You are on page 1of 23

Chapter 7

The Schrödinger Equation in One Dimension


Introduction
We have seen that associated with any particle is a matter wave described by the wave
function x, t. How this wave function describes a particle and its behavior is the
subject of quantum mechanics. The wave function contains within it all the
information that can be known about the particle. Our goals will be to (1) discover
how information may be extracted from the wave function and (2) learn how to obtain
this wave function for a given system.
The Schrödinger Equation
The quantum wave function requires a wave equation that describes the motion of
particles that behaves as waves when not looked at but are particles when detected or
observed. It was Schrödinger (1925-26) who first came up with a simple nonrelativistic
wave equation. Heisenberg came up with an alternate method but is more complicated
to physically, so we will not cover it. One cannot derive Schrödinger’s equation just as
one cannot derive Newton’s laws; they are postulated! One can get some “feel” as to
how the Schrödinger equation came about by considering two things:
 The wave function must be a linear combination of harmonic waves in order to
observe interference: Ψ(x, t) = a1ψ1+ a2ψ2. Why? Remember the double slit
experiment with the electrons: Prob(A + B) = A2 + B2 + 2AB. This result is only
possible when it is a superposition of waves.
 Schrödinger’s Postulate
The total energy of a nonrelativistic particle along with the de Broglie relations (E=ℏω
and p=ℏk) leads to
p2 2 2
k
 V E  U 
2m E  2m
p k
Since the wave function is a collection of harmonic functions (Fourier series)
ψ(x,t)   A cos kx  t 
derivatives with respect to position and time lead us to
ψ  2ψ
    total energy  k 2  KE term
t  x 2
p k
E 
Schrödinger’s Postulate: the wave equation for matter waves of mass m is
2
2 
  U(x)  i
2m x 2
PE
t
KE Etotal

Remarks
1. In QM, experiments that are performed are called operations and operators describe
the measured value. For example, the differential operator d/dx measures slope; ∇∙E
measures divergence in the E-field lines. The Schrödinger equation is an operator
equation that operates on the wave function such that
d 
p̂  i 
dx
2
2   p̂2 ˆ  Eˆ  
 
   U(x)  i    U
d 2m x 2
t  2m 
Eˆ  ˆ i
dt 
2. The only solutions to the Schrödinger equation are linear combination of harmonic
wave functions since this guarantees interference between matter waves. If the
quantum wave function was  = A cos (kx−ωt),  is NOT a solution to the
7-1
Schrödinger equation. A more general expression is required that takes on a
complex exponential form:
  x,t   Ae 
i kx t 
 A cos kx  t   i sin kx  t 
e cos x i sin x
ix

This is a solution to the Schrödinger equation. Proof:


2
2 
  U(x)  i
2m x 2
t  Ae 
i kx t 

Proof:
2
   
  
2
Ae 
i kx t 
 U(x) Ae 
i kx t 
Ae 
i kx t 
 i
2m x 2
t
 2
 i kx t 
  i2k 2  U(x)  i ( i) Ae 
 2m 
2 2
k
U 
2m
3. The quantum wave function should immediately strike you as bizarre: it explicitly
contains the imaginary number i = √(−1). Since this is imaginary, it means that it has
no physical reality and therefore, is not a directly measurable function like those
classical wave functions for strings, for example. Analogy: pink hippo standing next
to me. NO one can measure it and so therefore, it does not exist. The wave function
is very similar. When a particle is in a superposition state, the particle does not have
a position or velocity until you observe it. Therefore, the Born interpretation defines
the meaning of the quantum wave function that makes it “reasonable”: the
probability that a particle will be found in the infinitesimal interval dx about the
point x, denoted by P(x) dx, is
P  x,t  dx    x,t  dx   * dx
2

4. The wave function is a computational device with no physical significance. The only
physical significance is |(x,t)|2 = * , which is proportional to the probability
distribution P(x, t).Nomenclature:
  x,t    *   probability density
2
&   probability amplitude
5. Because the quantum particle must be somewhere, the summed probabilities must
add up to 1:

   x,t 
2
dx  1 Normalization Condition
Any wave function satisfying this equation is said to be normalized. Normalization is
simply a statement that the particle can be found somewhere with certainty.
Example 7.1
a. Sketch the plot of the wave function (x) = Ce|x|/x0, where C and x0 are constants.
b. Find the constant C in terms of x0 using the normalization between ±∞.
c. Calculate the probability that the particle will be found in the interval –
x0≤ x ≤ x0.
Solution
a. The given wave function is symmetric, decaying exponentially from the
origin in either direction, as shown.
b. The normalization requirement is
 
1    x,t  dx  C2 
2
e|x|/ x0 dx
 

7-2
Because the integrand is unchanged when x changes sign (it is an even function),
we may evaluate the integral over the whole axis as twice that over the half-axis x >
0, where |x| = x. Then
 x  1
1  2C2  e2x / x0 dx  2C2  0   C2 x 0 
solving for C
 C
0
 2 x0

1

wavefunction
 e|x|/ x0
x0
Units: [] = 1/√L.
c. The probability is the area under the curve of |(x)|2 from –x0 to x0 and is obtained by
integrating the probability density over the specified interval:
2
 x0  1  x/ x 
 x0 2  x0 2x/ x0
P  x dx  2 
x0 0
2
e 0  dx  e dx
 x0 0  x 
 0 


2 x0

x0 2
1  e2  
 P(  x 0  x   x 0 )  1  e2  0.8647  
Probability of 86.5% between  x 0  x  x 0

This example represents very clearly the direction of Chapter 7; once the wave function
has been solved for, one normalizes the wave function so the total probability is 1 and
then one calculates the probability of finding some experimental value (position,
energy…).
Wave Functions in the Presence of Forces
When a force acts on a particle, the potential function represents this force through F =
∂V/∂x. Quantum particles feel a potential (not a forces) the wave function (x, t) is
determined from the Schrödinger equation:
2 2


 U(x)  i
2m x 2
t
To “test drive” this equation, Schrödinger first applied it to blackbody radiation (electrons
acting as harmonic oscillators) and the hydrogen atom (bound states of electrons), with
good results. In both of these test cases, we referred to the state of the electrons as
being in a stationary state (or eigenstate) because the probability does not change
with time. Mathematically, the reason for this is that the potential energy function is
time independent, and this leads to the wave function being separable!
 The potential energy function is time independent: U(x, t) = U(x).
 A separable wave function means the wave function can be broken up into two
different wave functions that depend only on position (x) and time (t)
  x,t   ψ(x)  (t)
This leads to great simplifications in the mathematics. To get a detailed look at this
substitute in these two assumptions into the Schrödinger equation
2
 2 ψ(x)  (t)  ψ(x)  (t)
  U(x) ψ(x)  (t)  i
2m x 2
t
 2ψ(x)
2
i (t)
  U(x)   constant  E
2mψ(x) x 2 (t) t
depends only on position x depends only on time t

where we have collected like terms. The only way that two independent variables can be
equal to each other is if they are equal to the same constant defined as the total energy
E. By recognizing this fact, we have replaced a partial differential equation with two
ordinary differential equations:

7-3
 2
d2ψ(x)
   U(x)  E  time-indep Schrodinger equation
 2mψ(x) dx
2


 i d(t)
 E  time-dependent equation
 (t) dt
The process is to solve both of these equations for both wave functions, then combine
them into the total wave function (x, t). However, there is a nice subtlety that simplifies
the process. Note that we can only solve for the time-dependent equation but not the
time-independent one, since we do not have the form of the potential. Solving for the
time-dependent wave function is straightforward, and is given by
d(t) d(t) iE
i  E(t)    (t)  eiEt /
dt (t)
When this is combined with the time-independent wave function the probability density is
  x,t     x,t  *   x,t   ψ(x)*  (t) * ψ(x)  (t)
2

 ψ(x)*ψ(x)  eiEt / eiEt /  ψ(x)*ψ(x)    x,t 


2

time independent

the time dependent part drops out. That is, the probability density is time
independent and these types of states are called STATIONARY STATES. The
probability density is position depended only and is given by the time-independent wave
function (x). Physically interpretation: for a quantum wave function, the distribution of
matter (of electrons, atoms or nucleons in the nucleus) is time-independent or
stationary. The stationary states are the modern counter part to Bohr’s stationary orbits
and are precisely the states of definite energy. Because their charge distribution is
static, atoms in stationary states do not radiate energy.
The fundamental problem in one-dimensional QM is this: given the potential
function V(x), find the wave function (x) using the time-independent Schrödinger
equation
2
d2ψ(x)
  U(x)ψ(x)  Eψ(x); time-indep Schrodinger equation
2m dx 2
In most cases, it turns out that for many values of E the Schrödinger equation has no
solution (no acceptable solutions, satisfying the particular conditions of the problem, that
is). What limits and makes the solution “physical” is the boundary conditions (BCs)
imposed on ψ(x). In fact, this is exactly the reason why energy gets quantized in QM.
Here are the conditions for an acceptable (or “physical”) wave function:
1. (x) must exist and satisfy the Schrödinger equation.
2. Boundary Conditions:(x) and d(x)/dx must be continuous, finite, and single-
valued. One needs to know the value of  and ′ at point x = x0. This is exactly
analogous to the familiar statement with Newton’s 2nd law; to solve the 2nd law
one has to know the particles position and velocity at time t = t0.
3. (x) → 0 fast enough as x   so the normalization integral remains bounded.

Review of complex numbers


A complex definition of a number is that it must include i = √(−1). A complex number is
defined as
  a  ib and *  a  ib
where a is called the real part and b is the imaginary part. Some properties of complex
numbers are
 The magnitude of a complex number is defined as

7-4
2
   *   (a  ib)(a  ib)  a2  b2
 Euler relationship is
eikx  coskx  isinkx
An application of Euler’s relationship and the magnitude is finding the magnitude of the
complex number eikx.
eikx  coskx  i sinkx; (eikx )  eikx  coskx  i sinkx
| eikx |2  (eikx ) (eikx )  (coskx  i sinkx)(coskx  i sinkx)  cos2 kx  sin2 kx  1
Problem Solving Strategies for the Schrodinger equation
Step 1: Determine the potential function for the quantum particle
Step 2: Setup and solve the Schrodinger equation for (x) using assumed solutions
Step 3: Apply the BCs to get the quantized n and En, and normalize n.
Step 4: Solve the problem and interpret the solutions

Infinite Square Well (Particle in a Box)


Of the problems involving forces, the simplest is that of particle confinement. Classical
analogy: cart with two springs; the cart moves freely in-between the two end points but
cannot escape outside of the end points. An electrical analogy would be an electron
inside a thin conducting wire. To a fair approximation the electron would move freely
back and forth inside the wire but could not escape from it.
Step 1: Determine the potential function for the quantum particle
Consider a particle moving along the x-axis between the points x = [0, L], where L is the
length of the “box.” Inside the box the particles are free and experience no force acting
on it. That is,
dU
F  0  F    U  constant  0  E  K  U  K
dx
However at the endpoints it experiences strong forces that serve to contain it.
F  0  U  0  E  K UU
particle comes to rest
instantly at potential wall
A charged particle moving along the axis between the two sets of electrodes C that
serves to confine a charged particle.

The inner cylinder is grounded, while the outer ends are held at some high electric
potential U. A charge moves freely within the cylinder but encounters electric forces in
the gaps separating them. If V is large, the “walls” become very high and steep,
becoming infinitely high as U →∞. For this setup the potential energy function
looks like
0 0 x L
U(x)  
 x  0 and x  L
Physically, an infinite box requires an infinite amount of energy to remove a
particle from a perfectly “rigid box.” An infinite box seems unrealistic; a real bound
system might require a larger energy to pull apart, but surely not an infinite amount. It
appears that subatomic particles like neutrons and protons are made-up of quarks.
Infinite energy is needed to pull them apart.
7-5
From a classical viewpoint, the particle simply bounces back and forth between the
confining walls of the box. Its speed remains constant, as does its KE. Furthermore,
classical physics places no restrictions on the values of its momentum and energy. The
quantum description is quite different and leads to energy quantization. The way we get
energy quantization is by apply boundary conditions on the quantum particle: since the
potential energy is infinite outside the potential walls, the wave function is
required to be zero there. That is, we expect that
 0 for 0  x  L  the electron acts like a free particle
ψ(x)  
 0 for x  0 and x  L  the electron will never be found there
Classical Analog
A classic example of a boundary condition is that of a vibrating string fixed at both ends
that produce standing waves. If the string is fixed at x = [0, L], we have the same
boundary condition on the vibrating-string wave function y(x) = A sin(kx) such that y(x =
0) = y(x = L) = 0. The key point is the requirement y(x = L) = 0 leads to the resonate
standing-wave condition: an integral number of half wavelengths must fit along
the length of the string.

n   21    L n  1,2,3,...
classical standing wave condition

As we will see, in moving from the classical box to the quantum box, the same condition
follows for a particle in an infinite square potential well.
Step 2: Setup and solve the Schrodinger equation for (x) using assumed solutions
We now compute the wave function (x) for a quantum particle inside the quantum box
where U(x) = 0; we write
 d2 (x) 2mE
d2 (x)  V  0     2 (x)  k 2(x)
2
  U(x)(x)  E(x)    dx 2

2m dx 2  V   
  (x)  0
where
2 2
2mE k
k2  2

solving for energy
 E
2m
The most general solution is a linear combination of harmonic waves:
d2 (x)
2
 k 2(x) 
solution
 (x)  A sinkx  Bcoskx
dx
Step 3: Apply the BCs to get the quantized n and En, and normalize n.
Step 4: Solve the problem and interpret the solutions
This interior wave must match the exterior wave at the walls of the box for it to be continuous
everywhere. Therefore we require the interior wave to vanish at x = 0 and x = L:
ψ(0)  A sin0  Bcos0  B  0 (continunity at x  0)
ψ(L)  A sinkL  0 (continunity at x  L)
Focusing on the boundary condition at x = L,
n
A sink nL  0  sink nL  0  k nL  n 
solving
for k
 kn  ; n  1,2,3,...
L
quantized wave number

7-6
The wave number is quantized as a direct result of the boundary conditions.
Writing this quantized wave number in terms of wavelength leads to the “standing wave
condition:”
n 2 2
k 
solving for
wavelength
    L  n 2  L
L k n
standing wave condition

That is, there are an integral number of half-wavelengths that must fit into the box.
Interpretation: as a quantum particle moves through the potential well, it looks and
smells like a standing wave. These waves are the de Broglie waves of the quantum
particle in the quantum box.
The quantum wave function for a particle in the box is
 n 
n (x)  A sink n x  A sin  x  n  1,2,3,... unnormalized wave function
 L 
Quantization of the de Broglie waves (n) implies the quantization of the momentum and
energy: pn = h/n .Since the energy is E = K + U = K = p2/2m, the allowed energies for a
particle in a one-dimensional box are
pn2 2 2
kn  2 2 
En    n2  2 
 n2E1; n  1,2,3,...
2m 2m kn 
n  2mL 
L

where E1 is the ground state energy and En are sometimes called


the eigenvalues of the energy.
Interpretation of the quantization energy
The energy-level diagram plotted for these energy eigenvalues
are
Remarks
1. The lowest energy for our particle is obtained for n = 1:
 2  
2 2 2
E1  2
    2 
 2  Emin
2mL  2mL 
zero point
energy

This is consistent with the lower bound derived from the


Heisenberg uncertainty principle for a particle confined to a box of length L; that is
the zero point energy:
2
xp  1
2

 Emin 
2mL2
That is, E1 > Emin is larger than the zero-point energy by a factor of 2.
2. Notice that (quite unlike those energy levels of the hydrogen atom) the energy levels
are farther and farther apart as n increases and that En increases without limit as n→
.
3. The spectra lines of the quantum box (i.e. the wavelengths) are given by
hc
Emn  (m2  n2 )E1 
mn
where “m=1” is called the Lyman series, “m=2” the Balmer series… In fact, the

Applications of 1D potential wells


1. Penning Trap
Trapped ions can provide the basis for highly stable, accurate frequency standards. Ions can be trapped for
long periods of time which allows long interrogation times. The device is called a Paul and Penning trap (30
by 60 m) is used in the NIST optical clock to trap a single mercury ion.
http://www.boulder.nist.gov/timefreq/ion/freqstd/hg.htm

7-7
The Noble Prize in Physics 1989 was award to Wolfgang Paul, Hans Dehmelt for trapping a single ion for up
to a year in a Penning trap.
2. Nanostructures and Quantum Wells
There are devices at acts as one-dimensional quantum wells called quantum wires or nanotubes, which
offer the possibility of dramatically increasing the speed at which electrons move through a device in
selected directions. This in turn would increase the speed with which signals move between circuit elements
in computer systems. Nanotubes consist of sheets of hexagonally arranged carbon atoms wrapped around
each other to form a cylinder with a hollow core that are a few nm in diameter, yet up to 1 mm long! As the
image on the left shows, individual atoms are seen as the bumps on the surface of the tubes.

The image on the right is called the “World’s smallest electrical wire.” The carbon nanotube wire (blue) is on
platinum electrodes (yellow). The nanotube is 1.5 nm across, a mere 10 atoms wide. It was made by firing a
laser at a mixture of carbon, nickel and cobalt to make spherical carbon fullerene molecules join together to
form a tube.
http://www.pa.msu.edu/cmp/csc/nanotube.html

Example 7.2 Macroscopic energies


A small bead of mass 1.00 mg is confined to move between two rigid walls separated by
1.00 cm. (a) Calculate the energy of the “Lyman-a state” and its minimum speed of the
object. (b) If the speed of the object is 3.00 cm/s, find the corresponding value of n.
Solution
a. Treating this as a particle in a box, the energy of the particle can only be given by
 2 2 
En  n2  2 
 n2E1; n  1,2,3,...
 2mL 
The minimum energy results (n = 1), given m = 1.00 mg and L = 1.00 cm is
 
2
2
2
h2 6.626  1034 J  s
E1     5.49  1058 J
  
2
2mL2 8mL2 8  106 kg  102 m

E21  (2  1 )  5.49  10 J  16.5  1058 J


2 2 58

There is no way to be able to see the quantization of the bead. When the bead
moves within this box, the energy is all KE and so the minimum speed is

v1 
2E1


2  5.49  1058 J 
 3.31 1026 m/s  v1
m 6
10 kg  
This speed is immeasurably small, so that for practical purposes the object can be
considered to be at rest. Indeed, the time required for an object with this speed to
7-8
move the 1.00 cm separating the walls is about 3 x 1023 s, or about 1 million times
the present age of the Universe! It is reassuring to verify that quantum mechanics
applied to macroscopic objects does not contradict our everyday experiences.
b. If instead the speed of the particle is v = 3.00 cm/s, then its energy is
  
2
E  21 mv 2  21  106kg  3.00  102 m/s  4.5  1010 J
This too must be one of the special values of n. To find which one, we solve for the
quantum number n, obtaining

    
2
 h2  8mL2En 8  106 kg  102 m  4.5  1010 J
En  n 2
2 

solving for n
 n 
 
2
 8mL  h2 6.626  1034 J  s

n  9.05  1023
Notice that the quantum number representing a typical speed for this ordinary-size
object is enormous. In fact, the value of n is so large that we would never be
able to distinguish the quantized nature of the energy levels and behaves
continuously. That is, the difference in energy between two consecutive states
with quantum numbers n1 = 9.05 x 1023 and n2 = 9.05 x 1023 + 1 is only about 10–33
J, much too small to detect experimentally. Therefore, at large n-values, there is a
transition from quantum behavior towards classical physics:

This is another exam


lim QM 
 Classical mechanics
nl arg e

ple that illustrates the working of Bohr’s corresponding principle, which asserts that
quantum predications must agree with classical results for larges lasses and lengths.
Normalization and Probabilities
The actual probabilities can be computed only after (x) is normalized, that is, we must
be sure that all probabilities sum to unity:
 L  nx 
1  ψ(x) dx  An2  sin2 
2
 dx
 0  L 
The integrand is evaluated with the help of the trigonometric identity 2sin2 = 1 – cos 2:
2  nx
 dx  L 2nx  
1  cos    dx  2 L  0    sin2n  sin0  21 L
L L
0 sin 
 L


1

2 0 
  L 
1
4nπ
first term
second term

Putting this all together, we have

 nπx 
 dx  An   2 L   A n 
L 2
1  An2  sin2  2 1 solving for A
0  L  L
Returning to the wave function, we have that

2 n
ψn (x)  sin x n  1,2,3,...
L L

For each value of the quantum number n there is a specific wave function
describing the state of the particle with energy En. The figure below shows plots of n
versus x and |n|2 versus x for n = 1, 2, and 3, corresponding to the three lowest allowed
energies for the particle.

7-9
Seond excited state (n  3)

First excited state (n  2)

Ground state (n  0)

Remarks
1. Note how the number of nodes of the (x) increases steadily with energy; this is what
we should expect, since more nodes means shorter wavelengths and hence larger
momentum and KE.
2. The complete (x, t) for a standing wave has the form

A  i(kx t) 
(x,t)  (x)eit  A sin k nx  e it  e  ei(kx t) 
 
sinkx  eikx  eikx / 2i 2i  right
 left


Our quantum standing wave can be expressed as a sum to two traveling waves (one
moving to the right and one to the left). Exactly like the classical standing wave,
these de Broglie waves are not physical “things;” the only physical thing is ||2.
3. For n = 1, the probability of finding the particle is largest at x = L/2 – this is the most
probable position for a particle in this state. For n = 2, |n|2 is a maximum at x = /4
and again at x = 3L/4: both points are equally likely places for a particle in this state
to be found.
4. There are also points within the box where it is impossible to find the particle. Again
for n = 2, |n|2 = 0 at the midpoint, x = L/2; for n = 3, |n|2 = 0 at x = L/3, 2L/3… This
is weird!
Question: how does our particle get from one place to another when there is no
probability for its ever being at points in between?
Answer: Remember the double-slit experiment: quantum particles have wave
probabilities until you measure them – collapsing the wave function localized them into
particles. Do not ask how it gets from one place to another – you will be going down a
dead alley that no one has ever been able to figure out. This is the quantum enigma.

Example 7.3
A particle is in the ground state of an infinite square well. Calculate the probability that
the particle will be found in the region (a) 0 < x < ½L, (b) 0 < x < ⅓L, and (c) 0 < x < ¾L.
(d) Interpret the results.
Solution
The probability density for the ground state is P(x) = ||2 = (2/L) sin2 (x/L) and its
probability amplitude and density are plotted below:

a. The probability of finding the particle in the range of 0 < x < ½L is


7-10
x
P  0  x  21 L    P(x)dx   41   0 
1
L 2 1
L 2 L 1
 2
L 0  0
sin2 dx   sin2 udu 
2 2 2

0 L L 
P  0  x  21 L   1
2

b. For the range of 0 < x < ⅓L, we get

P  0  x  31 L    61   41 sin( 32 )
2 L 1
 2
 0
 sin2 udu 
3

L 
P  0  x  31 L   0.195

c. For the range of 0 < x < ¾L

P  0  x  34 L    83   41 sin( 32 )
2 L 3
 2
 0
 sin2 udu 
4

L 
P  0  x  34 L   0.909

d. By looking at the probability density curve for the ground state

As the range increases the probability for finding the electron also increases – this is
expected. However, note the probability for finding it between 0 < x < ⅓L is much
smaller than finding it between 0 < x < ½L.
Finite Square Well
We just saw that an infinite box requires an infinite amount of energy to remove a
particle trapped inside. For most systems a more realistic assumption would be that
there is a finite, minimum energy needed to remove a stationary particle from the box. If
we call this minimum energy V0, the potential function for the finite square well is

Even the finite well is unrealistic in that the potential jumps abruptly from 0 to V0 at x = 0
and x = L. For a real particle in a box (electron in a very thin conducting wire) the
potential changes continuously near the walls, more like a rounded well. As one might
expect, the properties of both wells are quantitatively similar.
Application and Interesting Point: Charged-Coupled Devices (CCDs)
Potential wells are essential to the operation of many modern electronic devices, though rarely is the well
shape so simple that it can be accurately modeled by the finite square well just discussed. The charge-
coupled device (CCD) uses potential wells to trap electrons and create a “faithful” electronic reproduction of
light intensity across the active surface.These devices consist of a two-dimensional array of moveable electron
boxes (or wells) created beneath a set of electrodes formed on the surface of a thin silicon chip.

7-11
The silicon serves the dual purpose of emitting an electron when struck by a photon and acting as a local trap for
electrons. The potential energy seen by an electron in this environment is shown by the curve on the right with the
depth coordinate increasing downward. Though far removed from a “box” potential, the well shape nevertheless
serves to confine the emitted electrons in the depth dimension. (Each well or picture element (pixel) in the array
also is isolated electrically from its neighbors, in effect confining the electrons in the remaining two dimensions
perpendicular to the figure.) The number of electrons in a given well, and consequently the number of photons
striking a particular point on the chip, may be read out electronically and the signal processed by computer to
enhance the image. For more than two decades now, CCDs have been helping astronomers see amazing detail
in distant galaxies using much shorter exposure times than with traditional photographic emulsions
If we focus our attention on a particular value of E, we can distinguish two important
regions of x; V E > 0 and V – E < 0. The dividing points between these regions are the
classical turning points (x = 0 and x = L) where E = V(x).

VE positive VE negative VE positive


The region where V – E > 0 is outside these turning points (x < 0 & x > L) and is often
called the classically forbidden region since a classical particle with energy E cannot
penetrate there. Note at the turning points the KE must be zero where the particle comes
instantaneously to rest. The region where V – E < 0 is the interval 0 < x < L and is called
the classically allowed region. We shall see that the behavior of the wave function (x) is
quite different in these two regions. Physically let’s try and see the behavior of the
wave function before actually solving the Schrödinger equation. Whether you are a
research physicist or a professional engineer, do yourself a favor, “see through the
problem before actually solving it.” Talk about the difference as a graduate student
verses a SLAC physicist.
The Schrödinger equation tells us
2
d2ψ(x) d2ψ(x) 2m  V0  E 
  V0ψ(x)  Eψ(x)    ψ(x)
2m dx 2 dx 2 2

This immediately tells us two things:


d2ψ(x) 2m  V0  E  V0  E  0    positive function   ψ(x)

    ψ(x)   
dx 2 V0  E  0    negative function   ψ(x)
2

What does this tells us about the wave function? In the region where V – E is positive,
the Schrödinger equation has the form
V0  E  0     positive function   ψ(x)
In an interval where (x) is positive, this implies that ‫(״‬x) is also positive and
hence that (x) curves upward. In the region where V – E is negative,
V0  E  0     negative function   ψ(x)
In an interval where (x) is negative, this implies that ‫(״‬x) is also positive
and hence that (x) curves downward.
We want to break-up the problem into regions:
7-12
ψ1, (x  0)

ψ(x)  ψ2, (0  x  L)
ψ , (x  L)
 3
A classical particle with energy E less than the well height Vo is permanently bound to
the Region 2. However, if the classical particle has energy E greater than the well height
Vo, it would be outside the well gap and enter the exterior regions (1 and 3), above the
potential well. The particle moves freely but with a reduced speed corresponding to the
diminished kinetic energy given by K = E – Vo. On the other hand, quantum mechanics
asserts that there is some probability that the particle inside the well can be found
outside this region! That is, the wave function generally is nonzero outside the well, and
so the probability of finding the particle is nonzero. This is not possible in a classical
model – these regions are usually called the forbidden regions. We now find the wave
functions for each of the regions using the Schrödinger equation.
Region 2
The solutions inside the potential well (Region II) have already been found earlier:
2
d2ψ2 (x)
 2
 0  E2ψ2 (x)  ψ2 (x)  Aeikx  Beikx
2m dx
2 2
k
E2   KE
2m
Region 1 & 3
The Schrödinger equation for these regions is
2
d2ψ(x) d2ψ(x) 2m  V0  E 
  V 0ψ(x)  Eψ(x)   ψ(x)  2ψ(x)
2m dx 2 dx 2 2

The roots of this equation are complex roots because E1 = E3 > Vo:
2m  V0  E  2m  V0  E  2m E  V0 
2 = 2
   2
 i 2

momentum of the particle is complex (not real)

Substituting this into the general solution:


2mE  V0  2mE  V0 
ii x ii x
ikx
ψ(x)  Ce  De  Ce  De  Cex  Dex
ikx 2 2

k 
where
2
2
V0  E   reduced KE
2m
Summarizing our solutions, we get
ψ1(x)  Cex  De x ; (x  0)

ψ(x)  ψ2 (x)  Ae ikx  Beikx ; (0  x  L)
 x x
ψ3 (x)  Ce  De ; (x  L)
Physically, what do these wave functions mean?
Region 2
This is the standard traveling wave:

Ae
ikx
 particle traveling to the right 2 2
k
ψ2 (x)   ikx where E  KE

Be  particle traveling to the left 2m

Regions 1 and 3
These waves are not traveling waves but are exponential wave forms that are
indefinable as x   :

7-13
 lim Cex  
x  x 
x
lim ψ(x)  Ce  De  
x 
 lim Dex  
 x 
In order to avoid infinite solutions, we require the following two conditions:
Condition 1: The wave function ψ( x) must be finite as x   :
lim ψ(x)  0
x 

This puts conditions on the constants A, B, C, and D.


Condition 2: The boundaries must be continuous. That means that there is a condition
at the boundaries between regions 1 & 2 and 2 & 3 such that
ψ1(0)  ψ2 (0) ψ2 (L)  ψ3 (L)
 
Region 1 & 2   dψ1(0) dψ2 (0) Region 2 & 3   dψ2 (L) dψ3 (L)
   
 dx dx  dx dx
Applying Condition 1 to our wave functions (in the next example, we will apply both
conditions), we get
 At the boundary x = 0, the wave function 1(x) must be finite as x →  . That means
 
lim ψ1(x)  lim Cex  Dex  0  D    D  0
x  x 

 Applying a similar argument at x = L, 3(x) must also be finite as x →  :


 
lim ψ3 (x)  lim Cex  Dex  C    0  C  0
x  x 

Summarizing what we just did, we have


 x a particle moving from the right to the
ψ1(x)  Ce  
 left penetrating the barrier at x  0 2
2

 where E   V0
ψ (x)  Dex  a particle moving from the left to the 2m
 3 
 right penetrating the barrier at x  L
The total wave function is
ψ1(x)  Cex ; (x  0)
 ikx
ψ(x)  ψ2 (x)  Ae  Be ; (0  x  L)
ikx

 x
ψ3 (x)  De ; (x  L)
Now that the dust has settled a bit, what was it that we actually did here? The
coefficients A, B, C and D are determined by matching 1 with 2 and 3 at the
boundaries x = 0 and x = L. This can only be done for certain values of E, corresponding
to the allowed energies for the bound particle. Again, these allowed energy states are
quantized energy values. The figure below shows the wave functions and probability
densities that result for the three lowest allowed particle energies. Note that in each case
the wave functions join smoothly at the boundaries of the potential well.

Comparison of the finite with the infinite potential well shows

7-14
The fact that  is nonzero at the walls increases the de Broglie wavelength in the well
(compared with that in the infinite well), and this in turn lowers the energy and
momentum of the particle. This is a consequence of the uncertainty principle:
x p  21
increase decrease
This observation can be used to approximate the allowed energies for the bound
particle. The wave function penetrates the exterior region on a scale of length set by the
penetration depth d, given by
2
1
ψ3 (x)  ex  penetration depth d  
 2m  V0  E 
This means that at a distance beyond the well edge, the wave amplitude has fallen to e–1
of its value at the edge and approaches zero exponentially in the exterior region. At this
point, the exterior wave is essentially zero beyond the penetration distance, the allowed
energies would be those for an infinite well of length L + 2d, or
n2h2
En 
8m L  2d
2

A word of warning with this equation: the allowed energies for a particle bound to the
finite well are approximate! This equation is valid only if L >> d. Note that d must be
energy dependent since a particle with higher energy will penetrate the barrier more than
a particle with a lower energy. Therefore, the approximation is best for the lowest-lying
states and breaks down completely as E approaches the barrier height Vo.
WEB Simulation French quantum applet – very good! Go to the single well and use the
trial function option. Move the energy-level bar with the mouse to explicitly show the
wave functions.
http://www.quantum-physics.polytechnique.fr/en/index.html
Application and Interesting Point: Quantum Corrals
Analogy: standing waves on a Chladni circular plate
WEB Simulation Dan Russell’s site
The discovery of the STM's ability to image variations in the density distribution of surface state electrons
created in the artists a compulsion to have complete control of not only the atomic landscape, but the
electronic landscape also. Here they have positioned 48 iron atoms into a circular ring in order to “corral”
some surface state electrons and force them into “quantum” states of the circular structure. Here are the
various stages during the construction of the circular corral.

7-15
WEB http://www.almaden.ibm.com/vis/stm/corral.html
Below is an image of the variations in the density distribution of surface state electrons. The ripples in the
ring of atoms are the density distribution of a particular set of quantum states of the corral. The artists were
delighted to discover that they could predict what goes on in the corral by solving the classic eigenvalue
problem in quantum mechanics – a particle in a hard-wall box.

Stadium-shaped enclosure, composed of individually placed iron atoms on a copper surface, tests the ability
of researchers to contain surface-state electron density and modify its quantum properties. The structure
shown does not contain the surface electron density, although similar, circular structures do.

Reflected & Transmission of Waves


In the previous sections, we focused mainly on a particle that is confined inside a
potential well. We have seen that for these bounded particles, the Schrödinger equation
implies quantization of the allowed energies – one dramatic differences between
classical and quantum mechanics. We now focus our attention to particles striking a
potential barrier. Unlike potential wells that attract and trap particles, barriers repel them.
Because barriers have no bound states, the emphasis shifts to determining whether a
particle incident on a barrier is reflected or transmitted. This present a second example
that also has dramatic differences from classical mechanics. In classical physics when a
particle moves towards a barrier, the particle cannot cross through the barrier. However,
in quantum mechanics we will find that a particle can “tunnel” through the barrier and
emerge on the other side. This barrier penetration or tunneling has dramatic
consequences in several natural phenomena that we will consider. I will arrive at the
tunneling effect by first considering a Step Potential and the followed by a Barrier
Potential.
Step Potential
Consider a region where
0 for x  0
V( x)  
V0 for x  0
I am interested in analyzing a particle going to the right with two different energy levels:
above the step potential (E > Vo) and below it (E < Vo).
Case (1) Above the Step Potential – E > Vo

7-16
Consider a particle traveling from the left towards the right, where the total energy is
higher than the Step Barrier potential. What are the wave functions at this particular
energy? From earlier calculations, we found that
 2 2
k1
ψ1 ( x)  Aeik1 x  Beik1x ; (0  x)  E  x0
 2m
ψstep ( x)   where 
ψ2 ( x)  Ce ; ( x  0)
ik2 x 2 2
  E  V  k2 x  0
 0
2m
Note that by definition, we do not have particles coming from the right towards the left,
therefore, there are no terms with eik2 x .
The question is what do I do with this information now? I am interested in
calculating the probabilities of an incident beam of particles approaching a step barrier
and either being reflected or transmitted past this barrier. In order for me to do this, I
need to consider the total number of incident particles coming in and compare them to
the number of particles that are either reflected or transmitted. Remember that the
probability density is the [probability amplitude] 2. The probability amplitude is related to
the coefficient of the wave functions. In other words,
ψ1 ( x)  Aeik1x  Be ik1x  A  probability amplitude for the incident wave
 
  B  probability amplitude for the relfected wave
incident wave reflected wave

ψ 2 ( x)  Ce
ik2 x
C  probability amplitude for the transmitted wave
 transmitted wave 
The probability densities for particles encountering a potential barrier are
 
ψ*ψ
R  * reflected 
incident particle probability  B  probability of an incident particle getting

2


 
ψψ
incident
total probability  A2  reflected from a potential barrier
and

T
 ψ ψ
*
transmitted

transmitted particle probability C  probability of a transmitted particle

2


 ψ ψ *
incident
total probability  A2  getting by the potential barrier
Clearly, to calculate these probability densities, we must find the constants A, B and C.
This is done by applying the boundary conditions between regions I & II:
 ψ I (0)  ψ II (0)

 dψ I (0) dψ II (0)
 dx  dx
Applying these boundary conditions, we get
 ψI (0)  ψII (0)
ψI (0)  ψII (0)  A B C
dψI (0) dψII (0)
 
dx dx
dψI (0) dψ II (0)  d (eik1 x ) d (e ik1 x ) d (eik2 x ) 
  A B C 
dx dx  dx dx dx  x 0
 ik1 A  ik1B  ik2C  k1 A  k1B  k2C
To solve for constants B and C in terms of A, we setup a system of equations
A B C 1 1  A  C 
      
k1 A  k1B  k2C  k1 k1   B   k2C 
and solve them using matrix techniques:

7-17
1  C 1  1 C 2k1
A      k1  k2   C  
 1 1   k2C k1  2k1 A k1  k2
det  
 k1  k1 
To get the ratio B/A, we solve the system of equations for B/C and use the previous
result:
1 1 C  1 k k
B     k1  k2   C  B  1 2 C
 1 1  1 2 
k k C 2 k1 2k1
det  
 k1  k1 
Putting together the two results of C/A and B/C, we get solve for the ratio B/A:
k k C 2k1
B  1 2 C and 
2k1 A k1  k2
B
1
 k1  k2   C k1  k2
 
2 k1
=
A 1
2 k1  k1  k2   C k1  k2
B k k
  1 2
A k1  k2
Calculating the reflection and transmission coefficients, we get
2 2 2
k k 
B k2 C 4k1k2
R 2  1 2  and T  
A  k1  k2  k1 A 2
 k1  k2 2
Remarks
1. Note that the total probability is equal to the sum of the reflection and the
transmission coefficients:
k12  k12  2k1k2  4k1k2  k1  k2 
2 2
k k  4k1k2
R T   1 2     1
 k1  k2   k1  k2   k1  k2 2  k1  k2 2
2

No surprise here.
2. Even though E > Vo, R is not zero! This is quite surprising. The classical analogy is
that of electromagnetic radiation interacting with a transparent material. Part of the
light is reflected and part is transmitted.
WEB Simulation Step potential applet – move the energy-level bar with the mouse
to explicitly show the different wave functions.
http://www.quantum-physics.polytechnique.fr/en/index.html
3. Note the change in wavelength after the wave is transmitted: the reason is that the
KE of the particle has decreased, and therefore, the momentum of the particle has
also decreased. The de Broglie wavelength tells us that p  h /  : decrease in p
implies a longer wavelengths. Furthermore, as the energy increases well beyond the
potential step Vo, the probability of being reflected decreases as shown by the applet.
Check it out!

Case (2) Below the Step Potential – E < Vo

7-18
Classically one expects that a particle will absolutely get reflected from the barrier. This
is also true in the quantum mechanical situation, however, it is not reflected from the x =
0 boundary. The wave function has a probability that it will penetrate the barrier some
distance, and then get reflected back. Let’s analyze this to see what it is that I am talking
about.
Using the same analysis as in the previous case, we ask what the wave function is as it
encounters the barrier. For particles with energies below the step potential (E < Vo)
 2 2
k1
  ik1 x  E 
 ψ ( x )  Ae ik1 x
 Be ; (0  x )  2m
ψstep ( x)   1  x
where Estep  
ψ2 ( x)  Ce ;
 ( x  0) V  E  
2 2

 0 2m
Focusing only on the transmission coefficient, when the boundary conditions are applied
we get
ψI (0)  ψ II (0)  A  B  C
dψ I (0) dψ II (0)  d (eik1x ) d (eik1x ) d (e x ) 
  A B C   k1 A  k1B   C
dx dx  dx dx dx  x 0
The systems of equations are
 A B C  1 1  A   C 
        
 k1 A  k1B   C   k1 k1   B    C 
so that the constants C in terms of A is
1 C 2k1
A   k1     C  
2k1 A  k1   
The probability density for transmission is given by

T
 ψ ψ
*
transmitted

Ce  x
2


2
C 2 x
e
 ψ ψ
*
incident
Aeik1 x
2
A

2k1
 T e2 x
 k1   
Plotting the wave function and probability density for this particle in the barrier results in

Interpretation
2
It is clear from the plots of ψ and ψ that the wave function does not go to zero at x = 0
but decays exponentially. The wave penetrates slightly into the classically forbidden
region x > 0 but eventually is completely reflected.
WEB Simulation Step potential applet – move the energy-level bar below the step
potential

7-19
http://www.quantum-physics.polytechnique.fr/en/index.html

Example 7.4 (P6.41)


A free particle of mass m with wave number k1 is traveling to the right. At x = 0, the
potential jumps from zero to Vo and remains at this value for positive x.
(a) If the total energy is 2Vo, what is the wave number k2 in the region x > 0? Express
your answer in terms of k1 and Vo.
(b) Calculate the reflection coefficient R at the potential step.
(c) What is the transmission coefficient T?
(d) If one million particles with wave number k1 are incident upon the potential step, how
many particles are expected to continue along in the x direction? How does this
compare with the classical prediction?
Solution
(a) The two energies before and after the barrier are
 2 k12
  x  0
2m
E  2V0   2 2
 k2  V  x  0 
 2m 0

These are two equations in k1 , k2 , and Vo; solve for k2 by setting up a ratio:
k1 
2 2
E  2V0  
2m  2V0 21m 2 k12 1
2 2 2 2 
  1 2 2  k2  k1
k2 k2  V0 k2 2
E  2V0   V0  V0  2m

2m 2m  
(b) The reflection coefficient is given by
2 2
2
k k 
B
2
 k1  1
k  1 1 
2 1
R 2  1 2   R   2

A  k1  k2   k1  1
k  1 1 
   
1
k2  k1 2 1 2
2

 R  0.0294
In other words, 2.94% of the incident particles are reflected from the potential step.
(c) The transmission coefficient is
R  T  1  T  1  R  1  0.0294
T  0.971
(d) This means that 97.1% of the million particles ( 9.71105 ) will continue to past the
potential step in the positive x direction. However, classically, 100% would continue
on.
Barrier Potential
Consider a region where
0 for x  0, x  a
V( x)  
V0 for 0  x  a
I am interested in analyzing a particle coming from the left with an energy level below the
barrier potential (E < Vo). The wave functions for the three regions are
ψ I ( x)  Aeik1 x  Beik1 x ; (0  x)  2 2
k1
  E 
  2m
ψ barrier ( x)  ψ II ( x)  Ce x  De x ; (0  x  a) where Ebarrier  
V  E  
2 2

ψ
 III
 ( x )  Fe ik1 x
; (0  x  a )  0
2m
One can calculate the transmission coefficients by applying boundary conditions:

7-20
ψ I (0)  ψ II (0) 
 A  B  C  D
dψ I (0) dψ II (0)   
 ikA  ikB   D   C
dx dx 
ψ I (0)  ψ II (0)   a
 Ce  De a  Feika
dψ I (0) dψ II (0)    a  x
  De   Ce  ikFeika
dx dx  
Setting up a matrix equation as before
0     0 
A
1 1 1 1
  B
 ik ik   0    0
C    
0 0 e  x e x eika     0 
  D  
0 a  a ika  
 0  e  e ike  F   0
 
one can solve this for the transmission coefficient using somewhat strenuous
calculations:
1
 

T
ψ ψ*
transmitted

Fe ik1 x 2


F
2 
 1 
sinh  a 
2 

 ψ ψ *
incident
Aeik1 x
2
A  E  E 
 4 V 1  V  
 0  0 

where sinh  a  (e a  e a ) / 2 .


Remarks
1. Note that this equation is valid only for energies E below the barrier height Vo. For E
> Vo, α becomes imaginary and sinh(αa) turns oscillatory, as expected.
2. Use the Quantum applet to see the behavior of the tunneling wave
WEB Simulation barrier potential applet
http://www.quantum-physics.polytechnique.fr/en/index.html
Example 7.5
Two copper conducting wires are separated by an insulating oxide layer (Cu O).
Modeling the oxide layer as a barrier potential of height 10.0 eV, estimate the
transmission coefficient for penetration by 7.00-eV electrons (a) if the layer thickness is
5.00 nm and (b) if the later thickness is 1.0 nm.
Solution
We calculate  for this case using
2
2 2me (V0  E ) 2  0.511  106 eV/c2  (10eV  7eV)
V0  E      8.87 nm 1
  
2
2m c/c 197.3 eV nm/c

The transmission coefficient is


1
 
 
1
 sinh 2  a 
 sinh 2 8.87 nm 1 a 
T  1    1  
 4 E 1  E    48 / 49 
  V   
V0  0 

(a) For α = 5.00 nm, we get T = 0.963 1038 ; a fantastically small number on the ode of
10-38!.
(b) For α = 1.00 nm, we get T = 0.657 107 ; We see that reducing the layer thickness by
a factor of 5 enhances the likelihood of penetration by nearly 31 orders of magnitude.
7-21
Interesting Points and Applications
STM

Alpha Decay

Ammonia Inversion

Harmonic Oscillator
The second problem solved by Schrödinger was the blackbody radiation problem.
Electrons where viewed as acting like harmonic oscillators in the walls of the blackbody.
Any particle limited to small excursions about any stable equilibrium point behaves as if
it were attached to a spring with a force constant K prescribed by the curvature of the
true potential at equilibrium. The particle is subjected to a linear restoring force such that
V  x 
F   Kx    V  x   12 Kx 2 K  12 m 2 x 2
x 
m

Plotting this potential function for a particle energy level, we find

Solving the Schrödinger equation is beyond the scope of this course, however, one can
get make an educated guess as to what one expects. Inside the potential well one
expects free-particle type behavior, whereas at the potential barrier there should be
exponential decaying wave functions. This is what is observed in the Schrödinger
solution:
2
d 2 ψ(x) 1
  2 m 2 x 2 ψ(x)  Eψ(x)  ψ( x)  Cn e  m x
2

2
/2
H n ( x)
2m dx
 En   n  12   n  0,1, 2,3,...

7-22
7-23

You might also like