5v)
where l5u and I5v are the increments in u and v produced by an increment
I5x in x. Hence
dy = lim (U + l5u)(v + I5v) - uv)
dx b-+O I5x
. (I5U
= lIm I5v I5Ul5v)
v-+u-+--.
6,x-+0 I5x I5x I5x
Now the limit of a sum is the sum of the limits, and that of a product is
the product of the limits. The first two terms contain the definitions of
du/dx and dv/dx and the third gives
( I·1m
6,x-+0
l.:)(
uU I·1m -I5v)
I5x
6,x-+0
x
then
dy 1
- = 1 - -2
dx x '
Example 2.4
Show that implicit differentiation of xy - x 2 = 1 gives the same result
as Example 2.3.
Differentiating the expression gives
xdy+ ydx -2xdx = 0
xdy = (2x - y)dx
dy 2x - Y 2X2 -xy x 2 + (x 2 -xy)
dx X x2 x2
But x 2 -xy = -1 so that, as before,
dy = 1-~
dx x2 •
Example 2.5
The rate of change of vapour pressure p of a liquid with temperature T
is related to the heat of vaporization !l.H by
dp !l.H
dT= RT2P.
This can be transformed by using both dp/p = d(lnp) and
d(I/T) = - (I/T2)dT into
d (In p) dpT2 !l.H
d(l/T) = -p dT = -R·
Example 2.6
Show that if the rate coefficient k of a chemical reaction depends on
temperature T according to
k=Ae- E / RT
where A and E are constants, then
dInk E
d(I/T) = - R
Example 2.7
Show that
Example 2.8
Let
y = (2X2 _1)3.
Example 2.9
Let y = eax , where a is a constant. Substitute
u = ax,
then
du/dx =a
Turning points: maxima, minima and points of inflection 41
and
y = e", dy/du = e".
Hence
dy dydu
- = - - = ae" = ae DX •
dx dudx
In general
if y = e/(x) then dy = e/(X)f'(x)dx. (2.20)
Example 2.10
Find the derivative with respect to temperature T of the expression
Z = 1/(e hv /kT -1).
Substitute
u = (e hv /kT -1)
then
du _~ehv/kT
dT kT2
and
1 dz 1 1
Z=-, - u2
u du (e hv /kT _1)2
Hence
dz dzdu hv e hv /kT
dT dudT kT2 (ehv/kT _1)2 .
x
Fig. 2.4
Example 2.11
Given y = x4+4,
dy = 4x3 = 0 at x = 0,
dx
d 2y = 12x2 = 0 at x = 0,
dx 2
d3y
-=24x=0 at x = 0,
dx 3
d4y
dx 4 = 24.
Hence the fourth derivative is the first one to be non-zero at the turning
point x = O. Since n is even and the derivative is positive the point
x = 0, y = 4 is a minimum.
Example 2.12
Show that the curve In y = x 3 + 1 cuts the y-axis at a horizontal point of
inflection.
Differentiation of the equation gives
dy
- = 3x 2 dx,
y
dy
dx = 3x 2 y.
The second derivative may be obtained either by differentiating the
equation a second time to give
44 Differential calculus
or by differentiation of the expression for dyldx to give, as before,
d2 y d (dY) dy
dx2 = dx dx = 6xy + 3x 2 dx = 6xy + 9x 4 y,
and similarly
Example 2.13
The velocity v of a chemical reaction is given by
v = kl(a _X)2 + k2x(a -x),
where kl' k2 and a are positive constants, x is variable. Find the
condition for a maximum in velocity, and the relation between kl and
k2 such that this occurs at x = a14.
The required condition for a maximum is dvldx = 0, d 2vldx 2
negative, where
dv
- = -2kda-x)+k 2(a-2x)
dx
=0
So that
kl a-2x
for maximum velocity.
k2 = 2(a-x)
Put x = a14, then
kl _ al2 _ 1
k2 - 3al2 - 3
for the maximum to be at x = i.
This is the equation for a reaction which is second order in the reactant
together with catalysis by-product.
Maxima and minima subject to constraint 45
Example 2.14
Find the ratio of height to diameter for a cylinder so that the surface
area is a minimum for a given volume.
Put V = 7tr 2 h and A = 27trh + 27tr2. We have two simultaneous
equations to satisfy, that V is constant and A is a minimum. This may be
solved in various ways. We could insert Vas a constant and eliminate h
by writing
to give
A = 2V/r+27tr2
as an equation in the variables A and r, and use the condition
dA/dr = o. A neater solution is obtained by differentiating the
expressions both for V and for A and putting d V = 0 for constant
volume and dA = 0 for minimum surface area. Thus
dV = 27trh dr + 7tr2 dh
r
=0 if dr =
dh 2h
and
dA = 27th dr + 27tr dh + 47tr dr
=0 if :~ = - h : 2r
Example 2.15
Find the minimum distance from the origin to the curve y = 2 - x 2 •
The first equation is obtained by expressing the distance r from the
origin to any point (x, y) as
r2 = x2 + y2
and minimizing by differentiating and putting dr = 0, to give
2rdr = 2xdx+2ydy = 0
so that
xdx+ydy=O. (2.21)
The second equation is the constraint that the point (x, y) must lie on
the curve. We express that also in differential form by writing the
equation of the curve with zero on the right-hand side and differentiat-
ing. Thus
y_2+X2 = 0,
dy+2xdx = O. (2.22)
We solve these two simultaneous equations in the usual way by
multiplying equation (2.22) by a constant factor IX (the Lagrange
multiplier) and adding to equation (2.21). This gives
xdx+ydy=O
and
2lXxdx+lXdy = 0
so that
x(l + 21X)dx + (y + lX)dy = O. (2.23)
Maxima and minima subject to constraint 47
We now choose ex so as to eliminate one of the differential variables,
such as by putting
(1 +2ex) = 0, ex = -1/2.
The other term in (2.23) must then also be zero, so that
(y+ex)dy = (y-t)dy = 0
and, since dy may take any value, y = t. The equation of the curve then
gives
x2 = 2- y =!
and the required distance is obtained from
Example 2.16
Find the shape of a right circular cone having minimum surface area for
a given volume.
If the cone has base radius r, vertical height h and slant height a, we
have a 2 = r2 + h 2 and the volume V and surface area A are given by
V = nr 2hl3
A = nr2 + nra.
For minimum A, dA = 0 giving
2nr dr + nr da + na dr = 0,
(2r+a)dr+rda = O. (2.24)
For a given volume, d V = 0 giving
t nrh dr + !nr2 dh = 0,
2hdr+rdh = O. (2.25)
48 Differential calculus
and from r2 + h 2 - a 2 = 0
rdr+hdh-ada = O. (2.26)
We now take (2.24) + ex (2.25) + P(2.26) to give
(2r + a + 2exh + pr)dr + (r - pa)da + (exr + fJh)dh = O.
Choosing ex and p to eliminate the terms in da and dh, we put
r-pa = 0, p = ria
and
exr+ph = 0, ex = - phlr = -hla,
so that
2r+a-2h2Ia+r2Ia = O.
But h2 = a 2 - r2, so that
3r 2 + 2ra - a 2 = (3r - a) (r + a) = 0,
and since only positive values have physical significance, a = 3r.
An algebraic solution of the equations is possible but Lagrange's
method gives the neater solution.
The most probable arrangement is the one that can occur in the
maximum number of ways; this is when dP = 0 or, more usefully, when
d In P = dP / P = 0
so that
d InP = d InN! -d Innl! -d Inn2! - ... = o. (2.28)
If it were simply a matter of placing objects at random into boxes the
most probable arrangement would be an equal number in each box.
The problem is complicated by constraints, that the total number of
particles is fixed and that the total energy is also fixed. These constraints
give
N = nl + n2 + . . . + ni + ... = constant (2.29)
and since the ni particles in the state with energy 8i contribute ni8i to the
total energy,
E = nl81 + n282 + ... + ni8i + ... = constant. (2.30)
The variables in this problem are the numbers nl, n2' ... , n;, ....
Since we have three simultaneous equations, all but two of these
variables will be independent. We apply Lagrange's method by adding
(2.28) to at times the differential form of (2.29) and p times the
differential form of (2.30). The numbers nl, n2, ... that we seek are
average values, and so are treated as continuous variables rather than
integers. Hence
dN = dnl +dn2 + ... +dni+ ... = 0 (2.31)
and
dE = 81 dnl + 82 dn2 + . . . + 8i dni + ... = o. (2.32)
Before doing the addition, we modify (2.28) by use of Stirling's
50 Differential calculus
approximation, equation (1.23)
In N! = N In N - N for large N,
giving
d InN! = In NdN.
Also, since N is constant, dN = 0, so that the term in N disappears from
(2.28). This leaves
(2.33)
We now add (2.33) + (X(2.31) + P(2.32) to obtain
(In nl + (X + peddn 1 + (In n2 + (X + pe2)dn2 + ...
+ (In ni + (X + pe;) dn i + . . . = O. (2.34)
We now choose the multipliers (X and p so as to eliminate n 1 and n2'
Equation (2.34) is then a sum of terms, each term containing only one of
the independent variables ni so that in order for the sum to be zero, each
term must be zero. Hence
(In ni + (X + pei)dni = 0,
2.9 Series
The use of series provides a valuable extension to algebraic methods.
We can use a convergent series of terms to represent quantities that
cannot be expressed in a simpler closed form, such as 1t, e or sin 0, and
also problems that cannot be solved by simple algebra often yield to a
solution in terms of series. The subject of series arises at this stage in our
discussion because the series with most physical significance can be
based upon Taylor's theorem, which is discussed in the next section.
A mathematical series contains terms that are related to each other,
so that a formula exists for calculating any particular term. Such series
are useful in practice only when successive terms become smaller and
smaller so that the series converges. We may then use the series either to
calculate its sum or as a source of successive numerical approximations.
Series 51
2.9.1 Geomecric series
This is a series of increasing powers of a variable in the form
s = 1 + x + x 2 + x 3 + .... (2.35)
Successive terms will become smaller and smaller, so that the series
converges, if -1 < x < 1. The sum s of an infinite number of terms
may be obtained by the device of multiplying each term in the series by
x and subtracting the result from the original series, thus
sx = x + x 2 + x 3 + X4 + .... (2.36)
Subtracting (2.36) from (2.35) we obtain
s(1-x) = 1, s = 1/(1-x).
Example 2.18
The theory of the heat capacity of a crystal gives rise to the series
1 +e- hv / kT +e- 2hv/ kT +e- 3hv / kT + ....
This can be seen to be a geometric series by making the substitution
x = e- hv / kT,
so that the sum of the series is
s = 1/(1-e- hv/ kT ).
f '( ) = df(x)
x dh'
Next
f"(a+h)=2a2+2x3a3h+ ... ,
so that at h = 0
f"(a) = 2a2'
Hence Taylor's theorem can be written in the form
y= f(x)
f(o+h)
I
I
I hf'(o)
A
_________________ JI
f(o)
h I
I
I
I
o o+h x
Fig.2.S
Example 2.19
Given that x = 1.5 is an approximate root of the equation
In X = x 2 - 2, use Newton's method to find a solution correct to four
decimal places.
Put
f(x) = lnx-x 2 +2
f'(x) = l/x - 2x.
f(1.5) = 0.1555, 1'(1.5) = - 2.3333
Xl = 1.5 - 0.1555/( - 2.3333) = 1.5666
(2.43)
Example 2.21
Let
f(x) = sin x f (0) = sin 0 = 0,
then
f'(x) = cos x, 1'(0) = cos 0 = 1,
f" (x) = - sin x, f" (0) = 0,
f'''(x) = -cos x, !"'(O) = -1,
so that
x3 x5
sinx = x - 3! + 5! - ... = x for small x. (2.44)
Example 2.22
Let
f(x)= In(1 + x), f(O) = In 1 = 0,
f'(x) = 1/(1 + x), 1'(0) = 1,
f"(x) = -1/(1 + X)2, 1"(0) = -1,
so that
This has been written as the variable y because the definition of the
logarithm gives
y = In(1 +x) if (1 +x) = e Y
so that we may also write
y2 y3
e Y = (1 + x) = 1 + y + 2! + 3! + ...
so that
y2 y3
X = Y + 2! + 3! + .... (2.47)
Now equations (2.46) and (2.47) describe the same relation between x
and y; they differ in that the first gives y as an explicit function of x, and
the second gives x as an explicit function of y. This is an example of the
inversion of a series, which is discussed further in the next section.
Example 2.24
The binomial expansion gives a useful approximation when an
expression contains an algebraic denominator, such as
1 1
(a + X)3 - a 3(1 + x/a)3
1 (
= a3
)-3
1 +~
Example 2.25
Given the relation between the equilibrium constant K and the degree
of dissociation 0: for a solution of a weak acid at concentration c
0:2C
K=-1-'
-0:
~ = -~±J(~)(1+ ~yI2,
so that the expression in the right-hand bracket is nearly unity, which
means that the binomial expansion will converge rapidly. Hence
~= -~ + J(~)(1+~ -~ 1~2+ .. .)
as required.
Example 2.26
Given the series (2.46), obtain the series (2.47).
By comparing (2.50) with (2.46) we obtain a2 = -1/2, a 3 = 1/3,
a4 = -1/4, ... , so that (2.52) becomes
x = y+ y2/2+ y3/6+ y4/24 + y5/120+ ....
which is (2.47).
An alternative method of inverting a series is by successive approxi-
mations. We write equation (2.50) in the form
x = Y -a2x2 -a 3x 3
and substitute successive approximations for x into the right-hand side.
The first approximation is x = y, which we substitute, retaining powers
of y up to y2, so that
Example 2.27
The equation of state of a real gas is a relation between pressure p,
molar volume v and temperature T. The virial equation of state is a
power series in the density l/v
pv = RT[1 + B(l/v) + C(1/v)2 + D(I/v)3 + .. .].
An alternative is to use a power series in pressure in the form
pv = RT+B'p+Cp2+D'p3+ ... ,
and series inversion is used to obtain relations between the coefficients
B', C, D', ... and the virial coefficients B, C, D, ....
60 Differential calculus
Write the series in the form
p = RT(ljv) + B'p(l/v) + C'p2(1/v) + D'p3(1jv) + ...
and the successive approximations are used to substitute for p on the
right-hand side.
First approximation,
p = RT(l/v).
Substitute, retaining powers up to (1/v)2.
Second approximation,
p = RT(l/v) + B'RT(1/v)2.
Third approximation,
p = RT(l/v) + B'(l/v)(RT/v + B' RT/V2) + C'(1/v)(RT/v)2
= RT/v + B'RT/V2 + (B'2 + RTC')RT/V3
Fourth approximation,
p = RT/v+ B'(l/v)[RT/v + B'RT/V2 + (B'2 +RTC')RT/v 3]
+ C'(l/v)(RT/v + B'RT/V 2)2 + D'(1/v)(RT/v)3
= RT/v+ RTB'/v2 + (B,2 + RTC')RT/V3
+ [B'3 + 3B'C'RT+D'(RT)2]RT/v4 •
The required relations between the coefficients are
B = B',
e= B'2 + RTC',
D = B'3+3B'C'RT+D'(RT)2,
etc.
Example 2.28
. sin x
11 m--
x-+O X
Example 2.29
Evaluate
lim (x In x).
Example 2.30
Find the condition for the hyperbola
x2 y2
---=1
a2 b2
to be rectangular (asymptotes perpendicular to each other).
We have
b2 x 2 _ y 2 a 2 = a 2 b2
and the asymptotes are straight lines of slope equal to the value of
dy/dx as x -+ 00 and y -+ 00. By implicit differentiation
2b 2 xdx -2a 2 ydy = 0,
dy b 2 x
dx = a 2 y'
The principles of Newtonian mechanics 63
which gives 00/00 as x and y tend to infinity. Differentiation of
numerator and denominator with respect to x gives
. m
11 dy l'
- = 1m 2
b2
X"" 00 dx X"" 00 a dy/dx
Example 2.31
An alternative way of evaluating limits is by the use of series
expansions. Example 2.10 gives an equation that occurs in the theory of
the heat capacity of a monatomic solid. The high-temperature limit, as
T -+ 00, of the expression
hv e h./kT
kT2 (e h./kT _1)2
Example 2.32
The conventional order of definition of mechanical properties is, first,
velocity as rate of change of distance, then acceleration as rate of change
of velocity and, finally, force as mass x acceleration. This order can be
reversed with advantage in physical applications; we can use a
transducer to measure the force that is exerted on a sufficiently massive
body as a result of acceleration of a containing vehicle. Knowing force
and mass we can calculate the acceleration; knowing acceleration as a
function of time we can calculate velocity, and knowing velocity as a
function of time we can calculate the distance travelled. This is the basis
of inertial navigation systems.
The expression -!mv 2 is called kinetic energy and the integral on the left
is called work. If the force is constant the work done becomes
F(X2 -Xl) so that
work = force x distance = change in kinetic energy. (2.57)
The forces with which we are concerned are often due to interaction
with a field, as with mass in a gravitational field or electric charge in an
electric field. The field strength is defined as the force on unit mass or
charge. In the earth's gravitational field the force exerted on a body of
mass m is called the weight wand the field strength is denoted by g, so
that
w=mg.
By comparison with (2.54) we see that g can be identified as the
acceleration of free fall under gravity.
The corresponding force F exerted by an electric field of strength E
on a charge e is
F= Ee.
When a body is given a displacement dx in the direction opposite to
the force F exerted on it by a field, work F dx is done on the body and its
energy increases. Energy that changes with position in a field is called
potential energy and is denoted by V. Since force is in the direction of
decreasing potential energy we have
dV
dV= -Fdx, F= - dx'
so that
force = - potential gradient. (2.58)
66 Differential calculus
If a body falls freely under gravity, or a charge is moved by an electric
field, the decrease in potential energy is equal to the work done, which
by (2.57) will be the increase in kinetic energy. Denoting kinetic energy
by T and potential energy by V we have
T+V = constant. (2.59)
For this to be true, all of the assumptions that have been made above
must be true. In particular, the only effect of the action offorce must be
to increase the velocity of a body. This will be so only in the absence of
friction. Movement through a resisting medium, such as a fluid, or
sliding on a rough surface introduces a resisting force which reduces the
acceleration of the body; some of the work done is converted into heat.
At the molecular level, linear motion is converted into random thermal
motion of the body and its surroundings, which is an irreversible
process in the thermodynamic sense. Thus, although the conservation
of momentum always applies to real processes, conservation of energy
in the simple mechanical sense does not unless some formulation which
includes heat is adopted.
The motion of a structured body in three-dimensional space, such as
that of a molecule in a fluid, is resolved into linear motion (translation)
of the centre of mass and rotation about the centre of mass. Rotational
properties are expressed in terms of polar coordinates (Section 1.7.5).
Our discussion of mechanics has, so far, neglected structure, as if the
masses concerned were point particles. The arguments used apply to
translation of the centre of mass of an extended body when it is treated
as if all the mass were concentrated at that centre of mass, and all the
forces were moved parallel to themselves so as to pass through that
centre of mass. An applied force not passing through the centre of mass
will tend to produce rotation of the body. We therefore introduce the
principle of moments; the moment of a force about a point is defined as
the product of the force and the perpendicular distance from the point
to the line of action of the force. For an extended body to be in
equilibrium under the action of forces, not only must the sum of the
components of the forces in any direction in space be zero, but also the
sum of the moments of the forces about any point must be zero,
clockwise and anticlockwise moments being given opposite signs.
A molecule may be regarded as a set of discrete masses at the
positions of the nuclei. Fig. 2.6 represents a diatomic molecule
consisting of masses m1 and m2 rotating about an axis perpendicular
to the paper through point O. Mass m1 moves with tangential velocity
V 1 in a circle of radius r 1 about O. In time interval dt the radius vector r1
The principles of Newtonian mechanics 67
Fig. 2.6
where 'i is the perpendicular distance from mi onto the axis. The
moment of inertia about a particular axis is then
1= 'Lmi'f.
The principles of Newtonian mechanics 69
When the molecule is planar we can choose rectangular coordinate axes
OX, Of through the centre of mass as in Fig. 2.7. Mass ml at point
P(x, y)will contribute m1 (PR)2 = m1 y2 to the moment of inertia about
the x-axis OX. Ifwe then choose arbitrary axes OX', Of' by rotating
the axes through an angle (), the point P will assume new coordinates
(x', y') relative to OX', OY' given by
y' = PN = PM -MN = PM -QR = ycos()-xsin(),
x' = ON = OQ + QN = OQ + RM = x cos () + y sin ().
y
yl\
\
\
\ P(X,y)
\
\
\
\
\
\
\
\
\
\
\
\
\ M
\
\ ,,.,,
, ' ....0' \ \ x
\
Fig. 2.7
= -2Lmix;y; = o.
The expression LmixiYi is called the product of inertia about the x- and
70 Differential calculus
y-axes, and we see that if this is zero the moment of inertia will be a
maximum or a minimum. The axes for which this applies are called the
principal axes of the molecule, and the moments of inertia about the
principal axes are called the principal moments of inertia of the
molecule.
In three dimensions we find corresponding conditions, that when the
axes are chosen so that the products of inertia about each pair of axes
are zero,
Ee
E m1+6 r1
0I
r2
-6
I~
~
... Ee
Fig. 2.8
The principles of Newtonian mechanics 71
perpendicular to the plane containing the forces. If the molecule were
inclined at an angle e to the direction of the field, as in Fig. 2.9, the
perpendicular distance between the forces would be r sin e so that, in
general,
torque = Eer sin e = J.lE sin e, (2.68)
£e
+e
£
ma
r----------~-i-n~9------~-e
£e
Fig. 2.9
the product re = J.l being called the dipole moment of the molecule.
Such a pair of equal and opposite forces will have no effect on the
position in space of the centre of mass of the molecule because, when
they are translated so as to pass through that centre, their resultant is
zero. The effect of the field is therefore to apply torque and produce
pure rotation of the molecule.
For linear motion equation (2.56) applies, so that
force = rate of change of linear momentum. (2.69)
In Fig. 2.8 the forces and the momenta are both vectors (Section 3.9)
passing through the mass points and perpendicular to the axis of the
molecule. We define angular momentum as the moment of the
momentum, so that mass m1 has linear momentum m1 V1 and angular
momentum m 1V1 r1 about an axis through O. From this definition and
(2.69) we obtain
moment of force = torque
= rate of change of angular momentum, (2.70)
the total angular momentum of the two masses about 0 being
angular momentum = m 1v 1 r 1 +m2v2r2 = (m 1 d +m2rDw = Iw.
(2.71)
72 Differential calculus
The rotational kinetic energy of our molecule is simply the sum of the
kinetic energies of the masses, so that
kinetic energy = tml vf + tm2v~ = t(m1rf + mzd)w Z = tIw 2. (2.72)
The relations used to describe angular motion are thus the same as
for linear motion if we replace force by moment of force, mass by
moment of inertia and velocity by angular velocity.
The motion of a structured body in three-dimensional space is
resolved into translation of the centre of mass and motion relative to
the centre of mass. For a rigid body we need to consider only
translation and rotation, but for molecules we must include also the
vibration of the chemical bonds. This vibration is discussed in Example
6.9, and makes additive contributions to both potential energy and
kinetic energy.
Problems involving collisions between structureless particles are
solved by using the principle of conservation of linear momentum.
When rotation is possible, we have a corresponding conservation of
angular momentum. When structured particles collide, equal and
opposite forces act on each one at the point of contact. If we calculate
the total angular momentum of the two bodies both about the same
arbitrarily chosen axis, the forces acting on impact contribute equal and
opposite moments about that axis. By equation (2.70) these produce
equal and opposite rates of change of angular momentum, so that the
total angular momentum about any axis remains unchanged.
In problems involving real molecules it may be necessary also to take
account of the energies of allowed quantum states; these impose
additional constraints, but the principles outlined above still apply.
CHAPTER 3
Differential calculus in
three or more dimensions;
partial differentiation
x
z,x plane at constant y
Fig. 3.1
so that
(4x - 3y)dx = 3xdy at constant z
(ox/oY)z = 3x/(4x - 3y).
This example shows that the chain rule of Section 2.6 does not apply
to the partial derivatives when a different variable is held constant in
each case; with partial derivatives a minus sign appears, thus
(oz/ox)y(ox/oY)z = 3x = - (oz/oY)x. This relation is derived analyti-
cally in Section 3.6.
When a function is defined in terms of particular independent
variables, such as z (x, y), it will be clear that partial differentiation with
respect to one variable implies that the other is to be kept constant. We
can then use the abbreviated notation oz/ox to imply constant y
without ambiguity. In physical applications, however, we normally
have a wide choice of independent variables. Thus for some property X,
if we write oX / aTit may mean at constant v or at constant p, these being
different quantities (the relation between them is discussed in Section
3.6). We therefore need either to use the full notation, as (oX/aT)v and
(oX/aT)p, or to begin with an explanatory statement such as 'using
independent variables v,T' so that oX/aT then means (oX/aT)v and not
(oX/aT)p.
The higher-order differential coefficients such as (0 2 z/ ox2 )y are a
straightforward extension of the rules for total differentiation, but a
new possibility also arises in the form of a mixed second derivative. This
may be written in the form 0 2 z/oxoy or more explicitly as
8y
z(x,Y)
O Z(X+h,Y+6Y )
fix
Fig. 3.2
An alternative approach to calculus 77
the paper. By changing x first and then y the change in z is
oz (oz 0 oz )
~z = ox ~x + oy + ox oy ~x ~y,
whereas change first in y and then in x gives
oz (oz 0 oz )
~z = oy ~y + ox + oy ox ~y ~x.
In each case the expression in the brackets is the slope at z (x, y) plus the
increase in slope on moving by either ~x or ~y. The two expressions will
be the same so long as the order of successive partial differentiation can
be reversed. The expressions must be the same if a unique surface exists
over which the alternative movements are made, and the conditions for
that are that the variables shall be independent and the function and its
derivatives shall be finite, continuous and single-valued.
dz = (!:)y dx.
dz = (!:)y (:;)x
dx+ dy. (3.5)
Example 3.1
The pressure of a gas depends on the temperature Tand molar volume
v. Thus p = p (T, v) so that
dp = (Z)v (::)T
dT+ dv.
dz = (!:), (:;)x
dx+ dy, (3.12)
from (3.9)
from (3.10)
(3.17)
Substitution of (3.16) and (3.17) into (3.12) and using (3.13) gives
dz = (!:) w,'
dx+ (:z)
y w,X
dy+ (!:) "X
dw, (3.18)
which would be obtained if we had assumed (3.8) in the first place and
written the total differential to include the superfluous variable.
The number of independent variables that properly apply to a
particular application is a physical problem, not a mathematical one.
The significance of the above is to show that we do not make any
mathematical errors as a result of a physical error in choosing that
number.
82 Three or more dimensions; partial differentiation
3.5 Exact differentials
This term applies to an expression of the form X dx + Y dy, and means
that a quantity, z say, exists whose total differential is of that form, so
that
dz = Xdx+Ydy. (3.19)
This will be so if the relation (3.19) is identical with
and the test for exactness is to see if this is so. This can be done by taking
the mixed second derivative and checking that the order of differenti-
ation can be reversed. Thus to confirm that
and Y= (oz)
oy x
we calculate
Example 3.2
Show that (2xydx + x 2 dy) is an exact differential, whereas
(x 2 dx + 2xy dy) is not.
Applying the cross-differentiation test, equation (3.21),
o
oy (2xy) = 2x
o
ox (x 2 ) = 2x
and the test succeeds with the first expression whereas from the second
expression
o
_(x 2 ) =0
oy
o
ox (2xy) = 2y
Example 3.4
Cross-differentiation of an exact differential is used to obtain the
Maxwell relations in thermodynamics from the Gibbs equations.
Given the exact differential
dG = vdp-SdT,
84 Three or more dimensions; partial differentiation
cross-differentiation gives
(&)
dz = -
au V,w,...
du + -
av
(&) U,w,...
(&)
dv + -
aw U,v, ...
dw +
( ~~ ) u, w, ... = (~:) v, w, .. .
At constant z, dz = 0, so that
Dividing by dy at constant z
(3.24)
which is the chain rule for partial derivatives mentioned in Section 3.1.
Example 3.5
Equation (3.24) can be used to obtain purely mathematical relations
between experimental properties that are defined as partial derivatives.
Relations between partial derivatives 85
The coefficient of expansion IX = (l/v) (av/OT)p, isothermal compress-
ibility ,,= - (l/v) (av/ap)T and thermal pressure coefficient y =
(op/OT)vare related by
ICy = _!(av) (a p ) = !(av) =
V ap T or v v or p
IX,
This equation relates partial derivatives when the variable being held
constant is changed.
Example 3.6
The molar heat capacity at constant pressure Cp = (aH/OT)p, and that
at constant volume Cv = (aU /OT)v are related because H = U + pv, so
that
86 Three or more dimensions; partial differentiation
To develop this into a relation between C p and C v we need to relate
(oH/O/)vand (oH/O/)p. From equation (3.26)
so that
(~~)T =T(:!)T+ V.
Using the Maxwell relation of Example 3.4, we obtain
( OH) = v _T(OV) ,
op T or P
so that
Example 3.7
For a mixture of substances 1 and 2 which is able to gain or lose
numbers of moles dn l and dn2 of substance, the total energy U is given
88 Three or more dimensions; partial differentiation
in differential form by
dU = TdS - pdv + J.ll dn l + J.l2dn2.
The variables S, v, n l and n2 are extensive. We write
U = U(S, v, nl' n2)
and Euler's theorem gives
au au au au
S as +vav+nlanl +n2an2 = U.
But from the given equation
au au au
as =T, av= -p, and -;--- - .-2,
II
un2
so that
U =TS-pv+nlJ.ll +n2J.l2.
On the other hand, for the Gibbs function G we have
dG = v dp - S dT + J.ll dnl + J.l2 dn2
where nl and n2 are extensive but p and Tare intensive. Euler's theorem
therefore gives
3.9 Vectors
A vector quantity is one that has magnitude, direction and sense, in
contrast to scalar quantities, which have magnitude only. Mass and
time are examples of scalar quantities, whilst displacement, velocity and
force are vectors. A vector quantity may be represented by an arrow of
length proportional to the magnitude of the quantity and in the
appropriate direction and sense; they are then added together by
joining the arrows head to tail as in Fig. 3.3, where displacements OP
and PQ are added to give OQ.
o~------~------~
A
Fig. 3.3
x
Fig. 3.4
Example 3.8
When a gyroscope spins with angular velocity c.o about a horizontal axis
it has angular momentum 1c.o, represented by a vector located in the
axis of rotation as shown in Fig. 3.5. If it is supported at one end only,
mg
I 'I
,, '
,
~ __~x____+T~____~~/~
mg
_ - - \ - -1. mgxdf
~
Fig. 3.5
92 Three or more dimensions; partial differentiation
the weight mg of the gyroscope is balanced by upthrust mg at the
support; these forces generate a couple mgx represented by a vector in
the horizontal plane perpendicular to the axis of rotation. From
equation (2.70).
couple = rate of change of angular momentum,
so that in time interval dt angular momentum mgxdt is generated,
represented by a vector into the plane of the paper in Fig. 3.5 in the
direction perpendicular to the axis of rotation. The two angular
momentum vectors are drawn in the figure as when viewed vertically
downwards; the resultant, and so the axis of rotation of the gyroscope,
rotates in the horizontal plane in an anticlockwise direction from that
point of view. This rotation of the gyroscope about the vertical axis is
called precession, the direction of precession being the opposite of what
would occur if the gyroscope were a wheel running on a horizontal
table.
This also applies to the effect of a magnetic field on a spinning
magnetic dipole; the field imposes a couple on the spinning dipole,
which therefore precesses about the direction of the magnetic field.
Nuclei that behave in this way are detected by the absorption of
radiation of frequency equal to the precession frequency in a nuclear
magnetic resonance spectrometer, the precession frequency depending
on the magnitude of the magnetic field strength at the particular
nucleus.
Fig. 3.6
This illustrates another useful principle, that plane surface area may be
represented by a vector of length proportional to the area and in the
direction normal to the surface. When plane surface area A is viewed
obliquely from the direction making an angle 0 to the normal to the
surface, the area seen is A cos O. If we define the line of sight by unit
vector r this may be. expressed in the form of a scalar product:
effective area = A.r = A cos O. (3.37)
The moment ofa vector about a point, such as that of force F about
o in Fig. 3.7, can be expressed as the vector product of F and the radius
94 Three or more dimensions; partial differentiation
o,
\
\ rsin 9
Fig. 3.7
so that
V= -p.E (3.40)
The condition that a scalar product of zero means that the vectors are
Vectors 95
E
Ee
..".
x:
+I----*+e
-e
Ee
Fig. 3.8
f: f(x)g(x)dx =0 (3.42)
and
oFz of,, (3.47)
oy = oz·
We may therefore express departure from the condition of the work
being independent of path as departure from (3.47). This is called the
curl of the vector field, defined by
_ I_
curI F - (OFz OF,,) _(OF" OFz) k(OFy
- - - +J - - - + - -OF,,)
oy oz oz ox ox -
oy .
(3.48)
The first bracket in (3.48) is the x component of the vector curl F, and it
depends on the changes in F in the (y, z) plane, and in fact measures
rotation of the field in that plane. The significance of curl F is that it
measures eddies or swirls in the field: if we move from one point to
another along a path that is opposed to the eddies we do more work
than if we move along a path that is in the same direction as the eddies.
Equation (3.48) contains the derivatives of the components of F in
perpendicular directions, such as ofz/ ox. The derivatives in the parallel
directions, such as of,,/ox, give the components of the increase in
magnitude of the vector. This is a scalar quantity called divergence,
defined by
d· F _ of,, of,, of.
IV - ox + oy + OZ ' (3.49)
and detects sources or sinks in the field; this would apply if heat were
being supplied at a point in a flowing fluid because thermal expansion
would then create increase in velocity and so a positive divergence of
the velocity field.
These concepts provide concise notation in problems involving
three-dimensional fields. The nabla operator (3.44) can be used for
vector fields by noting that it can be regarded as a vector, so that both
scalar and vector products of V with F can be written. It can be seen
from the above definitions that
V.F = divF and V x F = curiF. (3.50)
CHAPTER 4
Integration
r
x = a to x = b, we have a definite integral and the integration constant
I
cancels out.
a x b
Fig. 4.1
i 1/
follow the curve, and the sum of the steps between a and b becomes
dY dy
L
x=b b (b)
lim d bx = -d dx = dy = f(b) - f(a). (4.7)
6s-0 x=a X a X I (a)
100 Integration
y
y = f(x)
feb) ------------------
f(o)
a b x
Fig. 4.2
Thus
or
f(b) = f(a) +
i dx
abdY dx (4.8)
Example 4.1
If a quantity X depends on temperature T and pressure p, we write
X = X (T, p). The change in X from temperature Tl to temperature T2 at
constant pressure is
f X"
X,,+l
dx = n + 1 + C, n + -1,
fdX = x+C,
f eX dx = eX + C,
feaxdx = ~eax+c,
Example 4.2
Example 4.3
Integration by parts can sometimes yield the same integral again, which
can then be evaluated by the solution of an algebraic equation.
Denoting the integral by the symbol I, we could have
1= f sin 2 0 dO = - f cos 0 dO
sin 0 cos 0 + 2
1= ~ - ~ sin 0 cos 0 + c
2 2
i
so that, for example
"/2 7t
sin 2 0 dO = -.
o 4
r I -r
taking the value of the integral at the upper limit minus that at the lower
limit. Thus
u dv = [ uv v duo
Example 4.4
The integral of In x is obtained as the product of In x and unity:
= x In x - f ~ x dx
=xlnx- fdX
= x In x -x +c.
104 Integration
4.2.4 Integration by substitution
This applies when the integral can be regarded as the product or ratio of
two parts, one part being the differential of the other part.
Example 4.5
I = fX~!X4
Since d (x 2 + 4) = 2x dx, the factor of 2 being immaterial, we substitute
U = x2+4, du = 2x dx.
Then
I ="21 fdU
-;- = 1ln U + C
= 11n(x2 +4)+C.
Example 4.6
1= fnxx dx.
f
When
f(X)
1= g(x) dx,
I f f
= x:: 2 dx = (X - X22: 2 ) dx
f f
= x dx - X22: 2 dx
x2
= 2- ln (x 2 +2)+C.
Example 4.8
I=f x 8dx
-4x
3
Put
8 8 ABC
--:=---=
x 3 - 4x x (x + 2) (x - 2)
=-+--+--
x x+2 x- 2
Then
A (x + 2) (x - 2) + Bx (x - 2) + Cx (x + 2) = 8.
Equating coefficients of x 2, of x and constant terms gives
A+B+C = 0, B=C and A = -2.
Hence
B = C = 1,
and so
1= -2 dX f --+
f -+ dx f -dx-
x x+2 x-2
X2
=In ( ~
-4) +c.
4.3 Standard forms of integral and numerical methods.
Although it is generally possible to differentiate a continuous
function, integration only gives a closed expression in particular cases.
Extensive tables are available giving the integral of many standard
functions; a particularly comprehensive set is included in the
Handbook of Chemistry and Physics, published at intervals by the
Chemical Rubber Publishing Company, USA. One important stand-
1
ard form is
00
.J 1t
e -",'x'd x·=--
o 21X
and this is discussed in Section 5.2.
For integrals which cannot be evaluated in closed analytical form,
geometrical and numerical methods are available at least to obtain an
Multiple integration 107
approximate solution. An integral can be interpreted as the area under
a curve, so that we can draw the curve and measure the area between
particular limits to evaluate the definite integral. This applies also when
we want to integrate a quantity which appears as a curve on a recorder
trace. The area can be found by counting squares or vertical strips, or by
cutting out the shape and weighing it. Numerical methods are based on
the same principle, and details are given in texts on numerical analysis.
The Euler-Maclaurin formula is particularly useful and this is
discussed in Section 4.6.
both of which mean integrate with respect to x and then integrate the
result again with respect to x. This is a double integral and the solution
will contain two integration constants.
Multiple integration is most useful when we have more than one
independent variable. If we write
ff(X,Y)dX
rf rf
variable whilst regarding other variables as constants. We do this when
defining the multiple integral
Example 4.9
=~ [~J: = 332 .
Notice that the limits of integration must be applied successively, not
simultaneously.
= [i:I 15'
Since the variable y occurs in the limits of the integration with respect
to x it is essential that the inner integral be evaluated first. If the integral
were written in reverse order we would have
y2 Jl
JoI' Jo11 x 2
ydydx = JoI' [X2 I' x 2
-2- 0dx = Jo 2 dx
f
will cancel.
d
dx f(x) dx = fdf(X) f
~ dx = df(x) = f(x).
oy
a f f(x,y) dx = fof(X,y)
oy dx. (4.10)
This will have meaning when the integral is between limits, and
complications arise if the limits of integration contain the variable
of differentiation in which case Leibnitz's theorem is used, which
is
a ib(Y) ib(Y) of(x, y) ob oa
a
y a(y)
f(x,y)dx=
a(y)
aY dx+f(b'Y)a-f(a'Y)a·
Y Y
(4.11)
Example 4.11
Leibnitz's theorem can be used to obtain an alternative form for an
integral by differentiation with respect to a parameter other than the
variable of integration. Thus if we differentiate the integral
1
00
o
2 1
xe- ax dx = -
2a
with respect to a we obtain
- 1 00
0 x 3 e -ax dx
2
=
d 1 1
da 2a = - 2a 2 •
f~' f(x)dx
so that the integral may then be evaluated. We first expandf(x) as a
110 Integration
Taylor series about the value f(a) at x = a in the form
(x a)2
f(x) = f(a) + (x - a) I' (a) + 2! I" (a) + ... , (4.12)
Ifwe now choose a to be the mid point of the interval (xo, Xl) the terms
containing derivatives of odd order will cancel; we put h = (Xl - x o), so
that (Xl - a) = h/2 and (xo - a) = - h/2, giving
We now use Taylor series again to expressf(a),f" (a),f 4 (a), ... in terms
of the values at the end points of the interval; we adopt the device of
taking the sum of the values of the function at the end points, but the
differences of the values of the derivatives. Again using (4.12), we have
and this is used to eliminate f(a) from (4.13). We also write Taylor
expansions off' (X),!3 (x), .. .in the form
4 3 3) 2
f (rx)=};1 ( f (xd-f (Xo) -2h2 3!f 6 (rx)- ... , (4.16)
and so on. Substitution of (4.14), (4.15) and (4.16) into (4.13) gives
f~If(X)dX = ~h (f(Xd+f(Xo»)
1 (2!1- 1
- 22 3! )h2 (f ' (xd -I'(xo) )
+214 (1
3! (1 1) - (1
2! -3! 1)) h
4! -5! 4(f 3(xd-f 3)
(xo)
rxof(x) dx = rXI
+ r
X2
+ ... + rxo ,
Jxo Jxo JX 1 JX__ 1
and the right hand side of (4.17) then simplifies. The first term can be
expressed in terms of the sum of the values of f(x) at the n points; in
particular, if h = 1 this gives
(4.18)
112 Integration
and the numerical coefficients of subsequent terms can be shown to be
1/30240, 1/1209600, ....
The two most common applications of this formula are to the
numerical evaluation of convergent, but intractable, integrals which do
not possess a solution in a closed form (numerical quadrature), and to
the replacement of the sum of a convergent but intractable series by an
integral (e.g. in statistical mechanics).
The formula as written in (4.21) contains the derivatives at both ends
of the range of integration; if this range is from zero to infinity and if the
values of the derivatives approach zero as x tends to infinity, the
formula simplifies to
roo 1 1 1
Jo
00
Example 4.12
Use the Euler-Maclaurin theorem to evaluate r:
e-«' x' dx when
a = 0.5, and compare the result with example 5.1.
We putf(x) = e-«2 x2, from which we obtainf(O) = l,J(oo) = O,and
all of the derivatives of odd order are zero at both limits. Hence by
1
(4.22),
00
~ e _«'n' --.
e _«'X2 d x = £..., 1
o "=0 2
The integral is shown in example 5.1 to have the value .j 'It/2a. In order
to obtain rapid convergence we must have a small enough value of IX.
Putting IX = 0.5 we obtain
L e-«2n2= 1+e-O.S+e-l+e-2.2S+e-4+
00
"=0
The sum of this series to eight terms is 2.272454, so that we have
agreement with the Euler-Maclaurin formula.
This expression arises in calculating the contribution from trans-
lation in a particular direction to the partition function in statistical
The Euler-Maclaurin theorem 113
mechanics. We then require to evaluate the sum
n=1 n=O
1
The Euler-Maclaurin formula gives
_ (2nmkT)1/2 ~
- h2 a+ 2'
so that
Since the value of the integral is usually a very large number, the
correction of - 1/2 given by the Euler-Maclaurin formula may usually
be neglected and we say that the sum may be replaced by the integral for
the purposes of evaluation.
Example 4.13
Following the previous example, the relation which corresponds to the
rotational contribution to the partition function is
L
00
(21 + 1)e -J(J+l)8/~
J=O
(where () = h 2 /8n 2 Ik). This is also an intractable series, whereas the
corresponding integral can be evaluated in closed form.
When 1(1) = (21 + 1)e- J (J+l)8/T the terms up to «()/T)2 are
1(0) = 1, 1'(0) = 2 -()/T,
f3(0) = -12()/T+12«()/T)2- ... ,
f5(0) = 120 «()/T)2 - ... ,
and all relevant terms become zero as 1 -+ 00. The integral may
be evaluated by making the substitution u = 1(J + 1W/T to
give
1 00
o (21 + l)e- J (J+l)8/TdJ =
T
0'
114 Integration
The Euler-Maclaurin formula (4.22) then gives, to terms in (8jT)2,
f: (2J + l)e-
J=O
J (J+1)6/T =!.- +! - ~(2 -~)
8 2 12 T
- 30~40 ( 120 (~ y)
+ ...
T(
=fi
18 1(8)2
1+ 3Y +15
4(())3 )
Y +315 Y + ....
In this case, the sum may be replaced by the integral when T;}> (). This
applies except for small molecules at low temperatures, for which the
series expansion can be used.
CHAPTER 5
Applications of integration
ff(X)dX
gives the area between the curve y = f(x), the x-axis and the ordinates
x = a and x = b.
If, instead, we required the area between the curve and the y-axis, we
sum horizontal strips of length x and thickness dy to obtain
if
f(a)
(b)
xdy,
if
Fig. 4.2. This can be proved analytically as follows:
i b
f(x)dx+
(b)
xdy
f.b
a
ib Xdxdy dx
f(a)
= a f(x)dx+ a
= f [j(x) + x!'(x)] dx
116 _Applications of integration
= r:: d(xf(x»
X-b
[
= xf(x) ] x=a
= bf(b)-af(a).
The various manipulations used here should be carefully noted.
These relations give the area between a curve and one of the axes.
The area between two curves can be found as the difference between
the area under one curve and the area under the other.
The conditions necessary for this interpretation of integration to be
valid are that the function should be continuous and single-vaiued; this
means that for any given value of x, the function y = f(x) shall have a
single, unique, value. This condition will not be met if the equation
contains quadratic or higher-degree terms in y. Thus the equation
x 2 + y2 = r2
Fig. 5.1
If we simply integrated
the result would be zero, as the sum of equal positive and negative areas;
this is because area has the signs shown in the quadrants of Fig. 5.1.
Plane elements of area 117
When a function is not single-valued, we must restrict the range of
integration, the area in the first positive quadrant of our circle being
obtained by choosing the positive root:
= _r2 r
J"/2
sin 2 0dO
= 1[r 2 /4
by using the result of Example 4.3. The area of a circle drawn entirely
within the positive quadrant is then 1[r2.
is justified as the sum of areas of vertical strips of height Y and width dx.
In the same section, integration is interpreted as giving an increment in
Y by moving along the curve y = f(x), so that
y = s: dy. (5.2)
rf:
If we now combine equations (5.1) and (5.2) we obtain the area as a
double integral
In the same way, we can obtain the area of a circle, using plane polar
coordinates, by summing the areas of annular strips of radius rand
thickness dr as shown in Fig. 5.2. Given that the circumference of a
circle is 2nr the area of an annular strip becomes, as dr --+ 0,
dA = 2nrdr.
dA=271"rdr
Fig. 5.2
fdA = f: 2nrdr = na 2•
= 2n f: r dr = na 2 •
Plane elements of area 119
dA=rdOdr
Fig. 5.3
Example 5.1
An integral that arises frequently in physical applications has the form
1=
1
000 e -IX
2
x
2
dx. (5.4)
1 =
1
000 e -IX
2
Y
2
dy.
We now multiply the two integrals (5.4) and (5.5) together. Since they
(5.5)
This double integral is over the positive quadrant of the (x, y) plane. We
now transform to polar coordinates, using x 2 + y2 = r2; the element of
area dx dy becoming r dO dr. The limits for the same positive quadrant
120 Applications of integration
become 0 to n/2 in () and 0 to 00 in r, so that
This has the effect of introducing a factor r into the integral, which can
now be evaluated by the substitution
Hence
f
so that
[=
OO
e -'"
22
x dx = --.
.In (5.6)
o 20:
Example 5.2
The previous example can be used in the integration of similar forms,
such as
where
[=1
- fOO e-'" 2x 2dx
2 20: 2 0 .
· x
IIm---=== I·1m 1 2 2 =0'
X _00 e'" x x _ 00 20: 2 xe'" x
so that
P(r,O,z)
z
y
Example 5.4
The volume V of a sphere of radius a centred at the origin must be
found from that of the positive hemisphere, so that the limits of
integration are 0 to n/2 for e and 0 to 2n for tjJ:
= f: 2nr 2 dr = j nr3.
V = 4nr 3 /3.
The length of curve between points (Xl' yd and (X2' Y2) is then
s = i2 ds
where
ds = ..) (dx 2 + dy2)
= ..)[1 + (dy /dx)2] dx,
124 Applications of integration
1
so that
s=
2
ds=
iX, J[1+(dy/dx)2]dx.
1 x,
Example 5.5
The circumference of a circle of radius r centred at the origin is found
from that of the positive quadrant. The equation of the circle is
x 2 + y2 = r2
2xdx+2ydy = 0
dy x x
dx y
The length of arc is therefore
X2
Jo
S=f"'(1+ 2
r -x2 )dX=f"'J(:
Jo r-x 2)dx.
t
Putx=rcos(),dx= -rsin()d(),
P(r,lJ,¢»
Fig. 5.6
f:
We have
dA = 1,
f: dr = u,
(,,/2
Jo sin 2 ¢ d¢ = n/4,
("/2
Jo sin 3 () d() = 2/3,
so that
p = --m
1N
3 V
10
00
u2 v(u)du.
Fig. 5.7
(~y dVLdVRg(r).
The component of the attractive force - dw(r)/dr perpendicular to
the plane is - (dw(r)/dr) sin () sin ¢. The differential contribution to the
pressure from the intermolecular forces, dPinl' is then
. 3 () . 2
= - ( N)2
V
3 dw(r)
sm sm ¢r g(r)~drd()d¢d(XdA.
We again have a five-fold integral and the limits are 0 to 1 in A for unit
area, 0 to 1 in (X to include all pairs for a given direction of the radius
vector r, 0 to n/2 in () and ¢, with a compensating factor of 4 to cover a
complete hemisphere on each side of the plane, and 0 to co in r:
Pint =
-4 ( V
N)2 l<Xl
0
dw(r)
r3g(r)~dr Jo("/2 sin 3 ()d() Jo("/2 sin2¢d¢ fl
0 d(X JoedA,
=
N)2
-1n ( V Jo(<Xl r3g(r)~dr,
dw(r)
i
differentiation and integration so that
01 = b
of (x, y, y') dx = O. (5.17)
08 a 08
Now change in 8 at constant x, 4>(x) and '1(x) produces changes in y and
The calculus of variations 131
y
'. ,,
,
-"1= cp(x}+ £l'}(x)
...
y =cp(x) -'"
"', ...
,.
x
Fig. 5.8
in y' so that
oj ojdy oj dy'
-=--+--
oe oy de oy' de
and from (5.16)
dy dy' d dy d dy d'1 ,
de = '1(x), - = - - = - - = - = '1
de dedx dxde dx '
so that (5.17) becomes
[OJ Jb J.b d oj
J.•b'1 ,oj
a'dx =
y
J.b 81'1
aY
oj ,
dx = 81'1 -
Ya
da''1 dx .
aXy
(5.19)
(5.20)
Now equation (5.20) contains '1(x) which may be any function, so that
the only way in which the integral must be zero is when the bracketed
132 Applications of integration
part is zero. Hence
af _~ af = 0 (5.21)
ay dxay' ,
Example 5.6
Show that a straight line is the shortest distance between two points in a
plane.
Whenf(x, y, y') is defined by (5.14) we have
f (x, y, y') = y' (1 + y'2)
and since we are regarding y and y' as separate variables
af = o.
ay ,
1= it2
t,
Ldt = it2 (T- V)dt
t,
(5.22)
f L dt = 2 f f F dx dt = 2 ff F dt dx = 2 f p dx,
J
and p dx is action. An early formulation of quantum theory was in this
form, the Wilson and Sommerfeld rule being that p dx = nh over aJ
closed cycle of the motion. It was shown in ExaPlple 1.2 that such a
relation is dimensionally correct.
We can use the calculus of variations from the previous section to
show that the two formulations of the principles of mechanics are the
same. For motion of a particle of mass m in the x direction, the kinetic
energy T = ! mi 2 • When force F acts on the particle in the same x
direction the potential energy will change, so that we write that
potential energy as V(x). The Lagrangian function L then depends on x
134 Applications of integration
(through V) and on x (through n and we have
L(t, x, x) = !mx 2 - V(x). (5.23)
Euler's equation (5.21) then becomes
aL _~ aL = 0 (5.24)
ax dt ax '
and from (5.23) we have
aL
-=
av.
--=F
ax ax '
which is the force acting on the particle, and
aL . d aL .
ax = mx, so that --=mx
dt ax '
and we obtain F = mx, which is Newton's law. Thus Hamilton's
principle (5.22) is the same as Newton's law.
The rotation of a diatomic molecule was considered on the basis of
Newton's laws in Section 2.11, where it was assumed that masses m1
and m 2 were rigidly held at a separation r from each other. By using the
more powerful technique of equation (5.24) we can consider not only
that case but the more general one, where the masses are attracted, or
repelled, by each other with a force that is a function of the separation r.
The potential energy may be written in the general form V(r); this will
apply either to a chemically bonded molecule or to the general case of
the interaction between two particles in three-dimensional space.
We use the polar coordinates shown in Fig. 5.9 to define the positions
of masses m1 and m2 at distances r 1 and r2 from the centre of mass 0 of
the pair of particles. We suppose the two particles to be moving in
space, and so relative to each other, so that both distance r and angle ()
change with time. We define the reduced mass m* of the two particles
when they are at separation r as in (2.66) by
and the rotational kinetic energy from (2.64) and (2.67) becomes in
terms of the angular velocity {j = OJ
rotational KE = !m*r 2 82 • (5.25)
Since the distance r between the particles is now a variable, we also
have radial kinetic energy given by
(5.26)
Generalized dynamics 135
Fig. 5.9
aT = ~(-21mi2) = mi = p.
aqi di "
and if qi were an angular coordinate 0 we would have T = tIlP, and
again
aT d 1 '2
de '
aqi = (2 10 ) = 10 = Pi'
The condition in (5.31) that V is independent of qi means that we
have a conservative system (Section 3.9.2), or that the work done in any
displacement is independent of the velocity (no 'friction').
In these generalized terms we can define also the Hamiltonian
function H which is often used in quantum mechanics. The formal
definition is
(5.32)
Again for the simple case of motion in the x direction, this gives
pi=mi,
Generalized dynamics 137
and
so that
H = mx 2 -(!mx2 - V)
= !mx 2 + V
= T + V = total energy.
This means that the Hamiltonian function has the simple meaning of
being the total energy so long as the system is conservative.
We can express Lagrange's equation in terms of coordinate qi as
dH = qidpi - ~L dqi'
uqi
Hence
oH .
~=qi' (5.33)
uPi
and
oH oL .
- = - - = -po (5.34)
Oqi Oqi ,.
Equations (5.32) and (5.33) are simpler than the equation in terms of the
Lagrangian function L in that they have a degree of symmetry; in most
problems, however, the Lagrangian function is found first and then the
Hamiltonian. The expression (5.33) gives the rate of change of
generalized momentum and so may be regarded as a generalized force.
CHAPTER 6
Differential equations
fdY = f 2XdX
Y = X 2 +c.
The integration constant C is important in differential equations.
A simple physical example is when the rate of change of a quantity x
with respect to time t is proportional to the quantity itself, so that
dx
-= -kx
dt '
giving
fd: = - f kdt ,
In x = -kt+C.
The value of the integration constant can be determined by the use of
boundary conditions, such as knowing the initial value Xo of x at t = o.
Then by putting x = Xo and t = 0,
In Xo = C,
(:J
and
In = -kt,
An alternative interpretation of this form of relation is in terms of a
time constant; if an equilibrium system is disturbed by a small amount
Ax it will return to equilibrium at a rate which to a first approximation
may be assumed to be proportional to the displacement. Then
giving
Ax = Axoe- t /<,
so that after time T the displacement will have decreased to lie = 0.37
of its initial value. This is called the time constant, or relaxation time, of
the system and is seen to be the reciprocal of the first-order rate
constant k in reaction kinetics.
Not all differential equations possess an analytical solution, but
standard methods are available for particular classes of equation. This
classification is based upon the concepts of order, degree, linearity and
homogeneity.
140 Differential equations
The order of a differential equation is that of the highest derivative
occurring in it, dy/ dx being first order, d 2 Y/ dx 2 being second order, and
so on. The degree is the power of this derivative. A linear equation is
one containing only the first powers of a variable and its derivatives,
with no cross-terms, so that
d 2y dy
ao(x) dx 2 +adx) dx +a2(x)y =f(x) (6.1)
Example 6.1
dy
dx +4x = 1.
fdY = f(1-4X)dX
Y = X_2X2+C
Y = x(1-2x)+C.
The geometrical interpretation of this solution is a family of curves,
each of which corresponds to a particular value of the integration
constant C. Suppose we require the solution passing through the point
(1,2), then we substitute x = 1 and y = 2 into the equation so as to
eliminate the integration constant, giving
2 = -I +C, C=3
so that the solution to the differential equation becomes
y = x(I-2x)+3.
Example 6.2
An equation may sometimes be separated into integrable terms even
when the variables are not completely separated. Thus if
dy 2x-y
dx x+1
the above procedure gives
(x + I)dy = (2x - y)dx,
xdy + dy = 2xdx - ydx.
The variables are not separable as such due to the terms xdy and ydx,
but the combination xdy + ydx is actually the differential of the
product xy, so that
dy + d(xy) = 2xdx
which may be integrated term by term to give
f dY + fd(XY) = f 2XdX ,
y+xy = x 2 +C,
so that
x 2 +C
y=--.
I+x
142 Differential equations
Example 6.3
A chemical rea{;tion is described as 'first order' when the rate of
production of product is proportional to the first power of the reactant
concentration. This is a different use of the term 'order' from the
mathematical definition. For the reaction
A ~ B
(a-x) x
In(_a) = kt,
a-x
x = a(l _e- kt ).
Example 6.4
Multiply by dx to give
xdy+2ydx = O.
We now substitute y = ZX, so that
dy = zdx + xdz,
Equations of first order, first degree 143
and
x(zdx + xdz) + 2zxdx = 0,
zdx + xdz + 2zdx = 0,
3zdx + xdz = 0.
The equation will now be of the separable variables kind, to be solved as
before. Thus we collect all terms in dx on one side of the equation and
terms in dz on the other, to give
3zdx = -xdz,
dx dz
x
In x = -! Inz+C,
In x 3 z = C',
x 3 z = C",
where C, C' and C" are constants, which are actually related to each
other by C' = 3C = In C", but the only point of interest is that a
constant has been retained in the correct place. We finally return to the
original variables by the substitution z = y/x, to give
x 2 y = C.
The correctness of the solution may be checked by differentiation and
substitution into the original equation.
Example 6.5
An equation that is homogeneous in the terms containing the variables
but also contains an additive constant can be reduced to a simple
homogeneous equation by using a substitution to eliminate the
constant. Thus if
dy
x-+y=3
ax
we substitute z = = dy, to give
y - 3, dz
xdy+ (y -3)dx = 0,
xdz+zdx = 0,
dx dz
x z
Inx = -lnz+C
Inzx = C,
zx = C',
144 Differential equations
so that
x(y -3) = C.
If a differential equation either has this form or can be converted into it,
direct integration is then possible.
Example 6.6
dz = 2(x + y)dx + (2x + l)dy.
We first test whether or not this is exact, as in Section 3.5, by seeing if
cross-differentiation applies, which means looking for equality of
a a
ay (2 (x + y» and ax (2x+ 1).
z = f2(X + y)dx
= x 2 + 2xy + C(y),
Equations offirst order, first degree 145
this integration being the inverse of partial differentiation with respect
to x so that C(y) is a constant only so far as x is concerned; it may be a
function of y. Then
(aoz) .= a0 (x +2xy+C(y»
y x Y
2
dC(y)
= 2x+-- = 2x+ 1
dy
so that
dC(y) = 1, C(y) =y+C
dy
where C is now a numerical constant. The required solution is therefore
z = x 2 + 2xy + y + c.
This has the effect of converting the left-hand side of the equation into
an exact differential by making cross-differentiation apply (Section 3.5).
We show this by multiplying the left-hand side of (6.3) by a functionf(x)
to give
f(x)dy + P(x) yf(x)dx
and cross-differentiation gives
lnf(x) = f P(x)dx,
f(x) = e fP(x)cIx,
146 Differential equations
omitting any integration constant so as to obtain the simplest solution.
Example 6.7
dy
dx +2xy = 4x.
Example 6.8
Successive kinetically 'first-order' chemical reactions give linear dif-
ferential equations. An extension of Example 6.3 is
A~ B ~C
(a-x) (x-y)
where the rate coefficients kl and k2 are shown, together with the
concentrations of the various species at time t. The rate of production
of C is given by
y = a-
k2 a e -k,t+ Ce -k2t.
k2 -kl
Linear differential equations 147
The integration constant C may be found from the initial conditions
that Y = 0 at t = 0, giving
Y = a[1-(k 2 e -k,t-k 1 e -k,t)/(k 2 -kdJ.
Fig. 6.1 shows typical curves of (a - x), (x - y) and y, each plotted
against time t.
A-B-C
Fig. 6.1
Example 6.9
Put
dy m,x
dx=me,
then
em" (m 2 + m - 2) = O.
This equation will be satisfied by solving the auxiliary equation
m2+m-2 = 0,
Linear differential equations 149
which factorizes to give either m = 1 or m = - 2, so that the differential
equation is satisfied by either y = eX or y = e - zx. It is also satisfied
by y = C1e x and by y = Cze-zX, where C 1 and C z are arbitrary
constants. Furthermore, since the equation is linear, we may also
write
y = Clex+CZe-zx.
This last solution contains the two arbitrary constants expected in the
solution of a second-order equation and is the most general form, the
earlier solutions corresponding to values of zero or unity for the
arbitrary constants.
This method gives solutions whenever the auxiliary equation has real
roots. If the auxiliary equation were to give equal roots, however, we
would obtain only a solution containing one arbitrary constant. In that
case, multiplication of that root by Czx will be found to produce the
required second solution.
A solution to the differential equation may exist even when the roots
of the auxiliary equation are imaginary since these appear as imaginary
exponents in the solution. From Section 1.10
e iy = cos Y + i sin y
and
e -iy = cos Y - i sin y.
Now, the arbitrary constants used in the solution of the differential
equation may also be complex numbers and these will then produce
real solutions, as in the following example.
Example 6.10
Example 6.11
An equation of the form of Example 6.10 is obtained for simple
harmonic motion, when a particle of mass m experiences a restoring
force proportional to its displacement x. Then
d 2x
force = m- = -kx,
2 dt
d2 x k
d t 2 +-x = O.
m
Put
then
a = ±i.j(klm)
giving
x = Ae it.J (kim) + Be - it.J (kim),
V= - f Fdx = !kx 2 ,
r1 r2
14-----
I
- -------*------1 I
• ~ I • • m2
1 0
I 1
~-------------------~
ro
Fig. 6.3
and
so that
d2r
F= -m*-
dt 2
and the motion will be as described above with the usual substitution of
reduced mass m* for m, zero displacement corresponding to the
equilibrium separation r 0 so that x becomes (r - r 0).
This simple model of molecular vibration will apply only to small
displacements; in vibrationally excited states the restoring force will no
longer be proportional to the displacement and the curve of potential
energy against r will depart from the parabola corresponding to simple
harmonic motion.
Example 6.12
Simple harmonic motion will occur only in the absence of frictional
losses. The motion of real macroscopic systems will be subject to
resistance which, often to a good first approximation, is proportional to
the velocity. This applies both to the vibration of an object immersed in
a fluid such as air or a liquid, and to electrical oscillations in the
154 Differential equations
presence of electrical resistance. The process then becomes irreversible
and potential and kinetic energy are gradually con .. erted into heat, so
that the vibration decays at a rate depending on the resistance. This is
called damping. The restoring force is then given by
dx d2x
force = -kx -{3- = m -
dt dt 2 '
where the frictional, or damping, force is {3 times the velocity and acts in
the same direction as the restoring force when dx/dt is positive and so
the displacement increasing.
This is a homogeneous linear equation with constant coefficients and
the usual substitution x = eat gives the auxiliary equation
2 {3a k
a +-+-=0,
m m
with the solution
a= - 2~ [{3 ± ({32 - 4mk)J.
The behaviour then depends on the relative magnitudes of {32 and 4mk,
which means on the extent of the damping.
When heavily damped, {32 > 4mk and the auxiliary equation has real
roots,say -al and -a2; the solution then has the form of Example 6.9,
This also gives exponential decay, the damping force being just
sufficient to prevent oscillation, but allowing the decay to occur most
quickly (curve B in Fig. 6.4). This is called the critical damping
condition.
If {3 were zero we would have purely imaginary roots to the auxiliary
equation as in Example 6.11, and so simple harmonic motion. When
o < {32 < 4mk, less than the critical amount of damping is present and
the roots of the auxiliary equation will be complex, of the form
a = d±if,
Linear differential equations 155
x
Fig. 6.4
where d (for decay) andf (for frequency) are real numbers given by
d = - p/(2m) and f = [J (4mk - p2)J/2m
= J(~- :~2).
The solution is then the product of terms arising from the real and the
imaginary parts; the real part is exponential decay and the imaginary
part gives periodic oscillation of frequency f/(2n), the result being a
damped oscillation of form C in Fig. 6.4. We notice that the frequency
of the damped oscillation is lower than the value (1/2n)J(k/m) which
would be obtained in the absence of damping.
The critical damping condition is very important in practice because
that gives the fastest response without oscillation. This is achieved by
adjusting the value of p: in a simple moving-coil galvanometer system
by adjusting the circuit resistance to equal that for critical damping,
and in more complicated control systems by adjusting the control
signal input which depends upon the derivative dx/dt.
156 Differential equations
6.3.2 General linear equation with constant coefficients
This has the form
dRy dn - 1 y
ao dxn + a1 dx n- 1 + ... + anY = f(x). (6.6)
Example 6.13
d 2y dy
dx 2 + dx -2y = 1-2x.
Heref(x) = 1-2x, which is a first-degree polynomial, so that we try
the general first-degree polynomial y = ax + b as a solution.
Substituting this into the equation gives
a-2(ax+b) = 1-2x,
which is satisfied by a = 1 and b = O. A particular integral is therefore
y = x. The complementary function for this equation was found in
Example 6.9 to be
and the general solution is the sum of the complementary function and
Linear differential equations 157
a particular integral, or
y = CleX+C2e-2x+x.
Example 6.14
Show that the method of this section gives the same solution as the
method of Section 6.2.4 by solving the equation
dy
dx +y = 3x.
Example 6.15
We have discussed damped harmonic motion in Example 6.12 and have
interpreted the damping as a dissipation of kinetic and potential energy
as heat. A common physical situation is when such oscillation is
sustained by a forcing function in the form of a periodic input of energy
which may be applied mechanically, or as an applied alternating voltage
to an electrical circuit, or as incident radiation of a particular frequency
on a molecular system. In order to sustain oscillation the forcing
function must be 'in tune' with the oscillation; when we apply a
sinusoidal forcing function - F sin pt the equation becomes
dx d2 x .
force = -kx-Pdt-m dt 2 = -Fsmpt.
The complementary function will be as in Example 6.12. A particular
integral is sought by using a trial function of the same form as the right-
hand side of the equation, so we put
x = A cos pt + B sin pt,
dx
dt - Ap sin pt + Bp cos pt,
(6.9)
r(x+ 1) = to tXe-tdt
so that
1 3 3 3
= p + 1 yeO) - p+ p2 + P + 1
where we have used partial fractions for the last term. We now take the
inverse transform of each term. From a table of transforms, or by
evaluating the integral, we find
L(e- X ) = 1/(P+1)
which, together with (6.13), gives
y = y oe- x -3+3x+3e- x
= (Yo + 3)e- -3 + 3x,
X
Example 6.17
The Schrodinger equation in one dimension has the form
a 2t/! 8n 2m
ax 2 = - ~ (E - V)t/!.
which relate functions of frequency (w) and functions of time (t). The
complex exponential function is expressed in sine and cosine terms in
Section 1.10.
An appropriate text should be consulted for the derivation and use of
Fourier analysis and transforms. Their application in spectroscopy is to
transform measurements of the decay with time of the response
induced by an irradiating pulse into the required spectrum as a function
of frequency. The mathematical and physical principles may be
brought together by observing, first, that a pulse is equivalent to a
superposition of different frequencies with the major components near
the frequency of the pulsed radiation, so that a spectrum of absorption
frequencies can all be excited simultaneously and, secondly, that the
resulting decay, or relaxation in time, that is observed can be
transformed into a superposition of exponential decays in time of a set
of frequencies, the set being the required spectrum.
CHAPTER 7
7.1 Significance
Experimental measurements must be made and used with their
probable accuracy in mind. Some properties can and should be
measured more precisely than others; for example, it is as absurd to
measure the length of a piece of string to 0.01 mm as it would be to
weigh a block of gold to the nearest kilogram. The least that should be
done is to quote only figures that are believed to be reliable, so that a
result given as 2.1 implies that it is between 2.05 and 2.15, and 2.1234
implies that it is between 2.12335 and 2.12345. It is preferable to give an
explicit statement of precision, such as 2.123 ± 0.024, the result and the
uncertainty both being quoted to the same number of decimal places.
This implies that the results have been analysed with great care because
two figures are quoted in the uncertainty, which is the most that can
ever be achieved. It does not imply that it is not possible to obtain
greater precision in the measurement because improved equipment and
better instruments may well produce a more precise result; it does
imply, however, that such measurements will not give results outside
the bounds set on the existing measurement. This chapter is concerned
with the methods of analysing experimental data which are necessary to
achieve such a high level of confidence in the statement of results.
Experimental errors are of two kinds, systematic and random. A
typical systematic error would arise if length were measured using a
ruler calibrated at one temperature but used at a higher one. All
measurements would then be short by an amount proportional to the
166 Experimental error and method of least squares
length being measured. Systematic errors arise from inadequacy in the
calibration of instruments, or from inadequacy in the design of
apparatus or in the technique used.
The term 'random' is used in error analysis in a particular sense.
Instability, due to poor control, mechanical vibration or thermal or
electrical noise, produces a scatter in the experimental results. Small
errors occur more often than large ones, and an error of given size will
usually be equally often positive as negative. Such errors can be
analysed statistically in terms of the distribution of error about a mean
value; this distribution is a curve as in Fig. 7.1, which shows the number
of times a particular error occurs as a function of the size of that error.
Such a curve could be drawn if a large enough number of repeated
measurements were made.
The simplest case is the one already mentioned, when errors are of
random sign. The curve is then symmetrical about a maximum, which is
the arithmetic mean. This is the normal, or Gaussian distribution of
error, discussed in Section 7.7. If significant systematic error is present
in addition to such random error, the mean will not be the true value
and the curve may not be symmetrica1.
Fig. 7.1 shows the probability that a particular result will be obtained
in the measurement of a quantity x. We do not therefore expect
repeated measurements to give the same result, and the quality of a set
of measurements is assessed by the spread of results obtained. The
uncertainty in the final result needs to be known and the conventional
way of expressing that uncertainty is discussed in the next section.
Since the methods used to assess the size of experimental error are
statistical, the sample size, which means the number of independent
measurements made, is very important. The statistical theory is
developed by assuming at first an infinitely large number of measure-
ments. If we then add the results together, we sum over such a large
number that the differences between adjacent results becomes so small
that we can replace sums by integrals. For this reason sums and
integrals occur in this chapter, depending on whether or not we are
assuming an infinite number of measurements.
x = LXi, (7.1)
n
~
I
>c'
~
...r...:
~
s
168 Experimental error and method of least squares
and this is the best result obtainable from the measurements. The
experimental error in a particular result Xi is then the difference
between that value and the arithmetic mean, (Xi - x). The final result of
the experiment is to be expressed as x ± Ax, where Ax is a conventional
assessment of the uncertainty of the measurement x of x.
In order that Ax shall be a useful measure of uncertainty it should
reflect the spread in the values of Xi obtained. We cannot use for this the
average value ofthe error (Xi - x) because that is zero by definition ofx.
We therefore use instead the square of the deviation, (Xi - X)2, and
define the root-mean-square error, s(x), by
s(x) = -J II (Xi - X)2 /n]. (7.2)
The relation between this and the standard deviation q(x) is discussed
in Section 7.6.
The root-mean-square error can be calculated most easily from a set
of values of X by use of the relation
S(X)2 = (l/n) L (Xi - X)2 = (l/n) L (xf - 2XiX + x 2)
= Lxt/n-2xLxdn+x2
= Lxt/n-x 2. (7.3)
f +OO
_ 00 p(x)dx = 1. (7.4)
x = _ f +OO
00 xp(x)dx. (7.5)
The statistical analysis of experimental data 169
The root-mean-square error O'(x) in x is then given by
f:: L+:
arithmetic mean value z of z is then
+OO f+oo
cr(y)2 =f -00 -00 (y - y)2p(X, y)dxdy, (7.10;
(z -Z) = (oz)
oX y
(x -x)+ (oz) (y _
oy x
y)+~2 [(02~)(X
ox
_X)2
where
The mean squares of the errors U(X)2 and U(y)2 are called the variances
in x and y, and the new term u(x, y) is called the covariance of x and y.
The significance of covariance is that it determines whether or not the
variables are independent of each other. If they are independent, the
integral in (7.15) separates into the product of two integrals each with
respect to one variable only, and both of these integrals are zero by
definition of the arithmetic mean. The covariance is therefore zero if the
variables are independent.
In general, if z is a function of independent variables x, y, u, ... the
formula for propagation of error is
and since
_ Xl X2
Xn=-+-+
n n
the integral separates into the sum of n integrals of the form
n 1 f+oo _ n U(X)2 U(X)2
L L
_ 2 2
U(Xn) = 2 (Xi -Xoo) p(x)dx = -2- = --,
i=ln -00 i=l n n
so that
_ U(X)
U(Xn) = Tn· (7.17)
the middle term having disappeared because L (Xi - x,,) = 0 for the
arithmetic mean. We can then write the required relation between s(x)
and u(x); by definition
(7.20)
f:
or since P(Xi - x) is symmetrical about Xi = x,
00 P(Xi - x)d(xi - x) = !.
This gives an integral of the standard form discussed in Example 5.1,
from which
K f: oo
exp[ -A(X i -x)2/2]d(xi -x) =K J(n/2A) =!
K = J(A.j2n)
and
P(Xi - x) = J (A/2n) exp [ - A(Xi - x)2/2].
= 0- 2 by Example 5.2.
Equation (7.26) describes the normal or Gaussian distribution of
error. We can expect experimental measurements to follow this
distribution so long as only random sign errors are present.
A theorem that is important in practice is the central limit theorem,
which states that if mean values are taken of samples having any
distribution, those means follow a normal distribution. Such averaging
can occur in the use of measuring instruments, so that we then obtain a
normal distribution of error even when the quantity being measured
does not itself follow that distribution.
(ii' Jij )
I? X
6-- - - ~-x
xi
I
I
Fig. 7.2
dD = (:~)dm+(~~)dC = o.
A solution of this equation will be
(:~) = (~~) = 0
giving
and since we may now omit the subscript i without causing confusion,
Lxy-mLx 2 -cLx = 0, (7.31)
Equation (7.34) shows that the least-squares line passes through the
'centre of gravity' of the points, (x, y), and since (7.32) may be written as
L(Yi-mxi-c) =0
we see that minimizing the sum of the squares of the residuals
incidentally makes the simple sum of the residuals also equal to zero.
This will not be unity even when a true straight line exists because
small-sample error means that expressions (7.35}-(7.37) are not the
same as (7.9), (7.10) and (7.15). A value of r that is less than unity is
therefore evidence either of limitations in the number and precision of
the points, or that the supposed linear relation does not apply; it is not
possible to separate these two possibilities for a given set of results.
A correlation coefficient that is much less than unity does not imply
independence of the two variables since it may be that a curve rather
than a straight line fits the data. This may be seen by plotting the graph,
and the variables should then be changed to convert a curve into a
straight line before applying the least-squares analysis.
Writing the mean value of byf as (1';, and assuming this to be the same
for all y;,
~( -)2 2 2 2
«(jm)l = L. x;-x (1'y = (1'y = nay (7.44)
[I(X;_X)2]2 I(X;-X)2 nIx 2 -(Ix)2
where
-
(1 - r2)[ n ~>2 (Iy)2]
(7.45)
n(n -2)
The divisor (n - 2) in (7.45) arises because two numerical coefficients, m
and c, are derived from the data leaving (n - 2) degrees of freedom.
When the least-squares line is used to estimate the value Yo of y
corresponding to the value Xo of x we have
Yo = y+m(x o -x)
so that
(bYO)2 = «(jy)2 + (xo _x)2«(jm)2.
Again assuming a mean value (1'; for all values of y, the error in the mean
value of y is given by equation (7.17)
(by)2 = (1'; In
which, together with (7.44), gives
S:)2
(oyo = (1' 2y
(1-+
n
(xo - X)2)
I(X;-X)2
(1'; (xo - X)2 n(1';
= -+-=.-'--:---=---'-;;-
n nIx 2 -(Lx)2
(7.46)
(7.47)
Example 7.1
The result of applying this analysis to a set of points is shown in Fig. 7.3.
The mean point (x, y) has error bars at ± a y/ -J n. The least-squares line,
a, is drawn with alternative slopes (b and e) of ± 15m. The curves d and e
are drawn to show the errors ± 15Yo at particular values of Xo as given by
equation (7.46). The error ± be in the intercept on the y-axis is due
mainly to the second term in the bracket in (7.47), corresponding to
uncertainty in the slope, with an additional contribution from the first
term which corresponds to the uncertainty in y. The correlation
coefficient for the points shown is r = 0.775; it can be seen that as Iowa
value as this indicates a considerable scatter in the points.
x
Fig. 7.3
The method of least squares 183
7.8.4 Least-squares straight line with both variables subject to error
The calculation of the least-squares line in Section 7.8.1 assumes that
the error is vested in only one of the two variables and equations (7.33)
and (7.34) for m and c assume the error to be in y. Experimental
conditions may justify this assumption, otherwise we might equally
well assume the error to be vested only in x; we would then obtain a
different straight line by interchanging x and y in the equations.
Since the least-squares line passes through the 'centre of gravity' of
the points, the two lines obtained by assuming the error to be vested
either in y or in x will intersect at that point. The ratio of the slopes of
these lines is related to the estimate r of the correlation coefficient.
From equation (7.38)
s(x, y) I(x -x)(y - y)
m = S(X)2 = I(x _X)2 (7.48)
SI units
SI stands for the agreed international system of units, which is based
upon the kilogram (kg), metre (m) and second (s), together with the
kelvin (K) unit of temperature, the mole (mol) for amount of substance
and the ampere (A). The candela (cd) unit ofluminous intensity is also a
basic unit. Other units are derived from these basic units.
The kelvin is defined as 1/273.16 of the thermodynamic temperature
of the triple point of water. This means any subsequent refinements in
absolute temperature measurement will alter the size of the degree and
so, say, the absolute temperature of the boiling point of water, without
changing the triple point. The ice point is 0.01 K below the triple point
so that O°C is 273.15 K.
The mole is the amount of substance containing as many elementary
units as there are atoms in 0.012 kg of 12c. This allows experimental
measurement of the Avogadro constant by, in effect, counting this
number of atoms, and of the gas constant from measurements of the
properties of gases.
The unit offorce is the newton (N), defined as that which produces an
acceleration of 1 m s - 2 in a mass of 1 kg. Pressure is then force per unit
area in N m - 2, this unit being called the pascal (Pa). Energy is measured
in joule (J), which is 1 N m, and power in watt (w) which is 1 J s - 1.
186 SI units, physical constants, conversion factors
The ampere (A) is that current which produces a force of
2 x 10- 7 N m -1 between long straight parallel conductors 1 m apart in
vacuum. The volt (V) is then the potential difference such that a current
of 1 A dissipates 1 watt, and the coulomb (C) is 1 A s. The faraday
constant (F) can then be measured as the number of coulombs of charge
carried by one avogadro of electrons.
Standard gravity is 9.80 665 m s - 2 and a standard atmosphere is the
pressure produced by a mercury column 0.76 m high under standard
gravity at ODC; this is 101.325 kPa. The preferred unit is the bar, defined
as 100 kPa. Other units that are still in use are the calorie in various
forms, usually the thermochemical calorie of 4.184 J, and the
Angstrom, A, which is 10- 10 m. The word 'litre' is now regarded as a
special name for the cubic decimetre, with the recommendation that
neither the word litre nor its symbol, 1, should be used to express results
of high precision. This is because the litre was previously defined as
1.000028 dm 3 , this definition being rescinded in 1964.
Physical constants
Recommended values of the fundamental physical constants are
published by the International Union of Pure and Applied Chemistry.
Some of the values given in Pure and Applied Chemistry, 51, 1 (1979)
are:
Avogadro constant L 6.022045(31) X 1023 mol- 1
Boltzmann constant k 1.380662(44) x 10- 23 J K- 1
gas constant R 8.31441(26) J K- 1 mol- 1
Planck constant h 6.626176(36) x 10- 34 J s
Faraday constant F 9.648456(27) X 104 C mol- 1
mass of proton 1.6726485(86) x 10- 27 kg
charge on electron e 1.6021892(46) X 10- 19 C
mass of electron 9.109534(47) X 10- 31 kg
speed of light in vacuum 2.99792458(1) X 108 m S-l
permittivity of vacuum 8.85418782(5) x 10- 12 C 2 N- 1 m- 2
normal atmosphere 1.01325 x 105 Pa exactly
zero of Celsius scale 273.15 K exactly
standard acceleration of
free fall 9.80665 m s - 2 exactly
The notation used is that the figure(s) in brackets is (are) the standard
deviation of the last figure(s) quoted.
Sf units, physical constants, conversion factors 187
Conversion factors for units
1 amp = I A = I J s - I V - I 1 inch = 2.540 cm
1 atm = 101.325 kPa 1 joule = I J = I N m
1 cal = 4.184 J 1m 3 = 106 cm 3
1 coulomb = 1 C = 1 A s = 1 J V-I 1 micron = 1 p. = I p.m = 10- 6 m
0.30 debye = 10- 30 C m I newton = I N = 1 m kg s - 2
1 dyne = 10- 5 N lpascaI=1 Pa=INm- 2 =lm- I kgs- 2
1 electron volt = 1 eV = 1.6021 x 10- 19 J 1 pound (UK) = 1 Ib = 0.4536 kg
1 erg = 10- 7 J I torr = 1 mmHg = 0.1333 kPa
1 farad = 1 F = 1 A S V-I 1 volt = 1 V = 1 J A -I S - I
= 1 m - 2 kg - I S4 A 2 = 1 m 2 kgs- 3 A-I
1 gallon (UK) = 4.546 litre I watt = 1W = 1J s - I = I m 2 kg s - 3
= 1.2009 gallon (USA)
A IX alpha N v nu
B
r
p beta
y gamma
-
0
~
0
xi
omicron
n
,
i\ (j delta 1t pi
E E epsilon P p rho
Z zeta 1: (J sigma
H '1 eta T T tau
e 0 theta Y u upsilon
I iota <l> 4> phi
K K kappa X X chi
i\ ). lambda 'I' psi
M p. mu n '"
OJ omega
188 SI units. physical constants, conversion factors
Summary of useful relations
cos 2 A + sin 2 A = 1
cos 2 A - sin 2 A = cos 2A
sin 2A = 2sinA cos A
sin (A±B) = sin A cos B±cos A sin B
cos (A ±B) = cos A cos B+ sin A sin B
eXeY = eX+ Y
a-X = 1jaX
if y = eX then x = In y
In x = loge x = 2.3026 log x = 2.30261og 10 X
log 1o lO-x = -x
b 1
ax 2 +bx+c = 0; x = --±- .j(b2 -4ac)
2a 2a
.
SIn x
x3
= x -3-'. +,- ...
XS
5.
n(n - 1)
(1 + xt = 1 + nx + 2! x 2 + ...
n(n -1)
(a+xt=an+nan-1x+ 2! an- 2 x 2 + ...
x2
f(x) = f(O) + xl' (0) + 2! f" (0) + ...
· f(x)
11m - () = 1·x~a I' (x) ·ff() () 0
1m ~() 1 a = 9 a = or CX)
x~a 9 X 9 X
dy dy du
dx du dx
SI units, physical constants, conversion/actors 189
e iy = cos y + i sin y
e- iy = cos y -i siny
Index