Você está na página 1de 125

F L E X I B L E L E A R N I N G A P P R O A C H T O P H Y S I C S

Module M6.4 Waves and partial differential equations


1 Opening items 3 The Schrödinger equation
1.1 Module introduction 3.1 The time-dependent Schrödinger equation
1.2 Fast track questions 3.2 The time-independent Schrödinger equation
1.3 Ready to study? 4 Closing items
2 Waves and oscillations 4.1 Module summary
2.1 Oscillations 4.2 Achievements
2.2 Waves 4.3 Exit test
2.3 Functions of two variables Exit module
2.4 The general wave
2.5 Partial derivatives
2.6 Higher partial derivatives
2.7 The wave equation
2.8 Deriving the wave equation
2.9 Standing waves

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
1 Opening items
1.1 Module introduction
The purpose of this module is to provide an introduction to two topics: waves and partial differentiation.
Waves occur in many physical situations, including vibrating strings, ripples on the surface of a lake, sound
waves, electromagnetic radiation and so on. The study of waves will lead us to functions of two variables
(in this case position x and time t) and to the derivatives of such functions. Functions of more than one variable
are common throughout physics and finding the derivatives of such functions requires a knowledge of partial
differentiation.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
The structure of this module is as follows. In Subsection 2.1 we consider oscillations and the differential
equation of simple harmonic motion (SHM). Subsection 2.2 introduces waves and draws your attention to the
distinction between waves and oscillations. In Subsection 2.3 we consider functions of two variables and such
functions are used in the Subsection 2.4 to describe a general wave which propagates without changing its shape.
Subsections 2.5 and 2.6 introduce first- and second-order partial derivatives, respectively. (The techniques given
here have wide applicability throughout physics and not just to the study of waves.) Subsection 2.7 derives the
partial differential equation for a wave propagating in one dimension without changing its shape, and
Subsection 2.8 shows how such a wave equation arises in the study of waves on a string. Subsection 2.9 shows
how standing waves can occur on a vibrating string with fixed ends. Section 3 is devoted to the time-dependent
and time-independent Schrödinger equations that arise in quantum mechanics. It explains the relationship
between these two equations, describes the significance of some of their ‘wave-like’ solutions (called wave
functions), and demonstrates the crucial role of partial derivatives in quantum mechanics.
Study comment

Having read the introduction you may feel that you are already familiar with the material covered by this module and that
you do not need to study it. If so, try the Fast track questions given in Subsection 1.2. If not, proceed directly to Ready to
study? in Subsection 1.3.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
1.2 Fast track questions
Study comment Can you answer the following Fast track questions. If you answer the questions successfully you need
only glance through the module before looking at the Module summary (Subsection 4.1) and the Achievements listed in
Subsection 4.2. If you are sure that you can meet each of these achievements, try the Exit test in Subsection 4.3. If you have
difficulty with only one or two of the questions, you should follow the guidance given in the answers and read the relevant
parts of the module. However, if you have difficulty with more than two of the Exit questions you are strongly advised to
study the whole module.

Question F1
The functions F1 (x, y) and F2 (x, y) are defined by
∂φ ( x, y) ∂φ ( x, y)
F1 ( x, y) = − and F2 ( x, y) = −
∂x ∂y
q
where φ ( x, y ) = . ☞
4 πε 0 x 2 + y2

Find F12 + F22 .

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question F2
1
Show that f(x − v t) is a solution of the wave equation
∂2 f 1 ∂2 f
− 2 = 0
∂x 2 v ∂t 2

Question F3
The one-dimensional time-dependent Schrödinger equation for a particle of mass m with a potential energy
function U(x, t) is
˙2 ∂ 2 Ψ ( x, t ) ∂Ψ ( x, t )
− + U ( x, t ) Ψ ( x, t ) = i˙
2m ∂x 2 ∂t
Show that Ψ ( x, t ) = exp ( − mωx 2 (2 ˙) ) exp ( −iEt ˙) ☞
mω 2 x 2
is a solution of this equation with U ( x, t ) = where ω is a constant (so that U is a function of the single
2
variable x), and find an expression for E in terms of ω.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Study comment Having seen the Fast track questions you may feel that it would be wiser to follow the normal route
through the module and to proceed directly to Ready to study? in Subsection 1.3.

Alternatively, you may still be sufficiently comfortable with the material covered by the module to proceed directly to the
Closing items.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
1.3 Ready to study?
Study comment To begin the study of this module you need to be familiar with the following topics: the trigonometric
functions and trigonometric identities (specific identities will be supplied when needed); Cartesian coordinates and graph
sketching (including the concept of a tangent to a graph, and the determination of the gradient of a straight line); the
differentiation of functions of one variable (including the definition of a derivative in terms of a limit and the rules of
differentiation, particularly the chain rule); the derivatives of the trigonometric and exponential functions, polynomials and
powers; differential equations (including the role of boundary conditions in determining a solution to such an equation), and
for Section 3, the properties of complex numbers (including the rules for their addition and multiplication, and the meaning
of the modulus of a complex number). You need to have some acquaintance with the basic ideas of Newtonian mechanics,
including energy, momentum, power and Newton’s laws of motion; it would also be useful to have had some previous
experience of waves (including their wavelength, frequency and speed) since this module only provides a brief review of
their properties. In a similar spirit it would be helpful, but not essential, to have some knowledge of simple harmonic motion,
and Hooke’s law. Finally, you will also need to have some idea of the everyday meaning of the term probability, though you
are not expected to be familiar with any of the mathematical properties of probability. The following Ready to study
questions will help you to establish whether you need to review some of the above topics before embarking on this module.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question R1
Differentiate the following functions with respect to x:
(a) f0(x) = 3x 2 + 4x5
(b) g ( x ) = exp( λ x 3 ) , where λ is a constant,
1
(c) h(x) = cos (x + 3x3 )
x +1
(d) w ( x ) =
x2 + 2

(e) u( x ) = ( x 2 + c 2 ) , where c is a constant,


2

(f) v( x ) = [ x 2 + sin ( yx ) ] , where y is a constant,


3

(g) p( x ) = sin( ax − bt ) , where a, b and t are constants.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question R2
Show that:
x3 k dy 3 y
(a) if k is an arbitrary constant, y ( x ) = + 3 is a solution of the differential equation + = x2 .
6 x dx x
(b) if a and b are arbitrary constants, y (x) = a e −x + be−2x is a solution of the differential equation
d2y dy
2
+3 + 2y = 0.
dx dx

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question R3

 ( a + ∆x )2 − a 2 
(a) Evaluate the limit lim   . What does this limit represent?
∆x →0  ∆x 
dy
(b) If y = f0(x) then the derivative is sometimes written alternatively as f ′ ( x ).
dx
 f ′ ( x + ∆x ) − f ′ ( x ) 
What does the limit lim   represent?
∆x →0  ∆x

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
2 Waves and oscillations
2.1 Oscillations
An oscillation is said to occur when some quantity (such as position) repeatedly cycles about an equilibrium
value.

Although oscillations are often taken to be sinusoidal ☞ (see Figure 1), this is not necessarily the case and other
forms (which arise, for example, in the context of electric circuits) are shown in Figure 2 and Figure 3.

y y y

t t t

Figure 1 3Sinusoidal oscillation. Figure 2 3Triangular oscillation. Figure 33Square-wave oscillation.


FLAP M6.4 Waves and partial differential equations
COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
A typical physical example of an oscillating system is a mass on a
spring after the mass has been displaced from its equilibrium position
and released (as in Figure 4). After its release, the displacement of such
a mass exhibits an approximately sinusoidal oscillation similar to that in
Figure 1. Hooke’s law is often used to provide a mathematical model of
a spring, and using it we can derive an equation which describes the
motion of the mass. So-called ‘ideal springs’, which satisfy Hooke’s
law, have the property that the extension of the spring y, is proportional
to the restoring force F y that causes the mass to return towards its weight
(mass m)
equilibrium position. Thus, for such a spring
Fy = − k0y (1) y

where k is a (positive) constant. displaced


position at
If m is the mass, then Newton’s second law allows us to replace Fy time t
d 2 y(t )
by m
dt 2
, which shows that y(t) satisfies 3
Figure 4 An oscillating system; a mass
on a spring.
d 2 y(t )
m = − ky ( t ) (2)
dt 2

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
If we introduce the positive constant ω defined by ω = k / m we may rewrite Equation 2
d2y (t )
m = − ky ( t ) (Eqn 2)
dt 2
in the form
d 2 y(t )
= −ω 2 y(t ) (3)
dt 2
Equation 3 is a differential equation. It can be solved for y(t), and the general solution is
1 1 1 1
y(t) = A sin (ω1t) + B cos (ω1 t) (4)
where A and B are arbitrary constants. This general solution can also be written in the alternative
(and often more useful form)
1 1
y(t) = C sin (ω0t + φ) (5)

where C and φ are arbitrary constants, and we still have ω = k/m.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Equation 5
1 1
y(t) = C sin (ω0t + φ) (Eqn 5)
is an especially valuable form of the solution because it makes it easy to identify the physical significance of
each of the terms.
C is the amplitude and represents the maximum value of the displacement from the equilibrium position;

(ω1 t + φ) is called the phase and determines the stage that the oscillation has reached in its cycle at time t;

ω= k / m is called the angular frequency and is 2π times the number of oscillations completed per second;

φ is called the phase constant (or the initial phase) since it determines the value of the phase at t = 0, and hence
the displacement from equilibrium at that time (y(0) = Csinφ).
The sort of oscillatory motion described by Equation 5, and characterized by the parameters C, ω and φ , is
known as simple harmonic motion.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1

1 1
Show that y(t) = C sin (ω0t +φ) is a solution of the differential equation
d 2 y(t )
dt 2
= −ω 2 y(t )

Other examples of physical systems which display oscillatory behaviour include the simple pendulum and some
a.c. circuits, but the basic idea remains the same; some physical quantity, for example position, voltage or
current, cycles repeatedly about an equilibrium value.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
2.2 Waves

1 1
Perhaps the first thing that comes to mind when you see the word ‘wave’ is the wave that you see at the beach.
The undulations of the surface of the sea occur in a regular fashion, and appear to move towards the shore —
a phenomenon that surfers are able to exploit to the full. We say that the surface ‘appears to move’ because the
actual motion of the water is a good deal more complicated than you would at first think. Clearly something
moves towards the shore, but it is certainly not a coherent body of water, for otherwise the entire ocean would
end up on the beach. As further evidence of this you need only note that a swimmer tends to bob up and down as
a wave passes, and this would certainly not happen if the velocity of the water at each point in the wave was
directed towards the shore.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
There is a well known Tom and Jerry cartoon that illustrates what is y
t1
really happening. Tom is seated at a grand piano and runs his thumb
rapidly along the keys from left to right, while Jerry is inside the piano
and moves along with the wave of displaced hammers (see Figure 5).
The essential point to appreciate is that the hammers are simply
moving up and down, while the wave moves from left to right. The
wave is a travelling disturbance in the position of each hammer, not a
travelling set of hammers. x
(a)
A water wave is a very similar phenomenon. The motion of the water
is predominantly vertical and oscillatory, while the motion of the wave
y
is horizontal and progressive. t2 > t 1

3
Figure 5 Snapshots at two different times of the heights of all the hammers
in a piano show how vertical movements can produce a wave that travels
horizontally. Notice particularly that the vertical displacement y at any position
x (measured from the left-hand end of the keyboard) is dependent on two x
factors: the particular hammer and the time, i.e. the value of x and the value of (b)
t. The time t2 in (b) is later than t1 in (a).

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Another place where we are used to seeing waves is on strings, ropes and cables. These waves take many forms.
For instance, if you hold one end of a long string and give it a single flick, then a solitary wave (often referred to
as a pulse) will travel along the string. On the other hand if you jerk the string up and down repeatedly in a
regular way, you may produce a repetitive wave that travels along the string. Note that the disturbance does not
have to be repetitive in order to qualify as a wave. Both kinds of disturbance travel along the string, so both are
examples of travelling waves.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
y for a fixed value of t
Although travelling waves do not have to be repetitive, there can be
no doubt that the best known and most intensively studied kind of
travelling wave is one which has a sinusoidal shape at each
ymax λ
successive instant of time. Figure 6 attempts to illustrate such a
sinusoidal wave, such as one travelling along a string. As you can
see, the figure is in two parts. Figure 6a shows the displacement y at
all points x at a fixed time t; in effect, this is a ‘snapshot’ of the wave. x/m
Figure 6b, on the other hand, shows the way in which the
displacement y changes with time at a fixed position x. (a)
If you compare Figure 6b with Figure 1 you will see that at a fixed
value of x the disturbance caused by the sinusoidal travelling wave is
just a sinusoidal oscillation. Indeed, one way of picturing a y for a fixed value of x
sinusoidal travelling wave is as an array of simple harmonic
oscillators, where each oscillator is slightly out of phase, i.e. has a T
slightly different initial phase compared with its neighbours. ymax

3
1 11
Figure 6 (a) The ‘shape’ of a sinusoidal travelling wave at a particular t

1 1
instant of time — the wave profile. (b) The oscillation caused by a
sinusoidal travelling wave at a particular point — the wave form.
(b)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
y for a fixed value of t
A ‘snapshot’ of a wave at a particular instant, such as Figure 6a, is
generally referred to as a wave profile. A trace showing the changing
disturbance caused by a wave at a fixed point, such as Figure 6b is
ymax λ
called a wave form. Both of these terms will be used freely in the
rest of this module. Remembering that both are needed to
characterize a wave will help you to remember the distinction
between a wave and an oscillation. If you were to photograph a x/m
travelling wave on a string you would obtain a wave profile.
To determine the wave form you would have to record the (a)
displacement of a particular segment of string at different times.
y for a fixed value of x

T
ymax

3
1 11
Figure 6 (a) The ‘shape’ of a sinusoidal travelling wave at a particular
t
1 1
instant of time — the wave profile. (b) The oscillation caused by a sinusoidal
travelling wave at a particular point — the wave form.
(b)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
The need for both a wave profile and a wave form in order to characterize a wave points to the essential
mathematical difference between a wave and an oscillation. In the case of an oscillating mass, the displacement
from the equilibrium position y, is only a function of time; so it may be represented by y(t). On the other hand,
for a wave on a string the displacement from the equilibrium position is a function of position and time; so it
may be represented by y(x, t). Thus, even when it only travels in one dimension (x), the wave disturbance y is a
function of two variables. We shall consider such functions in more detail in Subsection 2.3, but first we shall
finish this subsection by briefly introducing some of the parameters commonly used to describe waves.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
y for a fixed value of t
If we stay with our example of a wave on a string, then the
amplitude A is the maximum displacement (in the y-direction) of the
string from its equilibrium position. The wavelength λ is the ymax λ
distance between two successive peaks (or troughs) on the wave
profile, as shown in Figure 6a. ☞ One might determine the
wavelength by photographing the string, and then measuring the
distance between the peaks. The period T is the time between two x/m
successive peaks (or troughs) on the wave form as shown in
Figure 6b, and this might be found by timing the interval between (a)
peaks at a fixed position on the string. The frequency f is the number
of troughs (or peaks) which pass a fixed location in one second, and y for a fixed value of x
is related to the period by
T
1 ymax
f = (6)
T

3
1 11
Figure 6 (a) The ‘shape’ of a sinusoidal travelling wave at a particular t

1 1
instant of time — the wave profile. (b) The oscillation caused by a sinusoidal
travelling wave at a particular point — the wave form.
(b)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
The units of frequency will therefore be s−1, which are called hertz (Hz) in SI units. Since f is the rate at which
peaks pass a fixed point, and since the distance between successive peaks is λ , it follows that the
speed of propagation v at which the waves move along the string is given by

v = fλ1 (7)

The angular frequency ω is related to the period by the equation


ω = 2π f = (8)
T

so ω = 2πf, and ω is in units of rad1s−1.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Similarly, the wavenumber σ and the angular wavenumber k are related to the wavelength λ by the equations

1
σ = (9)
λ

and k = (10) ☞
λ

It is probably worthwhile to try to remember Equations 6 to 10,


1
f = (Eqn 6)
T
v = fλ1 (Eqn 7)

ω = 2π f = (Eqn 8)
T
since there are many useful relationships that can be derived from them. The basic definitions are summarized in
Table 1 on the next page.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Table 1 3The parameters used to describe a wave.
Amplitude A The maximum displacement from equilibrium,
see Figure 6a or 6b.
Wavelength λ The distance that separates adjacent equivalent
points on the wave profile, see Figure 6a.
Period T The time that separates successive equivalent
points on the wave form, see Figure 6b.
Frequency f = 1/T
Angular frequency ω = 2πf = 2π/T
Wavenumber σ = 1/λ
Angular wavenumber k = 2π/λ
Speed of propagation v = fλ = ω/k

✦ Use the definitions that have been presented to show that v = ω0/k.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
y for a fixed value of t
Waves which cause a disturbance perpendicular to their direction of
propagation are known as transverse waves. Waves on a string,
ripples on a pond and electromagnetic waves are all examples of
ymax λ
transverse waves, but not all waves are transverse.
Sound waves in air are caused by compression and rarefaction of the
atmosphere as molecules move back and forth along the direction of
propagation. Such waves are known as longitudinal waves. x/m
For a sound wave, we can still represent the displacement from the
equilibrium position by y and draw graphs similar to Figure 6, but x (a)
and y are not physically perpendicular in that case.
y for a fixed value of x

T
ymax

3
1 11
Figure 6 (a) The ‘shape’ of a sinusoidal travelling wave at a particular
t
1 1
instant of time — the wave profile. (b) The oscillation caused by a sinusoidal
travelling wave at a particular point — the wave form.
(b)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
2.3 Functions of two variables
As we have seen, if y is the displacement from the equilibrium position for a one-dimensional wave, then y is a
function of two variables, the position x, and the time t. In order to emphasize that y is a function of both x and
t we may write it as y(x, t). Here is an example of such a function:

1 1
y(x, t) = A sin (at − bx) (11) ☞
where A = 2.371m, a = 161s −1, 1
and b = 3.23 m−1. Note that the units of a and b are such that when we substitute
values for t and x (in seconds and metres respectively) the argument of the sine function will be dimensionless,
as it should be.

✦ 1 1
Evaluate y(1.9 m, 12.0 s) from the definition of y(x, t) given in Equation 11. ☞

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Functions of two variables occur quite often in physical situations, ☞ h
here are some examples.
1 Suppose we want a mathematically precise description of a hill or a
mountain range. We could obtain such a description by first using a y
two-dimensional coordinate system to specify points on a horizontal
reference surface (e.g. sea-level), and then using a function of two
variables h(x, y), to represent the height of the Earth’s surface above
sea-level at a point with coordinates (x, y).
A function of this kind is difficult to illustrate on a sheet of paper but a
useful picture can often be obtained by means of a three-dimensional
graph of the sort shown in Figure 7.

3
Figure 7 The height h(x, y) above
the (x, y) plane.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
2 If a fixed electric charge q (measured in coulombs) is located at the φ
origin of a Cartesian coordinate system the electrostatic potential
at any point in the (x, y) plane with coordinates (x, y) is given by
q
φ ( x, y ) = (12)
4πε 0 x 2 + y 2
where φ is a function of the two variables x and y, and ε 0 is a
constant. This function is shown in Figure 8.

3
Figure 8 The function φ(x, y) of
Equation 12.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
3 The current I, the voltage V, and the resistance R in a circuit such
as that shown in Figure 9a are related by Ohm’s law
V
I = (13) I
R
In a situation in which V and R can be varied independently, I is a
function of these two variables and can be written as I(V , R).
This function is shown in Figure 9b.

I V

R
R
(a) (b)
Figure 9 3The current I as a function of V and R.
FLAP M6.4 Waves and partial differential equations
COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
It is sometimes convenient to display the arguments of the same function in different ways.
q
For example, if φ ( x, y ) = , as in Equation 12,
4 πε 0 x 2 + y 2
then, although φ is a function of the two variables x and y, it can also be considered as a function of the single
variable u = x2 + y2, and we can write
q
φ (u) = (14)
4 πε 0 u
where u = x2 + y2. In such a case we may avoid having to introduce the extra variable u by writing the function φ
in the form

φ ( x 2 + y2 ) =
q
(15) ☞
4 πε 0 x 2 + y 2

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1


If y ( x, t ) = sin  ( x − vt )  write y as a function of a single variable.
 λ

You will see shortly that the above discussion is relevant to the study of waves.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
2.4 The general wave
In this module we are interested in waves for which the displacement from the equilibrium position depends on
the time (t) and a single space variable (x). Such waves are called one-dimensional waves, and a typical
example is provided by a wave on a string of the kind we considered earlier. In general, a one-dimensional wave
consists of any disturbance which moves along a line, but we are more often interested in waves which maintain
their shape as they move with a constant speed. Our first example is a disturbance in the form of a parabola.

✦ Sketch the graphs of the following functions


1
(a) y(x) = (x − vt)2 when t = 0 s and v = 1 m s−111
1
(b) y(x) = (x − vt)2 when t = 1 s and v = 1 m s−1 1 1
1
(c) y(x) = (x − vt)2 when t = 2 s and v = 1 m s−11 1
Does the ‘pulse’ move to the left or to the right as t increases? In which direction would the pulse move if
( x − vt )2 was replaced by ( x + vt )2 ?

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
The continuous curve in Figure 11 shows a snapshot of a solitary wave y vT
pulse at some initial time t = 0, when it has a profile given by y = f0(x).
The dashed curve shows another snapshot of the same wave at a later
time t = T where T > 0. If the wave moves to the right at constant speed
v, without changing its shape, then during the time T that elapses
between the two snapshots each part of the wave will travel a distance
vT to the right. It follows that the disturbance y at any position x at time 0 vT x
t = T will be equal to the disturbance that would have existed at x − vT
at t = 0. Hence, at t = T the wave profile is given by y = f0(x − vT), 3
Figure 11 A general wave moving a
where f is the same function that described the wave profile at t = 0. distance vT to the right.
Of course, there was nothing special about the value of T that we chose,
so we can say that at any positive time t (including t = 0) any wave that
moves in the positive x-direction with unchanging shape and constant
speed v can be represented by
y = f0(x − vt)
Similarly a wave that moves to the left with unchanging shape can be represented by
y = f0(x + vt)
where y = f(x) describes the profile of the wave at t = 0.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
0
Question T1
A function θ (x) is defined by
0 for x < 0
θ(x) =  0 (16)
1 for x ≥ 0
and a function y (x, t) is defined by y (x , t ) = θ ( x − v t). Sketch graphs of y, as a function of x, for
1 1 1 1
vt = 0 m, 1 m, 2 m. Repeat the exercise for y(x, t) = θ (x + vt). ❏ 3
To a physicist, sinusoidal waves are of paramount interest for several reasons:
1 Many waves that occur in physics are sinusoidal, including monochromatic light (i.e. of a single
wavelength), sound generated by a fixed frequency oscillator and radio waves (when unmodulated by
speech, etc.).
2 Sinusoidal waves are comparatively simple to analyse.
3 In most cases, a complicated wave can be considered as the sum of sinusoidal waves of different frequency.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
A sinusoidal wave, moving to the right with constant speed v, as t increases, may be written in the general form

y ( x, t ) = A sin[ k ( x − vt ) + φ ] (17) ☞
where the constants A, k and φ, respectively, represent the amplitude of the wave, its angular wavenumber and
its phase constant.
Unless we are comparing two waves, the phase constant is of little interest and we may assume that it is zero; in
which case Equation 17 may be written in any one of the following equivalent forms (using Equations 6 to 10,
which are repeated for your convenience):

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
y( x, t ) = A sin [ k ( x − vt ) ] (18) f =
1
4 (Eqn 6)
T
y( x, t ) = A sin [ kx − ω t ] (19) v = f λ4 (Eqn 7)

y( x, t ) = A sin [ 2 π( σ x − f t ) ] (20) ω = 2π T 4 (Eqn 8)


1
  σ=
y( x, t ) = A sin  2 π  − f t  
x (Eqn 9)
(21) λ
 λ 

 t  k= 4
y( x, t ) = A sin  2 π  −  
x (Eqn 10)
(22) λ
  λ T

Each of these forms has its uses, but Equation 19 is probably the most practical, and you can use Equations 6
to 10 to generate the others when you need them. Figure 12 (on the next page) illustrates the function given in
Equation 19. Notice that each fixed value of t corresponds to a sinusoidal curve drawn on the surface. Such a
curve represents the wave profile at that particular value of t (i.e. a plot of y against x at a fixed value of t).
If we were to take a slightly larger fixed value of t then the wave profile would move to the right. A fixed value
of x also produces a sinusoidal curve on the surface, but this time it corresponds to the wave form.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
y (x, t)
1

−A
0

λ
Figure 124 T
A three-dimensional graph
of the function 2λ
y( x , t ) = A sin [ k x − ω t ] .

x 0 t

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
2.5 Partial derivatives
Given a function of a single variable, f0(x), its rate of change at x = a is equal to the gradient of the graph y = f0(x)
at x = a, ☞ and may be determined by evaluating the derivative dy/dx = f0 ′(x) at x = a. The rate of change of a
function of one variable is an important concept because it often arises in mathematical models of physical
systems, notably in the context of differential equations.
Rates of change are also important for functions of two variables such as f0(x, t), but in that context the presence
of more than one independent variable introduces additional complications. Functions of two variables may have
many different rates of change at a given point, and care must be taken to distinguish between them. In order to
do this we need to extend the notion of an ordinary derivative to that of a partial derivative. The main purpose of
this subsection is to introduce you to these partial derivatives, but before doing so we need to review the rules
normally employed in the differentiation of a function of one variable.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
The derivative of a function of a single variable
The rules of differentiation for a function of one variable are the following:
d ( cf ( x ) ) df ( x )
The constant multiple rule:4 =c
dx dx
d ( f ( x ) + g( x ) ) df ( x ) dg( x )
The sum rule:4 = +
dx dx dx
d ( f ( x )g( x ) ) dg( x ) df ( x )
The product rule:4 = f (x) + g( x )
dx dx dx
d  f ( x )  g( x ) df dx − f ( x ) dg dx
The quotient rule:4  =
dx  g( x )  [ g( x ) ]2
df (u( x )) df (u) du( x )
The chain rule:4 =
dx du dx

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
As an example of the chain rule, suppose that we are given a function f(u) and that we wish to differentiate
y = f0(x2 + 3x) with respect to x.
Using the chain rule we can write
dy df du
=
dx du dx
where u = x2 + 3x, and this gives
dy du df df
= = ( 2 x + 3) (23)
dx dx du du u = x 2 +3x

df
where the final term on the right means that must be evaluated for u = x2 + 3x.
du
The notation on the right of Equation 23 is rather cumbersome, and it is often preferable to use the notation
df
f ′ (u) in place of , in which case Equation 23 becomes
du
dy du df
= = ( 2 x + 3) f ′ ( x 2 + 3 x ) (24)
dx dx du

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Notice that this notation also has the advantage that we can avoid introducing the unnecessary variable u.
A more general statement of this form of the chain rule then becomes
d
f ( g( x ) ) = g ′ ( x ) f ′ ( g( x ) ) (25)
dx
(In the example on the previous page, g(x) = x2 + 3x so that in this notation g ′ ( x ) = 2 x + 3 .)

✦ Using the chain rule, write down the derivative with respect to x of each of the following functions:

1
(a) y = sin (3x + 5),

(b) y = sin1(ax + b), where a and b are constants

(c) y = sin1(kx − ωt) where k, ω and t are constants,

(d) y = f0(3x + 5) where f0(u) is an arbitrary function

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1

(a) Given that f0(x) = (x2 + 5x + 2)3 write down an expression for f0′ (x).

1
(b) Given that φ(t) = cos (at + b) where a and b are constants, write down an expression for φ0′(t).

dy
(a) Given that y = f(kx − ω0t) where k, ω and t are constants, write in terms of f0′.
dx

dy
(b) Given that y = f(kx − ω0t) where k, ω and x are constants, write in terms of f0′.
dt

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
The derivatives of a function of two variables
The last question, in which first x and then t was a constant indicates an important feature of the behaviour of
functions such as F(x, t) in which both x and t are variables. Such functions may generally have several different
derivatives. We can find one derivative by treating t as a constant and differentiating the function with respect to
x, and we can find an entirely different derivative by treating x as a constant and differentiating with respect to t.
This technique of holding one variable constant and differentiating with respect to another is called
partial differentiation and the resulting derivatives are called partial derivatives. Such derivatives are
distinguished from ordinary derivatives by using a special kind of ‘curly’ dee (∂0) when writing them. Thus, for a
given function F(x, t) we can introduce the two derivatives:
∂F  ∂F  ∂F
which may be abbreviated to  
 ∂x  t
☞ or more briefly and
∂x at constant t ∂x

∂F  ∂F  ∂F
which may be abbreviated to  
 ∂t  x
☞ or more briefly .
∂t at constant x ∂t

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
∂F
If we interpret the function y = F(x, t) as a wave then, at any given point (x, t), the partial derivative
∂x
∂F
represents the gradient of the relevant wave profile at the given value of x, while represents the gradient of
∂t
∂F ∂F
the relevant wave form at the given value of t. ☞ In other words, and represent the slopes of the three-
∂x ∂t
dimensional graph of y = F(x, t) in the direction of the x-axis and the t-axis respectively, at the given point (x, t).
These gradients in the x- and t-directions are indicated in Figure 13 on the next page. (Clearly, they are just two
of an infinite number of gradients that might be found at any given point, since we can identify an infinite
number of directions between the x- and t-directions.)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
the gradient of this
line is the value of
∂y
at x = xA, t = tA
∂x 1
y
0.5

0 A

−0.5
−1 5
4
3 4

3
2
Figure 134The graphical xA 2
tA the gradient of this
interpretation of the partial
1 line is the value of
derivatives with respect to 1 ∂y
x and t of a function at x = xA, t = tA
y = F(x, t) at a given point. ∂t
x 0 0 t

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1

∂F ∂F
Given that F( x, t ) = ( a sin k x + b cos ω t ) find
3
and and evaluate them at x = 0, t = 0.
∂x ∂t

Question T2
∂y ∂y
Given that y( x, t ) = A sin [ k x − ω t ] find and .
∂x ∂t
∂y ∂y
What is the physical significance of and if y(x, t) represents the wave of hammers moving inside the
∂x ∂t
piano in the Tom and Jerry cartoon?4❏

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
 ∂f 
Notice particularly that ∂ is used rather than d in the partial derivative   , and the t subscript can be used to
 ∂x  t
 ∂f 
emphasize that t is to be held constant. Similarly the x subscript in   can be used to emphasize that x is
 ∂t  x
∂f ∂f
held constant. Usually we shall write the partial derivatives as and , and drop the suffices. ☞
∂x ∂t
Finding partial derivatives
When dealing with partial derivatives we are by no means restricted to functions of the variables x and t.
We might, for example, choose to use x and y as the independent variables, and write
z = f ( x, y)
No matter what symbols we use to represent the variables, the process of finding partial derivatives involves
similar techniques to finding ordinary derivatives of functions of one variable, the only difference being that we
have to remember to treat one variable as a constant.
An example should help to clarify this.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Example 1
 ∂f   ∂f 
If f0(x, y) = x2 − y2 , find   and  
 ∂x  y  ∂y  x

 ∂f  ∂ ( x 2 − y2 ) ∂ ( x 2 ) ∂ ( y2 ) d ( x2 )
Solution   = = − = = 2x
 ∂x  y ∂x ∂x ∂x dx

 ∂f  ∂ ( x 2 − y2 ) ∂ ( x 2 ) ∂ ( y2 ) d ( y2 )
and   = = − = − = −2 y4❏
 ∂y  x ∂y ∂y ∂y dy
There are two points to note in this example.
First, when we differentiate a function of y alone with respect to x we get zero, because y is treated as a constant.
∂ ( y2 ) ∂ ( x2 )
In this particular case we have = 0 , and similarly = 0.
∂x ∂y
Second, when differentiating a function of x alone with respect to x, the partial derivative and the ordinary
∂ ( x2 ) d ( x2 )
derivative are identical, and for this reason we can replace ∂ by d, so that, for example, = = 2x.
∂x dx

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Example 2
∂f ∂f
Given that f0(x, y) = e 2x cos(3y), find and .
∂x ∂y

d ( e2 x )
Solution
∂f
∂x
=
dx
1 1
cos ( 3 y ) = 2e2x cos (3y)

d ( cos ( 3 y ) )
and111
∂f
∂y
= e2 x
dy
1 1
= −3e2x sin (3y)4❏

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question T3
∂f ∂f
Given that f0(x, y) = x 2 − xy + 2y2, find and .4❏
∂x ∂y

Question T4

1 1 1
Given that f0(x, y) = x 2 1sin y + y2 cos x, find
∂f
∂x
and
∂f
∂y
.4❏

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
It should be clear from these examples and questions that most of the rules of ordinary differentiation are
directly applicable to partial differentiation. However, the extension of the chain rule is a little more
complicated. Before using it in the context of partial differentiation let us briefly return to its application in the
ordinary differentiation of a function of one variable.

1
Given that f0(x) = sin (3x2 ), find
df ( x )
dx
.

1 1
Given that φ (y) = sin (c2y4) where c is constant, find
dφ (y)
dy
.

Now let us apply the same approach to a function of two variables. Naturally, we shall have to modify the chain
rule somewhat to apply it in this case.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Example 3
∂f ∂f
Given that f0(x, y) = sin(x 2 0y4), find and .
∂x ∂y
Solution
1
In this case, if we write f0(x, y) = sin (u), then u = x2 y4 is a function of two variables and the chain rule takes the
form ☞
∂f
=
∂ x 14
d (sin(u)) ∂ u
du244 ∂3x
1
= cos (u) × 2xy4 = 2 x y 4 cos(x{
123
2 4
y )
4 ∂u ∂x u
Using the chain rule

and
∂f
∂y
=
d (sin(u)) ∂ u
du ∂y
1 123
1
= cos (u) × 4x2 y3 = 4x 2 y 3 cos(x{
2 4
y )4❏
14 4244 3 ∂u ∂y u
Using the chain rule

In the above example, note that we use d when differentiating a function of one variable with respect to that
variable, and we use ∂ when differentiating a function of several variables. Apart from that the chain rule is
essentially unchanged.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question T5

1
Given that f0(x, t) = cosh (Ax2 t3), ☞ where A is constant, find
∂f
∂x
and
∂f
∂t
.4❏

Question T6
 x ∂f ∂f
Given that f ( x, y ) = log e   find x +y .4❏
 y ∂x ∂y

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Formal definitions
The ordinary derivative df/dx of a function of a single variable f(x) is formally defined by the following limit,
provided that the limit exists:
df  f ( x + ∆x ) − f ( x ) 
= lim  
dx ∆x →0  ∆x 
In a similar way, the partial derivatives of a function of two variables f(x, y) may be defined by ☞
 ∂f   f ( x + ∆x, y ) − f ( x, y ) 
  = lim  
 ∂x  y ∆x →0  ∆x 

 ∂f   f ( x, y + ∆y ) − f ( x, y ) 
  = ∆y→0
lim  
 ∂y  x  ∆y 

These definitions are mainly of interest to mathematicians, but you should be able to recognize them since they
can easily arise in a physical context.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
2.6 Higher partial derivatives
If we are told that f(x, y) = e2x cos(3y) then we may easily find the partial derivatives
∂f ∂f
= 2 e 2 x cos(3 y) and = −3e 2 x sin ( 3 y )
∂x ∂y
Now these partial derivatives are themselves functions of the two variables x and y, and so can be differentiated
partially with respect to x or y.
It is generally the case that a partial derivative of a function of two variables is also a function of these variables,
and if we choose we can differentiate a second time. ☞ For example:
If 1 1
f0(x, y) = e2x cos (3y)
∂f ∂f
then = 2 e 2 x cos(3 y) and = −3e 2 x sin ( 3 y )
∂x ∂y

so that

∂x
 ∂f 
  =
 ∂x 

∂x
( 2 e 2 x cos(3 y) ) = 4e2x cos(3y)1
∂  ∂f  ∂
while   =
∂y  ∂y  ∂y
( −3e 2 x sin(3 y) ) = −9e2x1cos(3y)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
It is common practice to abbreviate these second partial derivatives as follows:

∂  ∂f  ∂2 f
  = (26)
∂x  ∂x  ∂x 2
∂  ∂f  ∂2 f
and   = (27)
∂y  ∂y  ∂y2

These results are straightforward extensions of what happens with a function of one variable; here we simply
differentiate with respect to x (or y) while keeping y (or x) constant. However, there are two extra possibilities
which cannot occur with a function of a single variable.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Firstly, having found (in our earlier example)
∂ f  ∂ f ( x, y ) 
=  = 2e 2 x cos(3y)
∂ x  ∂ x  y= constant
144424443
Writing it in full to
remind you of what
it really means.

we could then differentiate the result with respect to y while keeping x constant. Doing this we obtain
∂  ∂f  ∂
  =
∂y  ∂x  ∂y
[ 2 e 2 x cos(3 y) ] = −6e 2 x sin(3 y) (28)

Notice that first we keep y constant and differentiate with respect to x and then we keep x constant and
differentiate with respect to y. (Notice also that we don’t use x and y subscripts to indicate which variable is
being held constant since doing so would make the notation unnecessarily complicated.) We normally abbreviate
these second derivatives as follows:

∂  ∂f  ∂2 f
  = (29)
∂y  ∂x  ∂ y∂ x

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Secondly however, we could differentiate in reverse order, so that in place of ∂2 f/ ∂y∂x we obtain
∂2 f ∂
∂ x∂ y
=
∂x
( −3e 2 x sin ( 3 y ) ) = −6e 2 x sin ( 3 y ) (30)

It is no coincidence that Equations 28 and 30 give the same result,


∂  ∂f  ∂
  =
∂y  ∂x  ∂y
[ 2 e 2 x cos(3 y) ] = −6e 2 x sin(3 y) (Eqn 28)

and in fact, for all the functions that you are likely to meet, it is generally true that

∂2 f ∂2 f
= (31)
∂ x∂ y ∂ y∂ x

The derivatives defined in Equations 29 and 31


∂  ∂f  ∂2 f
  = (Eqn 29)
∂y  ∂x  ∂ y∂ x
are called mixed partial derivatives.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Example 4

1
If V(x, y) is defined by V(x, y) = arctan (y/x) ☞ find
∂ 2V
∂x 2
and
∂ 2V
∂y2
.

∂ 2V ∂ 2V
Hence verify that V(x, y) satisfies + = 0 (32)
∂x 2 ∂y2
(An equation (such as Equation 32) which involves partial derivatives is known as a partial differential
equation. Equation 32 is known as Laplace’s equation (in two dimensions) and is crucial to the study of many
areas of theoretical physics, including electrostatics and hydrodynamics.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
1
Solution Using the standard derivative (derived elsewhere in FLAP)
d 1
arctan ( u ) = 4we obtain
du 1 + u2
∂V ∂  y  ∂  y
arctan    =
1
= 4from the chain rule
∂x ∂ x   x  y 
2
∂ x  x
1+
 x

× − 2  = − 2
1 y y
33111110 = (33)
2  x  x + y2
1+ 
y
 x

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
∂ 2V 2 xy
so =
∂x 2
( x 2 + y 2 )2
∂V 1 1 x
11 = × = (34)
∂y 2
+ y2
1+ 
x y x2
 x
∂ 2V 2xy
so = −
∂y 2
( x 2 + y 2 )2
∂ 2V ∂ 2V 2xy 2xy
It follows that + = 2 − = 04❏
∂x 2 ∂y2 ( x + y ) ( x + y 2 )2
2 2 2

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question T7

Given that V(x, y) = arctan 



1 y
x
, verify that
∂ 2V
∂ y∂ x
=
∂ 2V
∂ x∂ y
.4❏

Question T8

1 1
If V(r, θ) = (α0rn + β0r0−n) cos (nθ ) where α, β and n are constants, find
∂V ∂ 2V
,
∂r ∂r 2
and
∂ 2V
∂θ 2
.

Hence show that


∂ 2V 1 ∂V 1 ∂ 2V
+ + = 0 4❏ (35)
∂r 2 r ∂r r 2 ∂θ 2

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question T9
The pressure P, temperature T and volume V of a certain gas are related by van der Waals’ equation ☞
 P + a  V − b = RT

( ) (36)
V2 
where a, b and R are all constant. Regarding P as a function of V and T, find the values of P, V and T (in terms of
the constants a, b and R) for which the equations
∂P ∂ 2P
= 0 and = 0
∂V ∂V 2
are satisfied simultaneously.4❏

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
2.7 The wave equation
A general wave
In Subsection 2.4 we showed that a general one-dimensional wave that travels with unchanging shape at
constant speed v has the form y = f0(x ± vt). We now set out to find the equation for which this is a solution. ☞
For this restricted class of functions of two variables it is possible to simplify the process of finding the partial
derivatives. If we put u = x ± vt, then we can find the partial derivatives of f with respect to x in terms of the
derivative with respect to u. ☞

1
Given that f0(x, t) = sin (x + vt), find
∂f
∂x
, then put u = x + vt and calculate
df
du
.

∂f df
Hence show that = in this case.
∂x du

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
∂f df ∂u
It is no accident that = in the previous exercise. If u = x ± vt we have = 1, and we can use the chain
∂x du ∂x
rule to write
∂ f df ∂ u df
= = (37) ☞
∂ x du {∂x du
This is
equal
to 1

∂f df
Given that f0(x, t) = sin(x + vt), find , then put u = x + vt and calculate v .
∂t du
∂f df
Hence show that =v in this case.
∂t du

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
∂u
Once again, this result is no accident since if u = x ±vt then = ± v , and the chain rule gives
∂t
∂ f df ∂ u df
= = ±v (38)
∂ t du {
∂t du
This is
equal
to ± v

df
We can now use a similar process to find the second-order partial derivatives. Since can also be regarded as
du
a function of x and t, we can differentiate once again with respect to x to obtain
∂2 f ∂  ∂f  ∂  df  d  df  d2 f
2 =   =
 ∂ x  ∂ x  du 
=
 
= (39)
∂x ∂4
1 x2 43 123 142du
du 43 du 2
This is just From This comes from
the definition Eqn 37 Eqn 37 with df du
of the second in place of f
derivative

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Similar operations ☞ can be performed with respect to t, so that
∂2 f ∂  d 2 f ∂u ∂2 f
±v  = ±v 2
df d2 f
= = ( ± v )2 = v2 (40)
∂t 2 ∂t  du  du ∂ t ∂u 2 du 2
d2 f
We can now eliminate from Equation 39
du 2
∂2 f ∂  ∂f  ∂  df  d  df  d2 f
2 =   = = = (Eqn 39)
∂x ∂4
1
 x3 ∂ x 12
x 2∂4  du3 1
 
42du
du 43 du 2
This is just From This comes from
the definition Eqn 37 Eqn 37 with df du
of the second in place of f
derivative

and Equation 40 to obtain

∂2 f 1 ∂2 f
− 2 = 0 (41)
∂x 2 v ∂t 2

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
This is a second-order partial differential equation with the required solution, and is known as the
(one-dimensional) wave equation. Many interesting physical quantities can be shown to satisfy this equation,
and very often we would like to find its solution subject to various boundary conditions. From our previous
discussion we know that any function of the form y( x, t ) = f ( x ± vt ) is a solution of Equation 41,
∂2 f 1 ∂2 f
− 2 = 0 (Eqn 41)
∂x 2 v ∂t 2
but there are two points to appreciate here. First, this may not be the most general solution; and second, we
usually need to find some particular solution in order to solve a given problem.
So there are two important questions to consider: what is the most general solution of Equation 41, and how can
one select a solution that solves a given problem from this vast array of possible solutions? We shall not be able
fully to answer these questions in this module, but we shall show that certain specific functions are indeed
solutions of Equation 41. ☞ We shall also discuss some general features of solutions to the wave equation.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question T10
Show that if two functions f1 and f2 are solutions of the wave equation
∂2 f 1 ∂2 f
− 2 = 0
∂x 2 v ∂t 2
then any function of the form α0f1 + β0f2 , where α and β are arbitrary constants, is also a solution.4❏

Partial differential equations are very common throughout physics and there is a large literature devoted to
techniques which in some circumstances lead to explicit solutions. ☞ No matter what method is used to find a
solution, it can always be verified by substituting back into the differential equation. We shall be concerned with
showing that we really do have a solution to the wave equation, rather than with general techniques for finding
solutions.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
✦ 1 1
Verify that y(x, t) = A sin (ω0t − kx + φ) is a solution of a wave equation of the form
∂2 f 1 ∂2 f
− = 0 (Eqn 41)
∂x 2 v2 ∂ t 2
provided we identify f with y, and providedv2 = ω02 /k2.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
2.8 Deriving the wave equation
1 1
We have shown that the function y(x, t) = A sin (ω0t − kx + φ) is a solution of the wave equation (Equation 41),
∂2 f 1 ∂2 f
− = 0 (Eqn 41)
∂x 2 v2 ∂ t 2
but why should there be waves on the surface of a lake, on a string or sound waves in air? How do we know that
such physical systems are governed, at least approximately, by the wave equation?
As an example of how waves occur in physical systems, we now derive the wave equation for a stretched string.
Other physical systems, such as sound waves in air, can be analysed in a similar way.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
F T sin(θ + ∆θ)
1

We start by considering
a segment AB of the FT
string which is of
length ∆l, as shown in
B θ + ∆θ
Figure 14. The string is ∆l B
assumed to be stretched y A
tightly along the
∆l
horizontal x-direction, l
and then set in motion
in the y -direction x x
perpendicular to its x +∆x θ
A
length. (We ignore the x + ∆1 x
effects of gravity in the (a)
following discussion.)

FT
Figure 144
F T sin θ
1 1
(a) A vibrating string, and 0
(b) an enlargement of the (b)
segment AB.
0

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
The angle θ is the angle between the tangent to the curve and the x-direction, and it will be small if the
oscillations in the string are small. ☞
The following quantities are needed in our derivation:
x = position of one end of the segment as measured from a fixed point along the line of the string when at
rest;
t = time at which the segment is in the position shown;
y(x, t) = displacement from the rest position of the segment ΑΒ;
FT = the magnitude of the force acting on the ends of the string due to tension in the string
(assumed to be constant);
µ = the mass per unit length of the string (the linear mass density);
θ = the angle associated with the point at the ‘lower’ end of the segment;
θ + ∆θ = the angle associated with the point at the ‘upper’ end of the segment.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
F T sin(θ + ∆θ)
1

The mass of the


segment is µ ∆l which FT
is approximately µ1 ∆x if
θ is small (the size of θ B
has been exaggerated in ∆l B θ + ∆θ
A
the figure). Resolving y
the forces acting on the ∆l
segment, as in Figure l
14b, shows that there is
a net transverse force x x
x +∆x θ
which causes the A
x + ∆1 x
segment to accelerate in
the y-direction. (a)

FT
Figure 144
(a) A vibrating string, and F T sin θ
1 1

0
(b) an enlargement of the
(b)
segment AB.
0

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
In other words, we treat x as a constant and differentiate y twice partially with respect to t to obtain the
∂ 2y
acceleration of the segment.
∂t 2
Newton’s second law enables us to relate the acceleration to the force as follows
∂ 2y
µ ∆x = FT sin (θ + ∆θ ) − FT sin (θ ) (42)
{
mass
∂t 2 14444 4244444 3
{ y-component of the force acting on AB
of AB acceleration
of AB in
the y-direction

However for small angles we have sin( θ ) ≈ tan( θ ) so that Equation 42 can also be written
∂ 2y
µ ∆x 2 ≈ FT tan ( θ + ∆θ ) − FT tan ( θ ) (43)
∂t

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
F T sin(θ + ∆θ)
1

Suppose now that we


keep t fixed, and FT
consider the function
f0(x) defined by
B θ + ∆θ
f0(x) = y(x, t) ∆l B
A
for constant t y
the graph of which is ∆l
l
shown in Figure 14a.
The gradient of this
graph at the point A is x x
1
f0′0(x) = tan θ
x +∆x A
θ
x + ∆1 x
while the gradient of (a)
the graph at the point B
is f ′ ( x + ∆x )
1
= tan (θ + ∆θ) Figure 144
FT

(a) A vibrating string, and F T sin θ


1 1

(b) an enlargement of the 0


segment AB. (b)
0

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
This means that we can rewrite the right-hand side of Equation 43
∂ 2y
µ ∆x 2 ≈ FT tan ( θ + ∆θ ) − FT tan ( θ ) (Eqn 43)
∂t
to obtain
∂ 2y
µ ∆x ≈ FT f ′ ( x + ∆x ) − FT f ′ ( x ) (44)
∂t 2
µ ∂ 2y f ′ ( x + ∆x ) − f ′ ( x )
so that ≈
FT ∂ t 2 ∆x
If we now take the limit as ∆x tends to zero, the approximation becomes increasingly accurate, and therefore
(see Question R3) ☞
µ ∂ 2y
= f ′′ ( x ) (45)
FT ∂ t 2

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
∂ 2y
However, we defined f0(x) to be equal to y(x, t) for constant t, so that f ′′ ( x ) = , and this means that
∂x 2
Equation 45
µ ∂ 2y
= f ′′ ( x ) (Eqn 45)
FT ∂ t 2
becomes
µ ∂ 2y ∂ 2y
= (46)
FT ∂ t 2 ∂x 2
Now if we identify µ0/ F T with 1/v2 we see that Equation 46 is actually a form of the wave equation.
This shows that small amplitude waves on a stretched string propagate with a speed v given by
FT
v= (47) ☞
µ

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question T11
The y-component of an electric field Ey(x, t) and the z-component of a magnetic field Bz(x, t) are related by the
partial differential equations ☞
∂ Bz ∂ Ey
= − (48)
∂t ∂x
∂ Ey ∂B
µ 0ε 0 = − z (49)
∂t ∂x
where µ0 and ε0 are constants. Show that Ey and Bz both satisfy wave equations, and in each case find the speed
of the wave. (Hint: Differentiate one equation with respect to t and the other respect to x.)4❏
Study comment

It turns out that any pair of physical quantities, f(x, t) and g(x, t), which satisfy the pair of differential equations
∂f
∂t

∂g
∂x
3 3
and
∂g
∂t

∂f
∂x
where α and β are constants, lead to wave equations for f and g.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Energy flow for a wave on a string
If a wave propagates along a string there is an instantaneous flow of energy past a fixed point on the string. ☞
As a pulse moves past a particular point there is clearly energy present at that point (since the string is moving)
but, once the wave has moved on, the energy returns to zero (since the string is locally at rest). We can see this
mathematically by observing that the instantaneous rate of flow of energy past a point x on the string, i.e. the
power P, is equal to the product of the upward force on the point multiplied by the transverse component of the
∂y
string’s velocity .
(The component of ∂ t velocity of the segment along the string is assumed to be negligibly small.)

✦ Would it be more sensible to write ‘speed’ rather than ‘the transverse component of velocity’?

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
From our earlier discussion we know that the component of force in the y-direction at the point A is
∂y
− FT sin θ ≈ − FT tan θ = − FT (50)
∂x
and therefore the power is given by
∂y ∂y
P = − FT (51)
∂x ∂t
If we consider a wave moving from left to right, then y(x, t) = f(x − vt) and, putting u = x − vt, we obtain
(from Equations 37 and 38)
∂ f df ∂ u df ∂ f df ∂ u df
= = (Eqn 37) = = ±v (Eqn 38)
∂ x du { ∂x du ∂ t du { ∂t du
This is This is
equal equal
to 1 to ± v

∂y df ∂ u df ∂y df ∂ u df
= = −v and = =
∂t du ∂ t du ∂x du ∂ x du

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Hence, from Equation 51,
∂y ∂y
P = − FT (Eqn 51)
∂x ∂t
the power is given by
2
P = vFT  
df
(52)
 du 
The sign of P shows that the energy flows from left to right for a wave that travels from left to right.

✦ Show that P is negative for a wave moving from right to left.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
2.9 Standing waves
So far we have visualized waves as fixed shapes which change position,
and we have referred to such waves as travelling waves. However, not all
waves are travelling waves. Figure 15 shows three of the ways in which a (a)
stretched string with fixed endpoints can be made to move. In each case
the displacement of the string from its equilibrium position is a function
of position and time, and will satisfy the wave equation. Yet in this case
the ‘waves’ are not going anywhere; they do not appear to travel to the
left or to the right. In fact, they are called standing waves. On a standing (b)
wave there are some positions, called nodes, at which the displacement is
always zero.

Figure 15 Standing waves for a string fixed at its ends. This is the wave profile,
and the horizontal axis is distance. (c)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
As we will now demonstrate, a standing wave can be regarded as a
particular combination of progressive waves which are travelling in
opposite directions. ☞ Consider the sum of two progressive waves which
have the same angular frequencies and wavenumbers but are moving in (a)
opposite directions
1 1 1 1
y(x, t) = C sin (ω0t − kx + φ1 ) + D sin (ω0t + kx + φ2) (53)
and suppose we are looking for standing waves in the stretched string
shown in Figure 15. Let the string be of length L. The parameters in the (b)
expression for y(x, t) are not completely arbitrary since y(x, t) must satisfy
the condition that neither end of the string can move.
If x = 0 corresponds to one end of the string, then this implies that
y(0, t) = 0 for all values of t (54)
and therefore
1 1 1 1
C sin (ω0t + φ1) + D sin (ω0t + φ2) = 0 for all values of t (c)

Figure 15 Standing waves for a string fixed at its ends. This is the wave profile,
and the horizontal axis is distance.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
1 1 1 1
Since this equation is true for all values of t the functions C sin (ω0t + φ1) and −D sin (ω0t + φ2) must be identical,
and this implies that
C = − D4and4φ1 = φ2.
If we replace φ1 and φ 2 by φ, and D by −C our solution (Equation 53) to the wave equation
1 1 1 1
y(x, t) = C sin (ω0t − kx + φ1 ) + D sin (ω0t + kx + φ2) (Eqn 53)
reduces to
1 1 1
y(x, t) = C [sin (ω0t − kx + φ) − sin (ω0t + kx + φ)]
This expression can be converted into a more convenient form by using the trigonometric identity

1 1 α −β
sin α − sin β = 2 sin 
 2 
α +β
 cos 
 2 
 (55)

1 1 1 1
to give y(x, t) = −2C cos (ω0t + φ) sin (kx) (56)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
The point (0, L) corresponds to the other end of the string, which is also fixed; so y(L, t) = 0 for all values of t.
1 1 1
So, cos (ω0t + φ) sin (kL) = 0 (57)
Since this is also true for all values of t, we must have
1
sin (kL) = 0 (58)
The sine function is zero only when its argument is an integer multiple of π, so it follows that
kL = nπ4where n = 0, ±1, ±2, … (59)
Equation 59 is of profound importance because it tells us that the value of k cannot be chosen arbitrarily; only
certain values of k are possible if this end of the string is to remain fixed. Substituting the allowed values of
k = nπ/L in Equation 56 we find
1 1 1 1
y(x, t) = −2C cos (ω0t + φ) sin (kx) (Eqn 56)
all the allowed forms of sinusoidal standing waves that can exist on a string of length L with fixed end points

1 1
y(x, t) = A cos (ω0t + φ) sin  1
nπx 
 L 
4where n = 1, 2, 3,… (60)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Notice that in writing Equation 60

1 1
y(x, t) = A cos (ω0t + φ) sin  1
nπx 
 L 
4where n = 1, 2, 3,… (Eqn 60)

we have replaced −2C by A for convenience and we have dropped the non-positive values of n. The negative
values of n do not lead to expressions for y(x, t) which are independent of the positive values since

sin  −
nπx 
= − sin 
nπx 


L   L 
and all these waves can be reproduced by selecting an appropriate value for the phase constant φ .
Also note that the ‘wave’ corresponding to n = 0 in Equation 60 is the trivial case of the string at rest and is
therefore of little interest.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1

(a) What values of n in the function

1 1
y(x, t) = A cos (ω0t + φ) sin 1 1
nπx 
 L 
(a)

correspond to the three standing waves of Figure 15?

(b) What expressions for y(x, t) correspond to these standing waves in


(b)
Figure 15?

(c)

Figure 15 Standing waves for a string fixed at its ends. This is the wave profile,
and the horizontal axis is distance.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
The values for φ in the above function are indeterminate, since the wave profile tells us nothing about time.
Although each expression for y(x, t) given by a particular value of n is a solution of the wave equation
∂ 2y 1 ∂ 2y
− 2 =0
∂x 2 v ∂t 2
these waves do not appear to propagate with speed v since they are in fact standing waves. As we have just
demonstrated, each of these standing waves is actually the sum of two progressive waves travelling in opposite
directions with speed v. The standing waves cannot be expressed as a single function of f0(x ± vt), since they are
always combinations of two such functions with opposite signs within the bracket.

✦ Are the frequencies of the standing waves that we have found for the stretched string arbitrary?

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Study comment Make sure you study the answer to the above question. The answer is important in its own right, but it
also leads on to the important finding that:

The standing waves can be written in the form

y ( x, t ) = A cos 
nπvt
 L
+ φ  sin 

nπx 
 L 
where n = 1, 2, 3… 3(Eqn 61)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question T12
If the standing waves on a string are described by the function y(x, t) given in Equation 61,

y ( x, t ) = A cos 
nπvt
 L
+ φ  sin 

nπx 
 L 
where n = 1, 2, 3… 3 (Eqn 61)

show that the rate of energy flow past a fixed point x on the string is given by
2
P = FT v 
nπA 
sin  + 2 φ  sin 
2 nπvt 2 nπx 
 2L   L   L 
Also find the values of P at x = 0 and x = L and explain your results.

1 1 1 1 1 3
Hints: The previous section gives an expression for P in terms of partial derivatives, and you may need to use
the identity sin (2θ ) = 2 sin (θ) cos (θ ). ❏

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
3 The Schrödinger equation
Study comment The Schrödinger equation is the central equation of quantum mechanics. The purpose of this section is to

1 1
introduce the Schrödinger equation as another example of a partial differential equation with wave-like solutions, and to
examine some of its mathematical properties — it is not designed to teach you quantum mechanics. A very brief survey of
quantum mechanics is contained in Subsection 3.1, to enable you see something of the physical context in which the
Schrödinger equation arises, but for a comprehensive introduction to this very important part of physics you should consult
the appropriate blocks in the physics strand of FLAP.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
3.1 The time-dependent Schrödinger equation
1 1
In the early years of the 20th century it became increasingly apparent that classical physics — the mechanics of
1 1
Newton, the electromagnetism of Maxwell, the thermodynamics of Clausius, and so on — was incapable of
accounting for many of the phenomena involving atoms and nuclei that were being discovered at that time.
Because of this the period from 1900 to 1925 saw the development of a number of poorly formulated ‘theories’
that tried to explain the new phenomena in terms of an ad hoc mix of classical ideas and radical new proposals.
The need for a new theoretical framework to replace classical physics, one that would bring old and new ideas
together in a more coherent way, was widely perceived, but the experimental data were so fragmentary, and the
required shift in thinking so great that it was not until 1925/26 that the way forward became clear. It was during
these years that a relatively small number of European physicists including Werner Heisenberg (1901–1976),
Erwin Schrödinger (1887–1961) and Max Born (1882–1970) began to develop the new quantum physics, that
was destined to displace classical physics from its position as the most comprehensive and coherent description
of physical reality. Quantum physics is now a wide-ranging discipline with many subdivisions that influence
almost every other branch of physics. However, the first part of quantum physics to be developed systematically
was quantum mechanics, the quantum physical counterpart to classical mechanics, and it is this part of
quantum physics that we shall consider in this section. ☞

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Quantum mechanics addresses questions such as ‘how do the electrons in atoms behave?’ and hence, ‘how much
energy will the electrons in a given atom absorb or emit under given circumstances?’ These are questions that a
19th century scientist would have tried to answer using classical mechanics and classical electromagnetism, but
by 1925 it was clear that when classical physics was applied to individual atoms it was often inadequate,
inconsistent or just plain wrong. The ability of quantum mechanics to account for atomic phenomena caused a
revolution in physics, not merely because quantum mechanics agreed with experiment where classical physics
failed, but also because the answers it provided were often of a profoundly different kind and were arrived at in a
profoundly different way from those of classical physics. To clarify this difference let us examine the ways in
which classical and quantum physicists might have tried to formulate a simple one-dimensional model of an
atom, in which a negatively charged electron is bound to a positively charged nucleus by an electrical force.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
The classical physicist would have started with a mathematical formula for the electrical force F x on the
electron, or equivalently, a formula relating the electrical potential energy U(x) of the electron to its position x.
☞ The classical physicist would then use that force (or the related potential energy) to write down Newton’s
second law of motion as a differential equation. In a simple one-dimensional case this might have the general
form
d2x
m 2 = Fx
dt
or, in terms of the potential energy,
d2x − dU
m 2 =
dt dx
where m is the mass of the electron and x its position at time t.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Given an explicit expression for F x or U(x), the classical physicist could then solve the relevant differential
equation, subject to various boundary conditions, to find the position of the electron as a function of time x(t).
Equipped with a solution of this kind it would be relatively easy to determine all the other attributes of the
electron such as its velocity (vx = dx/dt), its momentum (px = mvx) and its energy ((mvx2/2) + U(x)), at any time.
Of course, having done all this, the classical physicist would probably find that the final results bore little
relation to the measured properties of electrons even if they were confined to one dimension.
Classical mechanics works well for macroscopic (laboratory scale) objects but often fails miserably in the
microscopic (atomic scale) domain.
For the quantum physicist the approach is very different. Force is a concept that sits rather uneasily in quantum
mechanics, so the quantum physicist would certainly start from the relevant potential energy function. Taking
into account the fact that the potential might change with time as well as position, the quantum physicist would
then write down the time-dependent Schrödinger equation. This is as fundamental to quantum mechanics as
Newton’s second law is to classical mechanics. In a simple one-dimensional case it would take the form

˙2 ∂ 2 Ψ ( x, t ) ∂Ψ ( x, t )
− + U ( x, t ) Ψ ( x, t ) = i˙ (62)
2m ∂x 2 ∂t

where the symbols have the following meaning.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
h=

h
11
= 1.055 × 10−34 J s is a physical constant (pronounced ‘h-bar’ and referred to as “Planck’s constant
divided by 2π”).
m represents the mass of the electron, as before.
U(x, t) is the potential energy function particular to the physical system being considered. If we wanted to
consider a different kind of atom then the function U(x, t) would have to be changed to represent the new
problem. (This would have been equally true in classical physics.)
Ψ(x, t), the subject of the equation, is a function of two variables (in this one-dimensional case) and is called
the wave function. ☞ It is the role of the wave function that really distinguishes quantum physics from
classical physics, so we shall have a lot more to say about it shortly. (In what follows we shall always
assume that the wave functions that satisfy the Schrödinger equation for a given potential energy function
U(x, t), satisfy an additional technical requirement known as normalization.)
i is the imaginary constant defined by i2 = −1 and often referred to as ‘the square root of −1’. Its presence is
another of the features that distinguishes quantum mechanics from classical mechanics.
x and t are position and time coordinates, as before, but it is no longer true to say that x is the position of the
electron at time t.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Aside Notice that Equation 62 differs from our one-dimensional wave equation (Eqn 41) in two fundamental ways. Firstly, it
involves the first not the second time derivative of the function. Secondly, it is a complex equation and its solutions are in
general complex functions. In a technical sense then, the Schrödinger equation is not strictly a wave equation, but it does
have wave-like solutions, and for this reason is often referred to as a wave equation.
˙2 ∂ 2 Ψ ( x, t ) ∂Ψ ( x, t )
− + U ( x, t ) Ψ ( x, t ) = i˙ (Eqn 62)
2m ∂x 2 ∂t
∂2 f 1 ∂2 f
− 2 = 0 (Eqn 41)
∂x 2 v ∂t 2

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Knowing the explicit form of U(x, t) the quantum physicist can solve the time-dependent Schrödinger equation,
subject to appropriate boundary conditions, to find a wave function Ψ(x, t). ☞ However, unlike the solution to
Newton’s second law x(t), a quantum mechanical wave function Ψ(x, t) will involve i (i.e. it will involve
complex numbers) and will not be a physically measurable quantity. How then is the quantum physicist to
determine the precise location of the particle at any time? The short answer, according to quantum mechanics, is
that the quantum physicist cannot do so, because the electron may not have a precisely defined location at every
possible time. This odd-sounding limitation is sometimes described by saying that ‘quantum mechanics is a
theory of behaviour, not a theory of attributes’. This means that rather than providing information about some
presumed attribute of the electron, such as its position at all possible times, what quantum mechanics provides is
information about the observable behaviour of the electron, such as the likelihood that it will be found in a
certain region of space at a particular time. According to quantum mechanics, this behavioural information is
contained in the wave function.

For instance, the probability ☞ that at some particular time t = T, the electron will be found in some small
1 1 1 1
region of fixed length ∆X centred on a given point x = X at time t = T, is | Ψ0(X, T) |02 ∆X, where | Ψ0(X, T) |
represents the modulus of the complex quantity Ψ0(X, T). Of course, there is nothing special about the particular
1
11
values X, T and ∆X that we have chosen, so we can say that in general the probability of finding the electron in a
1
small range ∆x, centred on x at time t is | Ψ0(x, t) |2 ∆x.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Unlike the classical trajectory x(t), this information about the probability of finding the electron within a certain
range of x values at a certain time is of little help in determining other important attributes such as energy and
momentum. However, in keeping with its nature as a theory of behaviour rather than attributes, quantum
mechanics does provide the answer to a behavioural question such as; ‘what are the possible outcomes of an
experiment to measure the energy of the electron?’ Perhaps surprisingly, it is a common feature of quantum
mechanics that confined particles, such as electrons contained within atoms, are only allowed to have certain
discrete energy values. These allowed energy values may be determined by examining all possible solutions to a
given time-dependent Schrödinger equation. A particular wave function satisfying that equation will then
determine the relative likelihood of each of those allowed energy values turning up as the outcome of an
experiment to measure the energy. Similar comments may be made about other observable quantities such as
momentum or position.
Thus a particular wave function may be said to describe a particular quantum state of the system being studied.
To a classical physicist a ‘state’ of the system would be defined by the values of attributes such as energy,
momentum, position etc., but to a quantum physicist the state is synonymous with a given wave function that
determines a particular set of probabilities of particular experimental outcomes. Even when the system is in a
particular state, attributes such as energy and momentum need not have definite values, only possible values
with definite probabilities of being observed under specified conditions.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Of course, you are not meant to understand quantum theory from this brief discussion. Rather, the points that
you should take away from this subsection are the following:
o Quantum mechanics involves solving a partial differential equation called the time-dependent Schrödinger
equation.
o The precise form of this equation differs from problem to problem, since each problem is characterized by a
potential energy function U that appears in the time-dependent Schrödinger equation.
o The solution to the time-dependent Schrödinger equation (subject to specified boundary conditions) is called
the wave function. In a one-dimensional problem it will be a function of two variables and may be written
Ψ(x, t). This is generally a complex quantity that is not directly measurable in an experiment.
o The answers to behavioural questions about a system, in so far as such answers exist at all, are obtained by
carrying out various mathematical operations that involve the wave function. The wave function may
therefore be regarded as a description of the state of the system under the given boundary conditions.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
3.2 The time-independent Schrödinger equation
Many quantum mechanical problems involve potential energy functions that do not depend on time. Under these
circumstances the one-dimensional time-dependent Schrödinger equation takes the form
˙ 2 ∂ 2 Ψ ( x, t ) ∂Ψ ( x, t )
− + U ( x ) Ψ ( x, t ) = i˙ (63)
2m ∂x 2 {
Not a
∂t
function
of t

It can be shown that because the potential is independent of t in this case, Equation 63 may have separable
solutions, that is solutions Ψ(x, t) that can be written as a product of a function of x and a function of t. ☞
When dealing with a separable wave function Ψ(x, t), the following notation is used to indicate its separability

Ψ(x, t) = ψ(x)φ(t) (64)


The function ψ(x) on the right-hand side of Equation 64 is then called the spatial part of the wave function
(though this is often contracted to spatial wave function), and the function φ(t) is referred to as the
temporal part of the wave function, (or just the temporal wave function). Note the distinction between the
upper case Greek psi (Ψ) used to denote the full time-dependent wave function and the lower case psi ( ψ) used
for its spatial part.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Using Equation 64

Ψ(x, t) = ψ(x)φ(t) (Eqn 64)


we may rewrite Equation 63
˙ 2 ∂ 2 Ψ ( x, t ) ∂Ψ ( x, t )
− + U ( x ) Ψ ( x, t ) = i˙ (Eqn 63)
2m ∂x 2 {
Not a
∂t
function
of t

in the following form:


˙2 ∂ 2 ∂ ( ψ ( x ) φ (t ) )

2 m ∂x 2 [ ψ ( x ) φ ( t ) ] + U ( x )[ ψ ( x ) φ ( t ) ] = i˙
∂t
Since t (and hence φ(t)) behaves as a constant when we differentiate partially with respect to x, and since x
(and hence ψ(x)) behaves as a constant when we differentiate partially with respect to t, we can rewrite this as
follows
 ˙2 ∂ 2 ψ ( x )  ∂φ
φ (t )− + U ( x ) ψ ( x )  = i˙ψ ( x )
 2 m ∂x 2
 ∂t

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
However ψ(x) is a function of the single variable x, so ∂ 2 ψ ∂ x 2 = d 2 ψ d x 2 and φ(t) is a function of the
single variable t, so ∂φ ∂ t = dφ dt .
 ˙2 d 2 ψ ( x )  dφ
Hence φ (t )− + U ( x ) ψ ( x )  = i˙ψ ( x ) (65)
 2 m dx 2
 dt
Notice that by restricting our attention to the separable solutions that can arise when the potential energy is
independent of time we have been able to replace the partial derivatives of Equation 63
˙ 2 ∂ 2 Ψ ( x, t ) ∂Ψ ( x, t )
− + U ( x ) Ψ ( x, t ) = i˙ (Eqn 63)
2m ∂x 2 {
Not a
∂t
function
of t

by ordinary derivatives.
Dividing both sides of Equation 65 by ψ0(x)φ0(t) we get the following
φ ( t )  ˙2 d 2 ψ ( x )  i˙ψ ( x ) dφ ( t )
− + U ( x )ψ ( x ) =
ψ ( x ) φ ( t )  2 m dx 2  ψ ( x ) φ ( t ) dt

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
1  ˙2 d 2ψ ( x)  i˙ dφ (t )
i.e. − + U ( x )ψ ( x ) = (66)
ψ ( x )  2m dx 2 φ (t ) dt
144444424444443 14243
This is a function of x only This is a function
of t only

This is a clever step because, as indicated, it produces an ordinary differential equation in which the only
variable appearing on the left-hand side is x, while the only variable on the right-hand side is t. Think about that
for a moment. How is it possible for a function of x (on the left of Equation 66) to be equal to a function of t
(on the right of Equation 66) when x and t are independent variables? There is only one way that this can
happen; and that is for each side of Equation 66 to be constant. A constant introduced in this way is called a
separation constant; in this case we shall denote it by C. We can then write Equation 66 as two ordinary
differential equations

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
1  ˙2 d 2 ψ ( x ) 
− + U ( x )ψ ( x ) = C (67)
ψ ( x )  2 m dx 2 
i˙ dφ ( t )
and = C (68)
φ ( t ) dt
It is easy to verify that the second equation, Equation 68, has the solution

φ(t) = Κ exp(−i0C0t/˙) (69) ☞


for some constant K.
Moreover, it can be shown that the constant C is of considerable physical significance; it represents the total
energy E, of the particle.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
So a solution to Equation 63
˙ 2 ∂ 2 Ψ ( x, t ) ∂Ψ ( x, t )
− + U ( x ) Ψ ( x, t ) = i˙ (Eqn 63)
2m ∂x 2 {
Not a
∂t
function
of t

that describes a state in which the particle has energy E is given by

Ψ(x, t) = ψ(x)φ(t)
where 1
φ(t) = K exp(−iEt/˙) (70)

˙2 d 2 ψ ( x )
and − + U ( x ) ψ ( x ) = Eψ ( x ) (71)
2 m dx 2

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
where Equations 70 and 71 have been obtained from Equations 69 and 67
1
φ(t) = K exp(−iEt/˙) (Eqn 70)
˙2 d 2ψ (x)
− + U ( x ) ψ ( x ) = Eψ ( x ) (Eqn 71)
2m dx 2
φ(t) = Κ exp(−i0C0t/˙) (Eqn 69)

1  ˙2 d 2 ψ ( x ) 
− + U ( x )ψ ( x ) = C (Eqn 67)
ψ ( x )  2 m dx 2 
by replacing the separation constant C by the energy E.
Clearly, if we want to find the explicit form of the solution Ψ(x, t), we must now solve Equation 71 to find ψ(x),
just as we solved Equation 68
i˙ dφ ( t )
= C (Eqn 68)
φ ( t ) dt
to find φ (t).

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Equation 71
˙2 d 2 ψ ( x )
− + U ( x ) ψ ( x ) = Eψ ( x ) (Eqn 71)
2 m dx 2
appears elsewhere in FLAP and is known as the time-independent Schrödinger equation. Finding and
interpreting solutions to this equation for various explicit forms of U(x) is often a major preoccupation in
introductory courses on quantum mechanics. We shall not pursue that subject here except to note the following
points:
o For a given potential energy function U(x) it is often possible to find a number of different solutions to the
one-dimensional time-independent Schrödinger equation, each of which corresponds to a different value of
the energy E. These different solutions may be conveniently labelled ψ1 (x), ψ2(x), ψ3(x), ψ4(x), … and the
corresponding values of E may be labelled E1, E2 , E3 , E4 , …
o For many choices of U(x) it will be the case that E1, E2 , E3 , E4 , … are separate and distinct numbers, i.e. the
possible values of the total energy are discrete. In this situation we say that the energy of the system is

1 1
quantized. However, it should be noted that quantum mechanics does not automatically demand the
discrete quantization of energy in all systems — it is perfectly possible for the time-independent
Schrödinger equation to have solutions that correspond to a continuous range of values for E, if U(x) has the
appropriate form.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
o The possible existence of many different solutions to Equation 71,
˙2 d 2 ψ ( x )
− + U ( x ) ψ ( x ) = Eψ ( x ) (Eqn 71)
2 m dx 2
with different values of E for a given choice of U(x), implies that there may be many different separable
wave functions Ψ(x, t) that satisfy Equation 63.
˙ 2 ∂ 2 Ψ ( x, t ) ∂Ψ ( x, t )
− + U ( x ) Ψ ( x, t ) = i˙ (Eqn 63)
2m ∂x 2 {
Not a
∂t
function
of t

If so these may be conveniently labelled Ψ1(x, t), Ψ2 (x, t), Ψ3 (x, t), Ψ4 (x, t), …, where
Ψ1 (x, t) = exp( −iE1t ˙) ψ1 (x), 3Ψ (x, t) = exp( −iE t ˙) ψ (x),
2 2 2

Ψ3 (x, t) = exp( −iE3t ˙) ψ3 (x), …

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
o Note that each of these wave functions describes a state in which the system has a fixed amount of energy.
o If Ψi(x, t) and Ψj(x, t) are both solutions to the time-dependent Schrödinger equation for a given potential
energy function, then
Ψ(x, t) = αΨ i(x, t) + βΨj0(x, t)
where α and β are complex constants, will also be a solution. In the case of Equation 63,
˙ 2 ∂ 2 Ψ ( x, t ) ∂Ψ ( x, t )
− + U ( x ) Ψ ( x, t ) = i˙ (Eqn 63)
2m ∂x 2 {
Not a
∂t
function
of t

this means that there may be solutions of the form


Ψ(x, t) = α exp( −iEi t ˙) ψi(x) + β exp( −iE j t ˙) ψj(x)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
o Assuming that i and j represent different numbers, so that Ei ≠ E j, it is clear that this wave function
Ψ(x, t) = α exp( −iEi t ˙) ψi(x) + β exp( −iE j t ˙) ψj(x)
describes a state in which the system has no definite energy. However, according to quantum mechanics, if
the energy of the particle is measured then the result will be either Ei or Ej and the probability of obtaining
either result in a sequence of identical measurements is predictable. In physical terms this is associated with
the system being in a superposition state. If the quantities α and β are themselves time-dependent, this can
often be associated with the system making a transition between the two states.
Aside Many authors refer to ‘the Schrödinger equation’, without specifying whether they mean the time-dependent equation
or the time-independent equation. Generally it is clear from the context which equation is relevant and the explicit labels
‘time-independent Schrödinger equation’ and ‘time-dependent Schrödinger equation’ are only used when there is likely to be
confusion.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
4 Closing items
4.1 Module summary
1 An oscillation occurs when a physical quantity repeatedly cycles about an equilibrium value. The physical
quantity is therefore a periodic function of time. An example is the simple harmonic oscillator which obeys
the ordinary differential equation
d 2 y(t )
= −ω 2 y(t ) (Eqn 3)
dt 2
which, for a given angular frequency ω, has the solution
1 1
y(t) = C sin (ω0t + φ) (Eqn 5)
where C and φ are constants.
2 A wave is a change in a physical quantity which propagates in space. The physical quantity is therefore a
function of both position and time.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
3 A function of two variables x and y, may be written as f0(x, y). The partial derivative of f with respect to x is
written as
 ∂f   ∂f  ∂f
  or   or
 ∂ x  y=constant  ∂ x  y ∂x
which means that f is differentiated with respect to x while holding y constant. Similarly, the partial
 ∂f   ∂f  ∂f
derivative of f with respect to y is written as   or   or
 ∂ y  x = constant  ∂y  x ∂y
which means that f is differentiated with respect to y while holding x constant.
4 To obtain the partial derivatives of a function of two variables, use the rules for differentiating a function of
one variable and treat the other variable as a constant.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
5 A wave which travels in the positive x-direction while maintaining its shape corresponds to a function of the
form y(x − vt), while a wave which travels in the negative x-direction while maintaining its shape
corresponds to a function of the form y(x + vt), where v is the speed of propagation in either case.
Such waves obey the (one-dimensional) wave equation
∂ 2y 1 ∂ 2y
− 2 = 0
∂x 2 v ∂t 2
A particularly important solution is the sinusoidal travelling wave
1 1
y(x, t) = A sin [k(x − vt) + φ] (Eqn 17)
where A, k (= 2π/λ) and φ are constants.
6 A wave equation for a vibrating string can be obtained by considering the forces acting on a short segment
of the string. A partial differential equation for the displacement of the string from its equilibrium position
is obtained in the limit as the length of this segment tends to zero.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
7 Any pair of physical quantities, f0(x, t) and g(x, t), which satisfy the pair of differential equations
∂f
∂t

∂g
∂x
33 33
and
∂g
∂t

∂f
∂x
where α and β are constants, lead to wave equations for f and g.
8 A standing wave is a particular combination of two travelling (or progressive) waves moving in opposite
directions.
9 The time-dependent Schrödinger equation, for a system consisting of a single particle of mass m moving in
one dimension with potential energy U(x, t), is the partial differential equation
˙2 ∂ 2 Ψ ( x, t ) ∂Ψ ( x, t )
− + U ( x, t ) Ψ ( x, t ) = i˙ (Eqn 62)
2m ∂x 2 ∂t
This is the central equation of quantum mechanics. Its solutions Ψ(x, t), subject to appropriate boundary
conditions, are called wave functions and describe the possible states of the system. A wave function is
generally a complex quantity and is not directly measurable, but it yields (probabilistic) information about
the behaviour of the system.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
10 If U is a function of x only so that U(x, t) = U(x), and E is the energy of the particle, the time-dependent
Schrödinger equation may have separable solutions of the form Ψ(x, t) = ψ 0( x) exp(−i0Et0/0˙), where ψ0(x) is
called the spatial part of the wave function and satisfies the time-independent Schrödinger equation
˙2 d 2 ψ ( x )
− + U ( x ) ψ ( x ) = Eψ ( x ) (Eqn 71)
2 m dx 2
For appropriate choices of U(x), this equation may have many different solutions which may correspond to
discrete values of E. When this occurs we say the energy of the system is quantized. This is the crucial
feature of quantum mechanics which is lacking in classical mechanics.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
4.2 Achievements
Having completed this module, you should be able to:
A1 Define the terms that are emboldened and flagged in the margins of this module.
A2 Understand the difference between an oscillation and a wave and give examples of each.
A3 Find the partial derivatives of a function of two variables.
A4 Recognize and derive the wave equation for a disturbance which propagates without changing its shape.
A5 Recognize, use and interpret typical solutions to the wave equation, such as
y ( x, t ) = A sin ( k ( x − vt ) + φ )
A6 Derive the differential equation for a wave on a string and relate the speed of propagation to physical
parameters for the string.
A7 Describe the role of partial derivatives in the time-dependent Schrödinger equation, explain its relationship
to the time-independent Schrödinger equation, and carry out simple operations involving these equations
and their solutions. (Note: you are not expected to understand quantum mechanics on the basis of the brief
introduction provided in this module.)

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Study comment You may now wish to take the Exit test for this module which tests these Achievements.
If you prefer to study the module further before taking this test then return to the Module contents to review some of the
topics.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
4.3 Exit test
Study comment Having completed this module, you should be able to answer the following questions, each of which tests
one or more of the achievements.

Question E1

(A3) 3Given that f ( x, y ) = yx 2


find
∂f
∂x
and
∂f
∂y
.

∂f
Express in terms of a limit, then evaluate this limit to verify your previous answer.
∂y

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question E2

(A1 and A3) 3Use the standard derivatives of functions of one variable to find ∂∂xf and ∂∂yf for
(a) f(x, y) = xy + 4xy − 2x,3(b) g ( x, y ) = e cos ( y )
2 x2 3

, 34(d) w ( x, y ) = log
x x +y 2 2
(c) h ( x, y ) = . e
x +y3 3 x 3

Question E3

3
(A12and11A3) If2 V ( x, y ) =
x2
x
+y 2
∂ 2V ∂ 2V
2show that2 2 +
∂x ∂y2
= 0.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question E4
(A2) 3Explain the difference between an oscillator and a wave, giving an example of each.
Question E5
3
(A3 and A4) Derive the second-order partial differential equation obeyed by any one-dimensional wave which
moves without changing its shape.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question E6
3
(A3 and A5) A fluid which is subject to a small disturbance in the x-direction obeys the equations
∂V 1 ∂P
= − (72)
∂t ρ ∂x
∂P ∂V
and = −K (73)
∂t ∂x
where V(x, t) is the velocity of fluid in the x-direction due to disturbance of fluid, P(x, t) is the excess pressure
due to disturbance of fluid, ρ is the the fluid density (a constant) and K is a constant (known as the bulk
modulus).
Show that both V(x, t) and P(x, t) obey wave equations, and find the propagation speeds in terms of the constants
K and ρ .

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1
Question E7
3
(A3 and A6) Show that the one-dimensional time-dependent Schrödinger equation for a freely moving particle
of mass m has a solution of the form
1
Ψ(x, t) = A exp[i(kx − Et/˙)]
(where A is a constant). Express k in terms of m, E and ˙. Hint: For a free particle U = 0.

Study comment This is the final Exit test question. When you have completed the Exit test go back to Subsection 1.2 and
try the Fast track questions if you have not already done so.

If you have completed both the Fast track questions and the Exit test, then you have finished the module and may leave it
here.

FLAP M6.4 Waves and partial differential equations


COPYRIGHT © 1998 THE OPEN UNIVERSITY S570 V1.1

Você também pode gostar