Вы находитесь на странице: 1из 41

Lecture Notes for Physical Chemistry II

Quantum Theory and Spectroscopty


Eric R. Bittner
March 1, 2006
Figure 1: Participants of the 1927 Solvay Conference on Physics and Chemistry. Pictured here are
many of the founders and pioneers of modern physics and quantum theory. Of the people pictured
here 18 would win Nobel Prizes for their discoveriesothers would not. One may even know how
to spell h-bar.
1
1 Classical Mechanics
Before launching into the details of quantum theory, it is important and necessary to review and
establish the physical and mathematical description of how particles in Nature move about. Our
basic concepts are rooted rmly in the classical laws of motion as establied by Isaac Newton over
300 yrs. ago. These laws of motion, and their generalization by Lagrange, Hamilton, and eventually
Einstein, are accurate and highly predictive for describing the motion of objects that are not too
small. For atoms and molecules, we will need to rene these laws of motion.
1.1 Newtons Postulates
Why do things move? Why does an apple fall from a tree. This is usually the rst sort of problems
we face in trying to study the motion and dynamics of particles and develop laws of nature that
are independent of a particular situation.
We understand the concept of force. We all have pushed, pulled, or thrown something. Those
actions require an action or force from the muscles in our body. Newton proposed a set of basic
rules or postulates which he thought could describe the rules that all objects obey under the
inuence of any kind of force.
Postulate 1 Law of Inertia: A free particle always moves without acceleration.
That is, a particle that is not under the inuence of an outside force moves along a straight line at
constant speed, or remains at rest.
Postulate 2 Law of Motion: The rate of change of an objects momentum is equal to the force
acting upon it.
d p
dt
=

F.
This is equivalent to

F = ma where a = dv/dt is the acceleration. Note, that in Newtons rst
postulate, we assume that the mass does not change with time.
Postulate 3 Law of Action: For every action, there is an equal and opposite reaction.

F
12
=

F
21
This is to say that if particle 1 pushes on particle 2 with force F, then particle 2 pushes on particle
1 with a force F.
In SI units, the unit of force is the Newton, 1N = 1kg m s
2
.
2
Let us apply these postulates to study some simple motions. First, we assume that the forces
depend only on position and time. Secondly, we require that the mass remain a constant. Under
these conditions, the second postulate becomes,

F =
d
dt
(mv) = m
d
dt
v = ma
Since the acceleration is the second-derivative of the position with respect to time, we have a
second-order ordinary dierential equation for the positions,
d
2
x
dt
2
= m

F
where

F =

F(x).
1.1.1 Example 1: Hookes law
Lets consider a simple spring placed in the horizontal direction with a mass m as shown in the
gure here. If we displace the mass to the right, and release, the mass will spring back. Lets take
the force to be proportional to the displacement, F = kx where k is a constant.
The resulting equations of motion are
kx = m x
where x denotes the second-time derivative. This we can also write as
v =
k
m
x
Thus, we look for solutions of the form
v(t) = a sin(t) +b cos(t)
with boundary conditions, v(0) = 0 at t = 0. Thus b = 0 and v(t) = a sin(t). Since dv/dt =
k/mx(t),
v = a cos(t) =
k
m
x(t)
thus,
x(t) = a
m
k
cos(t)
3
Since x(0) = x
o
is the original displacement,
a
m
k
= x
o
.
Also, we know x = v, so
x =
2
x
m
k
sin(t) = v(t)
We therefore conclude that
2
m/k = 1, =
_
k/m and a = x
o
. Thus, the particle moves
according to
x(t) = x
o
cos(t)
with velocity
v(t) = x
o
sin(t).
The frequency, is related to the period of oscillation, , by 2 = and = 1/. Thus, = 2/.
1.1.2 Example 2: Motion in a gravitational potential.
Let us now consider a two dimensional example where we have a particle launched upwards at some
initial velocity and we wish to predict where it will land. We shall neglect frictional forces.
The equations of motion in each direction are as follows. In the vertical direction,
m y = mg
where g is the gravitational constant and the force mg is the attractive force due to gravity . In
x, we have
m x = 0
4
since there are no net forces acting in the x direction. Hence, we can solve the x equation im-
mediately since v
x
= 0 and thus, x(t) = v
x
(0)t + x
o
= v
o
t cos(). For the y equation, denote
v
y
= y,
m
d
dt
v
y
= mg
Integrating, v
y
= gt +const. Evaluating this at t = 0, v
y
(0) = v
o
sin() = const. Thus,
v
y
(t) = gt +v
o
sin().
This we can integrate
_
dy =
_
(gt +v
o
sin())dt
i.e.
y = v
o
sin()t
g
2
t
2
So the trajectory in y is parabolic. To determine the point of impact, we seek the roots of the
equation
(v
o
sin()t
g
2
t
2
) = 0.
Either t = 0 or
t
I
=
2
g
v
o
sin().
We can now ask, what angle do we need to point our canon to hit a target X meters away? In time
t
I
the canon ball will travel a distance x = v
o
cos()t
I
. Substituting our expression for the impact
time:
X = v
2
o
cos()
2
g
sin() =
v
2
o
sin(2)
g
Thus,
sin(2) =
g
v
2
o
X.
One can also see that the maximum range is obtained when = /4.
1.2 Hamiltons Equations
It is often easier and more convenient to to express Newtons equations of motion as two rst order
dierential equations rather than a single second-order dierential equation. Both are equally valid.
However, it is far easier to to obtain a equations of motion in other coordinate systems that the
x, y, z Cartesian coordinates is we work with a more general set of equations. For this, we dene a
more general quantity for the energy of a system,
H = T(v
1
, v
2
, , v
N
) +V (q
1
, q
2
, q
N
)
where T is the kinetic energy which depends upon the velocities of the N particles in the system
and V which is the potential energy describing the interaction between the all the particles and
any external forces. V is the energy of position where as T is the energy of motion. For a single
particle moving in three dimensions:
T =
1
2
m(v
2
x
+v
2
y
+v
2
z
).
5
If we write the momentum as p
x
= mv
x
then
T =
1
2m
(p
2
x
+p
2
y
+p
2
z
).
Notice that we can also dene the momentum as the velocity derivative of T.
p
x
=
T
v
x
.
This denes a generalized momentum in such q
x
is the conjugate coordinate to p
x
and (q
x
, p
x
) are
a pair of conjugate variables. This relation between T and p
x
is important since we can dene the
canonical momentum in any coordinate frame. In the Cartesian frame, p
x
= mv
x
. However, in
other frames, this will not be so simple.
We can also dene the following relations
H
p
i
=
T
p
i
=
p
i
m
=
q
i
t
(1)
H
q
i
=
V
q
i
= F
i
=
(mv
i
)
t
(2)
where i now denotes a general coordinate (not nesc. x, y, z). In short, we can write the following
equations of motion:
H
p
i
=
q
i
t
(3)

H
q
i
=
p
i
t
(4)
These hold in any coordinate frame and are termed Hamiltons Equations.
1.2.1 Example: Hamiltons Equations in Polar Coordinates
Lets consider the transformation between polar and two dimensional Cartesian coordinates, x and
y.
x = r cos & y = r sin
Our Hamiltonian in x, y coordinates is
H =
m
2
(v
2
x
+v
2
y
) +V (x, y)
Thus,
v
x
=
dx
dt
= v
r
cos v

r sin (5)
v
y
=
dy
dt
= v
r
sin +v

r cos (6)
v
2
= v
2
x
+v
2
y
= v
2
r
+v
2

r
2
6
Thus,
H =
m
2
_
v
2
r
+v
2

r
2
_
+V (r, ).
We can now proceed to write this in terms of the conjugate variables,
p
r
=
T
v
r
= mv
r
(7)
p

=
T
v

= mv

r
2
(8)
Note that p

is the angular momentum of the system. Thus, we can write


H =
1
2m
_
p
2
r
+
p
2

r
2
_
+V (r, )
Next, consider the case where V (r, ) has no angular dependence. Thus,
p
r
t
=
H
r
=
p
2

mr
3

V
r
(9)
p

t
=
H

=
V

= 0 (10)
r
t
=
H
p
r
=
p
r
m
(11)

t
=
H
p

=
p

mr
2
(12)
Notice that p

does not change in time...i.e. the angular momentum is a constant of the motion.
7
The radial force we obtain from p
r
/t = F
r
F
r
=
p
2

mr
3

V
r
The rst term is constant (since p

= const.) and represents the radial force produced by the angular


momentum. It is always points outward towards larger values of r and is termed the centrifugal
force. The second term is the force due to the attraction between the moving object and the origin.
It could be the gravitational forces, the Coulombic force between charged particles, etc. We can
also write the force equation as
F
r
=
(mv

r
2
)
2
mr
3

V
r
=
mv
2

r

V
r
If the two forces counter balance each other, then the net force is F
r
= 0 and we have
mv
2

r =
V
r
.
Since v

=

= const., = t +const. where is the angular velocity. Thus, using v

= , we can
write
m
2
r =
V
r
.
Finally, we note that the linear velocity is related to the angular velocity by = vr,
mv
2
r
=
V
r
.
Hence
mv
2
= 2T = r
V
r
.
1.3 Virial Relation
Lets consider an important case where V (r) r
n
. First, lets dene a quantity, G = pr so that

G = rp +r p =
We can manipulate this a bit, viz:

G = mr r r + pr (13)
= 2T +Fr (14)
Thus,
d
dt
p(t)r(t) = 2T +Fr =
dG
dt
Now, we take the time average
1

_

0
_
dG
dt
_
dt = 2T +Fr =
1

(G() G(0))
8
If the motion is periodic (corresponding to bound motion) and is equal to that period,
1

(G() G(0)) = 0
and we can write
2T = Fr.
This is an important relation since if we have a potential of the form V = Cr
n
, then F = dV/dr =
Cnr
n1
. Hence rF = nV Thus, we have a relation between the average kinetic energy and the
average potential energy
T =
n
2
V (15)
This is what is called the Virial Relation. Remember this relation!
1.4 Angular Momentum
We noted above that if you have a radial force, then the angular velocity and angular momentum
were constants of the motion. In general, the angular momentum is dened as the cross product
between a radial vector locating the particle and its linear momentum

M = r p
Cross products are equivalent to taking the determinant of a matrix

M =

i

j

k
x y z
p
x
p
y
p
z

(16)
where

i,

j, and

k are the unit vectors along the x, y, z axes. Evaluating the determinant gives

M =

i(yp
z
zp
y
)

j(xp
z
zp
x
) +

k(xp
y
yp
x
) (17)
=

iM
x
+

jM
y
+

kM
z
(18)
For motion in the x y plane, the only term that remains is the M
z
term, indicating that the
angular momentum vector points perpendicular to the plane of rotation,
M
z
= (xp
y
yp
x
) = m(xv
y
yv
x
)
Since we noted that the angular momentum is a constant of the motion, we must have dM
z
/dt = 0.
Lets check.
dM
z
dt
= m(v
x
v
y
v
y
v
x
+xa
y
ya
x
)
where a
x
= v
x
is the acceleration in x. Thus,
dM
z
dt
= (xF
y
yF
x
).
9
If the force is radial, F
x
= F
r
cos() and F
y
= F
r
sin(). Likewise, x = r cos() and y = r sin().
Putting this in we have
dM
z
dt
= rF
r
(sin() cos() sin() cos()) = 0
Taking = t as above where is the angular frequency, and using v
x
= r sin(t) and y
y
=
+r cos(t) we can also write
M = m(v
x
y = v
y
x) = mvr(sin
2
(t) + cos
2
(t)) = mvr
1.5 Classical motion of an electron about a positive charge (nucleus)
Now we are ready to describe the motion and energy of a charged particle about a nucleus. This will
provide us with a classical description of the Hydrogen atom and any hydrogenic (hydrogen-like)
species. We shall need these results to begin our discussion of quantum theory.
For an electron bound to a nucleus of charge +Ze, the Coulomb force holding the electron to
its orbit is counter balanced by the centrifugal force as in the equations we derived above. The
Coulomb force is
F =
1
4
o
Ze
2
r
2
Note that we shall eventually use units such that 4
o
= 1, for now we keep the SI units. Since the
Coulomb force counterbalances the centrifugal force
1
4
o
Ze
2
r
2
= mr
2
= m
v
2
r
As above, we can see the vIrial relation between the kinetic and potential energies
T =
1
2
mv
2
=
1
4
o
Ze
2
2r
=
1
2
V
where
V =
1
4
o
Ze
2
r
is the Coulomb potential. Since the classical energy is E = T +V we can write
E =
Z
4
o
_
e
2
2r

e
2
r
_
=
1
4
o
Ze
2
2r
1.6 Appendix: Hamiltons equations in spherical polar coordinates
You may have noticed that dx
2
+dy
2
+dz
2
is the dierential line element, dl
2
in Cartesian coordinates
and that v
2
= (dl/dt)
2
= (dl)
2
/(dt)
2
. This allows us to easily deduce the kinetic energy function
10
in any arbitrary coordinate system that has a dene length (or metric). For example, in spherical
polar coordinates,
dl
2
= dr
2
+r
2
d
2
+r
2
sin
2
d
2
Hence,
v
2
=
_
dr
dt
_
2
+r
2
_
d
dt
_
2
+r
2
sin
2

_
d
dt
_
2
(19)
= v
2
r
+r
2
v
2

+r
2
sin
2
v
2

(20)
Thus, since T =
1
2
mv
2
,
p
r
=
T
v
r
= mv
r
(21)
p

=
T
v

= mr
2
v

(22)
p

=
T
v

= mr
2
sin
2
v

(23)
From here you can work out the equations of motion for a particle in spherical polar coordinates.
2 Quantum Theory
2.1 Bohrs model for Hydrogen
The above equations are great. However, theres one problem. According to electrodynamics, an
accelerating charge radiates electromagnetic energy. Hence, every classically bound electron should
spiral inwards towards the nucleus giving o a burst of x-ray and gamma ray radiation. This is
not observed. Moreover, there is no restriction about which radius we choose for the electron...we
can have any energy we want. However, we also know that Hydrogen emits and absorbs light only
at specic wavelengths. In deed, Balmer back in 1885 demonstrate that all of the absorption (and
emission) lines of H could be t to a single empirical equation

1
= 109677cm1
_
1
n
2

1
m
2
_
where n and m are integers n = 1, 2, 3, 4 and m = n+1, n+2, n+3, . The numerical constant
109,677 cm
1
is the Rydberg constant (denoted R
H
).
About the turn of the last century, results such as this and results demonstrating the photoelectric
eect and Plancks theory of blackbody radiation (described wonderfully in you text) indicated
that ther was something wrong with the way we view the physical world when it came to atoms
and molecules. Niels Bohr attempted to explain the spectral observations and combine Plancks
notion of quantized energies with a radation-less orbit of the electron about the nuclei. In doing
so, he made a number of remarkable leaps of faith. Bohr postulated the following:
11
1. A discrete spectrum implies discrete energy levels and that the energy absorbed or emitted
are the energy dierences between these levels.
E = h = W
f
W
i
2. Electrons undergo transition between these levels via absorbing or emitting light.
3. Electrons are bound to the nuclei via Coulombic forces and obey classical mechanics.
4. The Rydberg equation gives us the E, hence since = c
E =
hc

= 109677cm1
_
1
n
2

1
m
2
_
5. The energy levels are then
W
n
= hcR
H
1
n
2
6. Correspondence Principle: At large values of n the classical emission frequency must be equal
to the electrons orbital frequency, as required by classical electrodynamics.
Now Bohr makes a dramatic leap of faith. The W
n
are the quantum energies. Bohr assumes that
quantum = classical (24)
hcR
H
1
n
2
=
1
4
o
Ze
2
2r
(25)
According to Bohrs Correspondence principle, the angular frequency of an electron for large values
of n must be equal to the classical radiation frequency. (I now drop the 4
o
)
=

e
2
4mr
3
where = 1, 2, 3 . For large values of n the quantum frequency is
= R
H
c
_

1
n
2
+
1
(n )
2
_
(26)
=
2R
H
c(1 /(2n))
n
3
(1 2/n +
2
/n
2
)
(27)
where = 1, 2, 3, is the change in quantum number as the electron goes from a higher energy
orbit to a lower energy orbit. For large values of n
=
2R
H

n
3
Thus, the frequencies are integer multiples of some fundamental frequency
o
. Again, we use
quantum = classical (28)
2R
H

n
3
=

e
2
4mr
3
(29)
12
Now we start canceling terms. To eliminate n we use the expression for the energy levels
n =

R
H
hc
|W
n
|
,
and for r, the classical radius
W =
e
2
2r
which gives, r = e
2
/(2|W|). Turning the crank and eliminating variables where possible:
R
H
=
2
2
me
4
ch
3
which gives the Rydberg constant entirely in terms of fundamental physical constants: m: mass of
electron, c: speed of light, and h Plancks constant. Thus, we can write the energy levels in terms
of no adjustable parameters:
W
n
=
R
H
hc
n
2
=
2me
4
h
2
1
n
2
.
Furthermore, we can go on to show that since
W =
e
2
2r
=
R
H
hc
n
2
we can solve for the orbital radius r,
r =
e
2
2
n
2
R
H
hc
giving xed circular orbitals for the electrons
r
n
=
h
2
4
2
me
2
n
2
=
h
2
me
2
n
2
where n = 1, 2, 3 . The innermost orbital, with n = 1 is the Bohr radius, a
o
. Hence, r
n
= n
2
a
o
.
Finally, since the electron is orbiting about the proton, it must have angular momentum. From
classical mechanics M = mr
2
= mvr. Again, following our prescription
quantum = classical (30)

2
2
h
me
4
1
n
2
=
e
2
2r
(31)
from the energy expression. Now, take
m
2
r =
e
2
r
2
and solve for e
2
e
2
= mr
3

2
and plug this back into the energy equation above
quantum = classical (32)

2
2
h
me
4
1
n
2
=
mr
3

2
2r
=
M
2
2mr
2
(33)
13
where M is the angular momentum we derived above. Thus,
M
2
=
2
2
me
4
h

2mr
2
n
2
Taking our expression for the quantized radii from above,
M
2
=
_
h
2
_
n
2
= h
2
n
2
Thus, M = hn is the angular momentum! It too is quantized in units of Plancks constant over 2.
2.2 Extra: Do the orbitals need to be circular?
Here we pose an interesting question. is there any reason to believe that the electrons or-
bits need to be strictly circular (or rather elliptical if we account for motion about the center
of mass of the electron-proton system). A bit of dimensional analysis indicates that the units
of h are energy time. That is also equivalent to momentum length. If we imagine plot-
ting the position versus the momentum of a particle on the p x plane, then momentum
length corresponds to some area encompassed by a closed path as shown in the Figure below
The closed loop in red encompasses an area equal to
area =
_
p(x)dx
If we assume that energy is conserved, then the energy is a constant along the red loop, H(p, x) =
E = const.
Planck required that E = hn for the quantized harmonic oscillator levels, but that equation is
only applicable to harmonic systems. If quantization condition is general, then we should see it
14
appear in a more general context. Lets take a harmonic oscillator system as an example.
H(p, x) =
p
2
2m
+
k
2
x
2
= E
This is the equation for an ellipse.
1 =
p
2
2mE
+
k
2E
x
2
with major and minor axes a =

2mE and b =
_
2E/k. The area of an ellipse is
A = ab.
Hence, the are on the px plane (called phase space) is
A =

2mE
_
2E/k = 2E
_
m
k
=
2E

=
E

Since we assumed A to be quantized in multiples of h


A =
E

From Planck: E = hn thus,


A = hn.
Hence, the electrons need not move in strictly circular orbits, they only need to move in closed
paths such that
nh =
_
p(x)dx.
This is the Bohr-Sommereld quantization condition.
2.3 Wave-Particle Duality
In 1922, Arthur Holly Compton at the University of Chicago performed a series of remarkable
experiments by scattering x-rays from atoms. He observed that the scattering wavelength

was
always greater than the incident wavelength. To explain this, he assumed that the energy lost was
due to the particle-like collision between a phonon and an electron and that some of the incident
energy was transfered to the electron.
From energy conservation:
h = h

+
mv
2
2
Furthermore, the x and y components of the momentum must also be conserved.
p
x
= p

cos() +mv cos() = p (34)


p
y
= p

sin() +mv sin() = 0 (35)


Here, p and p

are the incident and nal momenta of the photon, and v is the nal velocity of the
electron and we take the incident momentum to be along the x axis.
mv cos() = p p

cos() (36)
mv sin() = p

sin() (37)
15
Figure 2: Compton scattering experiment
Squaring both and adding
m
2
v
2
= p
2
+p
2
2pp

cos
Going back to the energy equation:
m
2
v
2
= 2m(h h

)
and equating the last two equations:
2m(h h

) = p
2
+p
2
2pp

cos
Compton then postulated that the momentum of a photon is given by h/ = h/c. Thus, replacing
the frequency with c/,
2mch
_
1

_
= h
2
_
1

2
+
1

2
_

2h
2

cos
Taking common denominators,

h
2mc
_
1

2
+
1

2
_

h
mc
cos .
Finally, if the change in wavelength is small,

and


2

2
, we have

= +
h
mc
(1 cos )
which is the Compton scattering formula. The factor h/mc is the Compton wavelength (
c
=
0.024263

A). Since x-rays are on the order of 1-3



A, the assumption that =

is pretty accurate.
In fact, if you do a relativistic treatment, you do not need to even make that assumption.
16
Figure 3: Arthur H. Compton on the cover of 13-Jan-36 Time magazine
If we re-write this as an energy equation
E

=
Emc
2
mc
2
+E(1 cos )
Since m is the mass of the electron, mc
2
= 0.5MeV is the rest-energy of the electron. You can also
write this as
1
E

=
1
E
+
1
mc
2
(1 cos )
Plotting 1/E

vs 1 cos should give a straight line with the slope being 1/m
e
c
2
.
2.4 De Broglies Matter Waves
Louis de Broglie rationalized that if light could behave as both particle (as in the Compton ex-
periment) and wave-like (as in diraction experiments), then so should ordinary matter such as
electrons, H atoms, neutrons, etc. For light:
E = h = pc
where pc = E is from Einsteins relativity. Thus, = c/, which is what Compton also used.
Consequently, = h/p. For particles with mass, p =

2mE, thus the wavelength of a particle with


17
Table 1: Compton scattering data. This is data I measured back in 1987 in an Experimental
Physics course in college using a
137
Cs source. Here we measured the energy of the scattered x-ray
from a sample. From this, you can calculate the rest-mass of the electron and the incident energy
of the x-ray emitted by the source.
(degrees) E

(MeV)
30 0.562
45 0.464
60 0.3955
75 0.339
90 0.29
105 0.244
120 0.223
135 0.205
mass is
=
h

2mE
.
This served as a the basis of de Broglies PhD thesis in 1923. When asked during his PhD examina-
tion, just how one may observe such matter waves he replied that one should be able to diract
very light particles such as electrons or He atoms from a surface, just as one can do x-ray dirac-
tion. This suggestion was tested by Davsson and Germer and eventually led to the development
of a number of powerful analytical techniques for analyzing the structure of crystal surfaces (most
commonly LEED). For this both de Broglie (in 1929) and Davisson (with Thompson in 1937) were
awarded Nobel Prizes.
To come full circle, we ask How many de Broglie wavelengths does an electron have in the H
atom?
=
h

2mE
(38)
= h(2me
2
/2r)
1/2
(39)
= h
_
2me
2
_
me
2
h
2
n
2
__
1/2
(40)
=
hhn
me
2
(41)
= 2n
h
2
me
2
= 2na
o
(42)
Thus, for n = 1, corresponds to the circumference of the rst Bohr radius. Hence, we can imagine
the electron as a standing wave on a ring of radius a
o
18
Figure 4: Experimental data from 1987 Compton scattering experiment at Valparaiso Univ.
2.5 Quantum Theory
Bohrs model of the Hydrogen atom was successful in that it gave us a radically new way to look
at atoms. However, it have serious shortcomings. It could not be used to explain the spectra of He
or any multi-electron atom. If could not predict the intensities of the H absorption and emission
lines. With de Broglies hypothesis that matter was also wave-like, there arose a question at the
1927 Solvey conference of what is the wave equation? De Broglie could not answer this, however,
Erwin Schrodinger gave it go. Over the following year, he deduced the general form of the equation
that bears his name and applied it successfully to the hydrogen atom. What emerged were a new
set of postulates, much like Newtons, that laid the foundations of quantum theory.
Postulate 1 Matter is described by a wave function, (x, t), which is a function of time and space.
It is nite, single-valued, smooth, and continuous at all points.
This wave function is in general a complex valued mathematical object. It is square-integrable in
that
_
+

|(x, t)|
2
dx <
Hence, this integral can be re-cast so that
_
+

|(x, t)|
2
dx = 1
for all time.
Postulate 2 For every physical observable, there is a corresponding quantum mechanical operator,

O, that is Hermitian. The commutator of conjugate variables is equal to ih.


19
Figure 5: Louis de Broglie
This is Bohrs correspondence principle. Below are the classical quantities and their corresponding
quantum operator.
Table 2: Quantum and classical operators
Classical Quantum
Position x x
Momentum p
x
h
i

x
Time t t
Energy E hi

t
Postulate 3 Quantum equations are derived from classical equations by direct substitution of the
quantum operators in to the classical equations in Cartesian coordinates, then allowing the resulting
new operator to act upon, .
For example, the Hamiltonian operator, which gives the energy of the system, is given classically
as
H(p, x) =
p
2
2m
+V (x).
20
We can immediately write the quantum operator form as

H =
_

h
2
2m

2
x
2
+V ( x)
_
= hi

t

This is the time-dependent Schrodinger equation that describes the time evolution of the wave
function, (x, t).
Postulate 4 The quantity |(x, t)|
2
dx is proportional to the probability of nding the system in a
volume dx about x at time t. (Born interpretation)
1
If (x, t) is normalized by
_

|(x, t)|
2
dx = 1
then |(x, t)|
2
is a normalized probability distribution. This invokes a statistical interpretation of
the wavefunction such that (x, t) is the probability amplitude.
Since we are discussing probability distributions, the expected value of any physical measurement
can be derived from the normalized wave function by taking the appropriate integral of the form
O =
_

(x, t)

O(x, t)dx.
If the operator is simply a function of x, then
O =
_
O(x)|(x, t)|
2
dx.
Other wise, if it is a dierential operator,
O =
_

(x, t)(

O(x, t))dx.
For example, lets consider the case the expectation value of the Hamiltonian operator

H on
(x) =
_
2

_
2
xe
x
2
/2
is normalized, (you can check). New we take the expectation value of

H by writing

H =
h
2
2m
_
+

(x)

(x)dx +
_
+

|(x)|
2
V (x)dx
Since in this case is real, we can dispense with the

and write

= . The second integral


will depend upon the actual functional form of the potential. The rst we can evaluate by either
taking the second-derivative of and then integrating, or integrate by parts. I demonstrate the
latter.

dx =
_
(
x

)dx =

+
_

dx
1
Incidentally, Olivia Newton-John is Max Borns granddaughter. I guess we now know what she meant by Lets
get physical.
21
the boundary term vanishes if is a proper wavefunction. Hence

h
2
2m
_

dx =
h
2m
_

dx =
h
2m
_
(

)
2
dx
Inserting

and squaring
_
(

(x))
2
dx =
2

_
e
x
2
_
1 x
2
+x
4
_
dx =
3
2
If we take V (x) = kx
2
/2 as our potential.
k
2
_
x
2

2
dx =
3
2
Thus,

H =
3
2
_
h
2
2m
+
k
2
_
=
3
2
h
_
h
2m
+
m
2h
_
2.6 Stationary State Schrodinger Eq.
For stationary probability distributions, P(x)dx does not change in time. That is to day, if we have
a stationary quantum system, then ||
2
at one time must be the same function at some later time
and also at some earlier time. To see this, take (x, t) and write it as
(x, t) = f(t)(x)
and substitute this into the time-dependent Schrodinger equation. For time independent Hs

H = f(t)

H = ih(x)

f
Divide both sides by ,

= ih

f
f
Since the left hand side involves only spatial derivatives and the right hand side only time-derivative,
the left and right hand side must be equal to a common constant. Thus,
ih

f = Kf
Weve solved equations like this in kinetics, so, we know immediately that
f(t) = Ce
iKt/h
where C is some constant of integration. Likewise,

H = K.
and we have
(x, t) = C(x)e
iKt/h
and |(x, t)| = |(x)|
2
is no longer a function of time.
22
What is K? To determine K, consider the energy expectation value, assuming is already nor-
malized:
E = ih
_

(

t
)dx (43)
= ih
_
|C|
2
|(x)|
2
e
+iKt/h
iK
h
e
iKt/h
dx (44)
= |C|
2
K
_
|(x)|
2
dx (45)
= K (46)
where we have taken |C|
2
= 1. Thus, the eigenvalue equation

H = E.
gives E as the energy eigenvalue and as the eigenfunction. This is the time-independent Schrodinger
equation.
2.7 Quantum Math
We have introduced a number of terms and mathematical concepts that are most likely unfamiliar
to most of you. So, let us take the time to learn these since we will need to use them repeatedly in
the rest of this course. Part of the diculty with quantum mechanics is that there is essentially a
new mathematical language that needs to be mastered in order to move forward with your study
of this topic.
First of all, were dealing with complex valued functions and quantities. Complex numbers are
extensions of the real numbers in two dimensions. If we take the real numbers as lying along the
x axis, and the imaginary numbers to lie along the y axis, we can dene a complex number z in
terms of its real and imaginary components.
z = x +iy
where i =

1. We can also dene a complex number as a magnitude and phase


z = re
i
expanding the exponential as a Taylor series:
z = r(1 +i

1!
+i
2

2
2!
) (47)
= r((1 +i

1!


2
2!
+ ) (48)
= r(cos() +i sin()) (49)
A complex conjugate is formed by taking i i. This is denoted by a

.
z

= x iy = re
i
Note that z

z = r
2
. Thus, |z|
2
= r
2
= x
2
+ y
2
. Hence we have the geometric interpretation of r
being the polar radius and being the polar angle dening some point on the x, y plane.
23
2.7.1 Operators
Operators are mathematical operations. Its an instruction or set of instructions on what you are
supposed to do to what is it acting upon. Typically, operators act on what ever is to their right,

Of = g
where f and g are functions and

O is some operation. For example, we have can write the derivative
operator as

D =
d
dx
thus,

Df =
df
dx
= f

= g.
Or, we can have the multiply by x operator,
x = x
so that
xf = xf = g.
We also have the identity operator,

If = f
and so on.
Linear operators are such that,

O(f +g) =

Of +

Og
the action of the operator can be distributed over the various objects. Clearly, the derivative
operator and the multiplication by x operators are both linear, but the operator

Lf =
d log(f)
dx
=
1
f
df
dx
is not a linear operator since

L(f +g) = (

Lf) + (

Lg))
Order of Operations is Important! If you can switch the order of operations, the operators commute
with each other. Many operators do commute with other operators. However, as we shall see later
on, the fact that certain quantum mechanical operators do not commute plays a crucial role in
both the implementation and interpretation of quantum theory. Consider the commutation of the
derivative operator and the multiply by x operator:
x

Df = xf

Likewise

D xf = f +xf

Hence,

D xf = x

Df.
24
We can test whether or not two operators commute by subtracting one arrangement from another:

D xf x

Df = [

D, x]f = xf

f xf

= f
The [A, B] is called a commutation bracket and whenever you see one, you immediately can write
[A, B] = A.B B.A. Note that in testing the commutation of the two operators, we needed to
supply a test function, f. From the above example, have an important result,
[

D, x] = 1.
2.7.2 Eigenvalue/Eigenvector Equations
We shall also encounter operator equations of the form

Of = af
where a is a numerical constant. This particular type of operator equation is termed an eigenvalue
equation and f is an eigenfunction of operator

O. For example, let f = exp(imx).

Df =
d
dx
e
imx
= ime
imx
= imf
hence f is an eigenfunction of the derivative operator with eigenvalue im. Likewise,

D
2
sin(kx) = k
2
sin(kx)
But,

Dsin(kx) = k cos(kx) = xsin(kx)


hence, sin(kx) is an eigenfunction of the second-derivative operator, but not the rst derivative
operator.
Hermitian Operators: Hermitian operators are operators with real eigenvalues.
They also have the property that
_

(x)(

G(x))dx =
_
(

G(x))

(x)dx =
_

(x)(

G(x))dx.
Thus,
_

(x)(

G(x))dx =
__

(x)(

G(x))dx
_

A proof of g e is as follows: Let

G = g
where g is some real number, then


G =

g = g||
2
Thus,
_

(

G)dx = g
_
||
2
dx
We could have also written,
_
(

G

)dx = g

_
||
2
dx
Thus, g

= g by the relation above.


25
2.8 Elementary Solutions
We are now ready to solve some simple problems. We will use these simple problems as building
blocks for more complex problems later on. The following sections will follow a basic form. Well
rst write the Scrodinger equation, give the solution of the resulting dierential equation and the
energies, and then discuss.
2.8.1 Free Particle
Here V = 0 ever where and we have only the kinetic energy operator.
H = T =
h
2
2m

2
=
h
2
2m
_

2
x
2
+

2
y
2
+

2
z
2
_
= T
x
+T
y
+T
z
Since the Hamiltonian is separable, the wavefunction must factor into separate components: (x, y, z) =

x
(x)
y
(y)
z
(z) and the energy is the sum of the kinetic energy in each direction, E = E
x
+E
y
+E
z
.
We have then three identical Schrodinger equations:
T
x

x
(x) = E
x

x
(x) (50)
T
y

y
(y) = E
y

y
(y) (51)
T
z

z
(z) = E
z

z
(z) (52)
corresponding the the dierential equations:

h
2
2m

2
x
2

x
(x) = E
x

x
(x) (53)

h
2
2m

2
y
2

y
(y) = E
y

y
(y) (54)

h
2
2m

2
z
2

z
(z) = E
z

z
(z) (55)
They are all the same...so pick one. Lets take the one in x.

(x) =
2mE
x
h
2
(x) = k
2
(x)
where k =
_
2mE/h
2
has units of inverse length and is referred to as a wavenumber.
Weve encountered solution to this dierential equation already and you can easily show that they
must have the form
(x) = Acos(kx) +Bsin(kx)
where A and B are to be determined by the boundary conditions. We can also write a more general
solution as
(x) = Ae
ikx
+Be
ikx
26
A and B are dierent constants than before. These we must determine by imposing boundary
conditions on the problem. Lacking boundary conditions, k can take any value and is a continuous
variable. Thus, we can say that lacking boundary conditions restricting the value of k,
E
x
=
k
2
x
h
2
2m
Hence, for three dimensions,
E =
h
2
2m
(k
2
x
+k
2
y
+k
2
z
)
We can pick A and B to be anything we wish, so lets take B = 0 and let

x
(x) = Ae
ikx
(Ill neglect the subscript on the k
x
when its in the exponential for clarity. ) Now, take the
momentum expectation value in x:
p
x
=
h
i
_

x
(x)dx
_
|
x
|
2
dx
Putting in the various functions;
p
x
=
h
i
_
e
ikx
(ik
x
)e
+ikx
dx
_
dx
= hk
x
Thus, the wavefunction e
ikx
corresponds to a momentum eigenstate with the particle moving in
the +x direction. The other case then corresponds to a particle moving towards the x direction
with momentum hk
x
.
2.8.2 Particle in a well in 1, 2, and 3 dimensions
Now let us impose boundary conditions by placing our particle inside a conning potential well
with V (x) = 0 or x [0, L] and V = otherwise. This imposes the condition that at x = 0 and
x = L, (x) = 0 otherwise the system would have innite energy. Inside the box, the Schrodinger
equation is the same as above, and have the solution
E =
h
2
k
2
2m
and
(x) = Ae
ikx
+Be
ikx
.
Now, apply the boundary conditions at x = 0
(0) = A+B =
thus, A = B. Also,
(L) = A(e
ikL
e
ikL
) = A

sin(kL) = 0
27
Since sin(kL) = 0 whenever kL = n where n = 0, 1, 2, we have a quantization condition:
k =
n
L
where n is any integer. We have to exclude the case where n = 0 since is would force sin(0x) = 0
everywhere, and our particle must be somewhere. Thus, we have the discrete energy levels:
E
n
=
h
2
n
2

2
2mL
2
For the normalized eigenfunctions, we require
1 = A
2
_
L
0
sin
2
(x
n
L
)dx = A
2
L
4
_
2
sin(2n)
n
_
Since n = 1, 2, 3 , the sin(2n) = 0 n. Thus, A =
_
L/2 and we have

n
(x) =
_
2
L
sin(nx/L)
as our eigenstates.
Let us rst check that these are in fact the eigenstates. Inserting
n
into the Schrodinger equation
above: n

n
=
n
2

2
L
2

n
=
2mE
n
h
2

n
.
Now lets check orthogonality:
_
L
0

m
(x)
n
(x)dx =
2
L
_
L
0
sin(mx/L) sin(nx/L)dx
Evaluating the integral:
2
L
_
L
0
sin(mx/L) sin(nx/L)dx =
2ncos(n) sin(m) 2mcos(m) sin(n)
(m
2
n
2
)
The numerator is equal to 0 in all cases since n, m 1, 2, . If m = n, then we have the integral
we had above, which is normalized to 1. Thus, we conclude:
_
L
0

m
(x)
n
(x)dx =
nm
where
nm
is the Kronecker-delta function which is equal to 0 if m = n and 1 if m = n.
Now, lets evaluate some simple expectation values:
The rst two we do not even need to do explicitly (can you see why??)
x =
_
L
0

m
(x)x
n
(x)dx =
L
2
28
p =
_
L
0

m
(x)x
n
(x)dx = 0
This one we have also already done.
p
2
=
_
L
0

m
(x)p
2

n
(x)dx =
h
2
n
2

2
L
2
OK, so this one we actually have to evaluate:
x
2
=
_
L
0

m
(x)x
2

n
(x)dx =
L
2
_
4
3
n
3
6 cos(2n)n +
_
3 6n
2

2
_
sin(2n)
_
12n
3

3
Taking n to be integer
x
2
=
1
6
L
2
_
2
3
n
2

2
_
Now, consider the r.m.s deviation in position and momentum.
_
x
2
=
_
x
2
x
2
=
1
12
L
2
_
1
6
n
2

2
_
_
p
2
=
_
p
2
p
2
=
hn
L
Taking the product:
_
x
2
p
2
=
h
2
_
1
3
_
n
2

2
6
_
_
2
Now, taking n 1, the minimum uncertainty in the product of x and p is
_
x
2
p
2
=
h
2
_
1
3
(
2
3) = 1.28987
h
2
A plot of this is shown in Fig. 6
This idea of a minimum uncertainty is a general result and unique to quantum mechanics. In a
more general sense, it is related to the Schwartz inequality
x
2
p
2
([x, p])
2

Since [x, p] = ih, we can conclude that


x
2
p
2
([x, p])
2
h/2
This has a profound implication in that it means that one can not simultaneously determine both the
position and the momentum of a particle to arbitrary precision. If you have a precise determination
of the momentum, you can make no precise determination of the position. Likewise, if you localize
the particle to some nite region of space, then there is a nite degree to which you can determine
the momentum (hence) kinetic energy of a particle. One can conclude, that any bound quantum
29
Figure 6: (x)
2
(p)
2
for 1 d particle in a box in units of h. The dashed line corresponds to h/2.
mechanical system can never be exactly at rest at the bottom of the potential well. The particle
must remain in some motion...else the uncertainty principle would be in violation.
2 and 3 dimensional wells:
We can extend the results from the last section to arbitrary numbers of dimensions. When we do,
we have the possibility that for a given set of dierent quantum numbers we can have states with
the same energy.
For a two dimensional system:
E(n
x
, n
y
) =
h
2

2
2m
_
n
2
x
L
2
x
+
n
2
y
L
2
y
_
and for 3 dimensions:
E(n
x
, n
y
, n
z
) =
h
2

2
2m
_
n
2
x
L
2
x
+
n
2
y
L
2
y
+
n
2
z
L
2
z
_
30
Lets take the case where L
x
= L
y
= L
z
for a 3D box. The energies for the rst 5 levels are listed
in Table 3. Eigenfunctions with identical energies are degenerate. The functions themselves are
Table 3: Energies for 3d Box
quantum numbers energy in h
2

2
/2mL
2
(2,2,2) 3 2
2
= 12
(1,2,2),(2,1,2),(2,2,1) 1
2
+ 2
2
+ 2
2
= 9
(1,1,2),(1,2,1),(2,1,1) 1
2
+ 1
2
+ 2
2
= 6
(1,1,1) 1
2
+ 1
2
+ 1
2
= 3
dierent and are still orthogonal to each other.
Quantum wells: The idea that the spacing and degeneracy of quantum states can controlled by
the size of the system is the basis of a number of nano-technological devices. See the on-line
encyclopedia Wikipedia for extensive details and links.
2.9 Harmonic Oscillator
We shall not derive the solutions to the time-independent Schrodinger equation in detail in this
case. The potential in this case is the parabola:
V (x) = kx
2
/2
This corresponds to a Hookes law restoring force for a displaced oscillator. In molecular terms,
we can represent the vibrations of bonds within a molecule in terms of a potential energy function
derived from the electronic degrees of freedom. Well discuss this in more detail later on in this
course. However, you should know that for a diatomic species, as you displace the system from
its equilibrium bond length, its energy will increase. The sketch below shows a typical form of the
potential as a function of bond-length r. If we expand the potential about the minimum, r
e
, which
is the equilibrium bond-length and set our zero of energy to be the bottom of the well, then we can
expand the potential in a Taylor series expansion as
V (r)
1
2
_
d
2
V
dr
2
_
r=re
(r r
e
)
2
+
ignoring the higher order terms. The rst derivative term vanishes since we are at a minimum.
The second derivative is the curvature at the bottom of the well, k. This is our Hookes law force
constant with the classical oscillator frequency =
_
k/. Thus, our Schrodinger equation is
_

h
2
2

2
r
2
+
k
2
(r r
2
)
2
_
= E
Note that the mass here is the reduced mass. This comes about when we consider the motion of
the two atoms in a center of mass frame. (See page on how to derive reduced mass equations.)
31
Finally, we dene a distortion coordinate x as r r
e
as the displacement from equilibrium. Since
dx = dr, the Schrodinger equation becomes
_

h
2
2

2
x
2
+
k
2
x
2
_
(x) = E(x)
The eigenfunction solutions of this equation are
e

y
2
2
_
2
n
n!
H
n
(y)
4

(56)
with y = x/a and a =
_
h/. H
n
(y) is a Hermite polynomial. A brief table of these polynomials is
given below. They can be generated using the Mathematica command: HermiteH[n,z], or generated
by taking the derivatives
H
n
(x) = e
x
2 d
n
dx
n
e
x
2
Note, that in contrast to the particle in a box case above, the lowest quantum number is n = 0.
32
Table 4: Hermite Polynomials, H
n
(z), up to n = 5.
n H
n
(z)
0 1
1 2z
2 4z
2
2
3 8z
3
12z
4 16z
4
48z
2
+ 12
5 32z
5
160z
3
+ 120z
The eigenvalues are then E
n
= h(n + 1/2). Hence,
H
n
= h(n + 1/2)
n
for the harmonic oscillator.
Sketches of the wavefunction solutions are shown in Fig. 7
2.10 The Principle of Superposition
Given a complete set of eigenfunctions, I can generate any arbitrary wavefunction by taking a linear
combination of the eigenfunctions:
(x) =

n
c
n

n
(x)
where
c
n
=
_

n
(x)(x)dx
is the probability amplitude of nding the system in the
n
eigenstate. In other words, we can
imagine populating the dierent eigenstates of a system with dierent probabilities of nding the
system in that particular state.
If we have two states populated, we can write
(x) = c
1

1
(x) +c
2

2
(x)
For example, if we take the rst two harmonic oscillator states and require the two coecients be
such that |c
1
|
2
+|c
2
|
2
= 1 we can have the following states:

=
1

2
(
0
(x)
1
(x))
These are shown in Fig. 8
Notice, that by combining the two waves, we can create new wavefunctions with the centers shifted
to one side or the other. These new wavefunctions are no longer stationary solutions of the SE.
33
They will evolve in time. One can visualize the time-evolution on the Java applet on the course
web-site.
Lets consider the time-evolution of the superposition of two harmonic osc. eigenstates under the
harmonic osc. hamiltonian. The formal solution of the time-dependent SE is
(t) = e
iHt/h
(0)
where (0) =

. The exponentiated Hamiltonian can be evaluated in a variety of ways. Perhaps


the most straightforward is to expand the exponential in a Taylor series:
(t) =
_
1
it
h
H +
1
2
_
it
h
_
2
H
2
+
_

Since H
n
= E
n

n
we can write
_
1
it
h
H +
1
2
_
it
h
_
2
H
2
+
_

n
=
_
1
it
h
E
n
+
1
2
_
it
h
_
2
E
2
n
+
_

n
= e
itEn/h

n
Thus,
(t) =
1

2
_
e
itEo/h

0
e
itE
1
/h

1
_
Since E
n
= h(n + 1/2)
(t) =
1

2
e
it/2
_

0
e
it

1
_
where I have pulled out a common factor of e
it/2
. Thats the time evolution of the wave function.
Lets ask about the time-evolution of various physical quantities.
1. Energy
E(t) =
_

(t)H(t)dx =
1
2
__

0
H
0
dx +
_

1
H
1
dx e
it
_

0
H
1
dx e
+it
_

1
H
0
dx
_
Since
n
are eigenfunctions of H and are orthogonal, the last two integrals are exactly equal
to 0 and we have
E(t) =
1
2
( h/2 + 3h/2) = h
which is independent of time.
2. Avg. Position For this we write
x(t) =
_

(t)x(t)dx (57)
=
1
2
(e
it
+e
+it
)
_

0
(x)x
1
(x)dx (58)
= cos(t)
_

0
(x)x
1
(x)dx (59)
The integral we do need to evaluate.
_

0
(x)x
1
(x)dx =

h
2
34
Thus,
x(t) =

h
2
cos(t)
2.11 Tunneling
Another intrinsically quantum mechanical phenomenon is that a particle can penetrate through a
potential barrier, even though it has less energy than the barrier itself.
Consider the case shown in Fig. ?? where V is the potential which is non-zero only over a small
region 0 x a. Imagine we have a particle with momentum p = hk impinging upon the barrier
from the left. In each of the three regions, the wavefunction has the following functional forms:
Region I: x < 0 Here,

I
(x) = Ae
ikx
+Be
ikx
where the rst term is the incoming component and the other is the reected component.
Region II: 0 xa

II
(x) = Ce
x
+De
+x
where =
_
2m(E V
o
)/h. Since we will not want exponentially growing solutions, we take
D = 0.
Region III x > a

III
= Ee
ikx
we have only out-going components.
Now, we need to match boundary conditions that at each boundary. Recall that (x) must be
continuous, as well as

(x). This, at x = 0

I
(0) =
II
(0) (60)
(61)
2.11.1 Angular momentum
2.12 Dirac NotationOptional
Often times in quantum mechanics we adopt a short hand notation for the wave functions and
operators by considering them as generalized functions (or rather vectors) in an abstract functional
space. This allows us to make a rather important connection between working with functions and
dierential operators and working with vectors and matrices. Paul Dirac made this important
35
connection in the development of quantum theory. The notation may take some getting used to,
but it is extremely powerful and will greatly simplify calculations.
This idea is that | represents the quantum state. The | is termed a ket and integrals over
wavefunctions are formed by forming a bra-ket.
|O| =
_

(x)O(x)dx.
The wave function is the projection of | onto the real x-axis:
(x) = x|
Other properties:
| = (|)

where denotes the Hermitian conjugate. Likewise,


|x =

(x)
2.12.1 Using Bra-ket notation
The Dirac or bra-ket notation is very powerful in setting up calculations. For example, the proof
we had above is simply if
G| = g
and
|G = g

|
then
|G| = g

| = g|
Now suppose, G| = a| and G| = b| where a = b. Thus,
|G| = b| = a|
Thus, either a = b or | = 0. Hence, the states | and | are orthogonal. In fact, all eigenstates
of the same operator are orthogonal.
| =
_

dx = 0
I will use the Dirac bra-ket notation when needed.
One can easily build up calculations using the Dirac notation using a few simple tools. For example,
I can represent the identity operator (multiplication by 1) in the following way:
If |
n
is an eigenvector of operator H, H|
n
= E
n
|
n
then if |
n
is a normalized states and

m
|
n
=
nm
, then I can dene the identity operator as
I =

n
|
n

n
|
36
where the sum is over all the eigenvectors of H. If we have a continuous spectrum of eigenvalues,
then I can write this as an integral
I =
_
dx|xx|.
This is a useful identity, since it allows us to take a statevector in one representation and re-write
it in terms of the eigenstates of a dierent operator. For example, we can write (x) = x| as the
wavefunction of the system. It may or may not be an eigenstate of any operator, but I can write
it in terms of eigenstates by application of the identity operator. Lets write the wavefunction in x
as a function of momentum states, |p using First, we insert the identitiy:
p| = p|I| =
_
dxp|xx|
Then note that p|x = e
ipx
/(2)
1/2
= x|p

is an eigenstate of the momentum operator. Thus,


the momentum wavefunction
p| = (p) =
1
(2)
1/2
_
dxe
ipx
(x)dx
is the Fourier transform of the wavefunction in position.
Moreover, if I have the eigenstates of some operator, I can expand in terms of these states by
| =

n
|
n

n
| =

n
c
n
|n
where c
n
=
n
| are the expansion coecients computed by taking the integral
c
n
=
_

n
(x)

(x)dx
37
Figure 7: Harmonic Oscillator Wavefunctions for n = 0, 1, 2 along with the corresponding proba-
bility densities
38
Figure 8:

=
1

2
(
0
(x)
1
(x))
39
Figure 9: Java applet showing time-evolution of harmonic osc. superposition states
Figure 10: A potential step barrier
40
Figure 11: Brief derivation of reduced mass equations
41

Вам также может понравиться