Anda di halaman 1dari 16

Quantum Solutions For A Harmonic Oscillator

Particle in a Box Analogy:

V(x) = k x2/2
Quantization -

Zero Point Energy -

Nodes -

Symmetry of ψ -

The Schrödinger equation for a particle of mass μ is


d 2 ψ (x)
2
kx2
− + ψ (x) = E ψ (x) ,
2μ dx2 2

which looks a lot like the first homework eigenvalue problem

52
ACTUALLY,
En = (k/m)1/2(n + ½ ) n = 0,1,2,3, . . .
β x2
Ψ n (x) = N n H n (β x) e −
2

where β2 = μk/ 2, and the Hn are Hermite polynomials. (See text)


H0 = 1 even
H1 = 2 βx odd
H2 = 4(βx)2 - 2 even
H3 = 8(βx)3 - 12βx odd

What do the solutions look like?

Again, to get a sense of the wavefunctions and the time dependence, look at the falstad website
http://www.falstad.com/qm1d/ and select “Setup: Harmonic Oscillator” from the drop down
menu.

53
What about various integrals of the harmonic oscillator functions?

ψ i ψ j = δ ij (WHY??)

where δij = 1 for i = j


0 for i ≠ j the Kronecker delta

what about x = ψ i x ψ i (WHY??)

Note that both the particle in a box and the harmonic oscillator eigenfunctions have a
definite parity and that it alternates with the quantum number. This result has to do with the
symmetry of V(x).
We can generalize this result with operators. Let the parity operator  be such that
Âf(x) = f(-x)

Then Â2 f(x) = ÂÂf(x) = Âf(-x) = f(x) = 1 f(x)


Since any function f(x) is an eigenfunction of Â2 with eigenvalue 1, the only possible
eigenvalues of  are ±1.

Suppose f(x) is odd about x = 0. Then Âf(x) = (-1) f(x)

Suppose g(x) is even about x = 0. Then Âg(x) = (1) g(x)

What is f g ? WHY??

54
Three Dimensional Harmonic Oscillator
Now let's quickly add dimensions to the problem. Consider a three dimensional harmonic
oscillator for a particle of mass m with different force constants kx, ky, and kz in the x, y and z
2 2 2
kx x ky y kz z
directions. The potential energy is V(x, y, z) = + +
2 2 2
and the Hamiltonian is given by
∂ 2 − 2 ∂ 2 − 2 ∂ 2 + kxx 2 + k yy + kzz2
2 2
Hˆ = − .
2 m ∂ x 2 2 m ∂ y2 2 m ∂ z 2 2 2 2
This Hamiltonian is clearly separable into x-, y- and z- equations, each of which looks
just like the one we have already solved. Thus the solutions are
kx 1 ky F1 I
kz 1 F I F I
En , n , n =
x
m
nx +
y z
2
+
m H
ny + +
2 mK nz +
2 H K H K
with n x = 0,1,2,.. n y = 0,1,2,.. n z = 0,1,2,...

and ψ n (x, y, z) = N n Hn (x) e−β x N n Hn (y) e−β y N n Hn (z) e−β z


2 2 2
x y z
x ny nz x x y y z z

It's really that simple! Now consider the isotropic 3-D oscillator with all three force constants
k
equal. Then kx=ky=kz=k and we have V(x, y, z) = x2 + y + z2 =
2
2 k 2
2
r . Here c h
r = x 2 + y2 + z 2 , the distance from the origin. The result is
k F n + n + n + 3I F 3 I
En , n , n =
x y z
m H x y
2K
z
H
= h ν class n x + n y + n z +
2 K
with n x = 0,1,2,. . n y = 0,1,2,.. n z = 0,1,2,...

Now we see new energy degeneracies since the problem has high symmetry.
Consider the level E= (9/2) hνclass . Then (nx + ny + nz +3/2) = 9/2
or nx + ny + nz = 3 . The possibilities are as follows:

55
nx ny nz

3 0 0
2 1 0
2 0 1
1 2 0
1 0 2
1 1 1
0 2 1
0 1 2
0 3 0
0 0 3

The level is 10 fold degenerate!


The potential V(r) is a very special potential, acting in three-dimensional space in a
direction along the line from r to the origin. If we imagine placing a proton at the origin,
and allow an electron to move around it, then we have the Coulomb potential we have seen
earlier, V(r) = -e2/r. This is a special case because the proton is so heavy that the center of
mass is essentially located on the proton, at the origin. We will quickly wave our hands
through a generalization to obtain center of mass coordinates.
Potentials of the form V(r) are called Central Potentials, and are of great importance in
quantum mechanics. We will solve all central potential problems as a simple equivalent
one-dimensional problem and in the process learn about angular momentum.
First, we deal with the separation of center-of-mass motion from internal motion.

56
Separation Of Center Of Mass Motion

Consider as a concrete example an ideal spring with masses m1 and m2 at the ends of the
spring, and constrained to move only in the z-direction. Let 0 be the unstretched (equilibrium)
length of the spring. We have the following picture.
m1 m
2

z
z1 z2

The potential energy of the spring can be written in terms of z1 and z2 as


a f 2
V( z1 , z2 ) = 12 k z2 − z1 − 0 and the Hamiltonian is given by
pz21 pz22 1
+ k ⎡⎣( z2 − z1 ) − 0 ⎤⎦
2
ˆ
H= +
2m1 2m 2 2
This Hamiltonian is certainly not separable in z1 and z2 coordinates. However, there is a
change of variables to z and Z, one that makes the expression separable. The quantity z is an
internal distance (z2-z1) while Z is the location of the center of mass of the system,
m1 z1 + m2 z2
Z =
m1 + m 2
After a little algebra, we can show that

57
2 2
pz k pZ
H ( z1, z 2 ) = H(z, Z) = + (z − 0) 2 +
2μ 2 2( m1 + m 2 )

where μ = m1 m 2 the REDUCED MASS


m1 + m 2
so the Hamiltonian is separable into motion of the center of mass (COM) and internal motion.
d + k (z − 0) −
2 2 2 2 2
i.e. Hˆ = − d
2 μ dz 2 2 2 ( m1 + m 2 ) dZ2
and the wave function ψ(z, Z) can be written in product form,
ψ(z,Z) = φ(z) fcm(Z)
where f(Z) describes the free (unquantized) motion of the center of mass in Z and φ(z) is the
solution to the other part. Now E = EZ + Ez.
In effect, the one problem separates into two independent problems:
1. The motion of a fictitious particle of mass (m1+m2) positioned at the center of mass,
and subject only to the external forces acting on the system.
2. The motion of a fictitious particle of reduced mass μ, subject only to the internal
forces. The center of mass of this system does not move.
Note that if we put this system in a box, then the Z motion would be quantized. This should
answer your worry about where the second quantum number went, since there are two particles
in one dimension.
So the harmonic oscillator problem becomes what we have seen before,


d φ(z)
22

+
k z− a f φ(z)
0
2

= E φ(z)
2μ dz2 2
with the known solutions, now simply centered about 0 rather than zero. This same procedure
will always work in 3 dimensions, and the motion of the center of mass reduces by three the
number of internal quantum numbers. These center of mass motions when the "particle" is
confined to a box give rise to the translational heat capacity expression we will derive in the
statistical mechanic portion of the course.
We will always implicitly separate out the center of mass motion in the future.
Soon we will see that when the 3 dimensional potential V(x,y,z) depends only on
the distance from the origin, i.e. V(x,y,z) = V(r), a central potential, then the Hamiltonian will
be separable in spherical polar coordinates, (r, θ, φ). The θ and φ parts of the solution will be
the same for any radial potential: coulomb, harmonic, rigid rotor or whatever!!
The θ and φ parts will refer to the angular momentum properties of the system, no
matter what system V(r) describes. For example, if V(r) is the 1/r Coulomb potential (H atom),
then the quantum numbers arising from the θ and φ parts give rise to the and m quantum
numbers we know well.
Before that time, however we make a short foray into commutation relations and the
basics of angular momentum.

58
Commutation Relationships

We have seen something of commutation relations. We need to say a little more about
why they are important.
First remember that if a state function is an eigenfunction of operator Â, then every
measurement of the A property gives the eigenvalue ai. This means that the property A is an
exact description of that aspect of the system. (energy, angular momentum, whatever) We wish
to find as many sharp properties of the system as possible.
In our 3D harmonic oscillator we saw that the function
-βx2/2 -βy2/2 -βz2/2
Ψnx,ny.nz(x,y,z) = Nn Hn(βx)e Hn(βy)e Hn(βz)e
is simultaneously an eigenfunction of [KEx, + V(x)] , [KEy +V(y)] and [KEz +V(z)]. The
harmonic oscillator Hamiltonian commutes with all of these operators. The properties are all
"sharp", i.e., characterized by zero standard deviation in a measurement. We will always want
to utilize such sharp such properties to characterize a system.

For an H atom, they might be total energy, total angular momentum and one component
of angular momentum. How do we identify such properties?

Theorem 1: If  commutes with Ô ( i.e., [Ô,Â] = 0), then there exists a complete set of
functions which are simultaneously eigenfunctions of Ô and Â.

Theorem 2: If two operators Ô and  have a common, complete set of eigenfunctions, then
they commute, i.e. [Ô,Â] = 0.

The proof follows from the fact that any function can be expanded in the common,
complete set of eigenfunctions.
Theorem 3: If Hermitian operator  commutes with Ô, and if φ1 and φ2 are eigenfunctions of Â
with eigenvalues a1 and a2 , then the MATRIX ELEMENT O1,2 vanishes unless a1 = a2..
Here O1,2 = φ1 O φ 2 , the definition of a matrix element.
Proof:
First Âφ1 = a1φ1 and Âφ2 = a2φ2
Now φ1 OA φ 2 = a2 φ1 O φ 2 .
Also φ1 OA φ 2 = φ1 AO φ 2 since they commute
φ1 AO φ 2 = Aφ1 O φ 2 = a1 φ1 O φ 2
Thus a2 φ1 O φ 2 = a1 φ1 O φ 2
or (a2-a1 ) φ1 O φ 2 = 0.

59
Since a1 ≠ a2, the matrix element must vanish. This theorem will be extremely useful in
applying symmetry to assist in obtaining wavefunctions. It also begins to show the importance
of matrix elements in quantum mechanics.
As a follow up, consider the harmonic oscillator problem
2 2 2
Hˆ = − d + kx
2 μ dx 2 2
with state functions ψn (x) as we know.
Now Hij ≡ ψ i Hˆ ψ j

Central Field Potential


Consider the case where there are two particles of mass m1 and m2 interacting via a
potential V(r), which depends only on the distance between the particles. (a central field) The
Hamiltonian operator is given by
2
H cent = − ∇ + V(r)
2


If we start with rectangular (x,y,z) coordinates, we must transform this expression to spherical
polar coordinates, r, θ, φ where 0 ≤r<∞ , 0≤θ≤π and 0≤φ≤2π. This is easily done as shown:

x = r sinθ cos φ y = r sinθ sin φ z = r cosθ

We can transform the x, y and z derivatives into r, θ, and φ terms. After much
manipulation, sweat, blood and tears, we would obtain the time independent Schrödinger

60
equation for a central field. Alternatively, we look up the form of ∇2 in spherical polar
coordinates.
In either case, we finally obtain the time independent Schrödinger equation for a central
field:

LM 1 ∂ F r ∂ I + 1 ∂ F sin θ ∂ I OPψ(r, θ, φ)
2

2μ N r ∂ r H ∂ r K r sin θ ∂θ H ∂θ K Q
− 2
2
2


2
1
+ 2 2 2 ψ (r, θ, φ ) + V(r)ψ (r, θ, φ ) = E ψ (r, θ, φ )
r sin θ ∂ φ
Various forms of V(r) give us important problems:
V(r) Problem

I. kr2 /2 3D Isotropic Harmonic oscillator

II. 0 for r =R Rigid Rotor


∞ otherwise

III. -ze2/r Hydrogen-like atom


What do we do now?
What we will find out is that all of the angular part of the Schrödinger equation above is really
the square of the angular momentum operator, L . The Schrödinger equation will in this case
simplify to
2
∂ ⎛ 2∂ ⎞ L2 ψ (r,θ , φ ) + V(r)ψ (r,θ , φ ) = Eψ (r,θ , φ )
− ⎜r ⎟ψ (r,θ , φ ) +
2μ r ∂ r ⎝ ∂ r ⎠
2
2μ r 2
When we learn more about angular momentum, we will see that this equation (which has
all the θ and φ mess hidden in the L part) is an enormous simplification.

61
Angular Momentum

Classically the angular momentum L of a particle of mass m is given by L = r × p , a


vector quantity which can be expressed as a determinant,
i j k
L = x y z
px p y pz
= i(y pz − z py ) + j(z px − x pz ) + k (x py − y px )
We of course know that an isolated system has a constant angular momentum vector. We
dp
also remember that the angular momentum version of F = ma = involves torque τ and
dt
is written
dL d Lx d Ly d Lz
τ= . If τ =0 , then dL = 0 ⇒ = = = 0
dt dt dt dt dt

Thus L , Lx, Ly, and Lz are constant in this case. They are "Constants of the Motion" Think
about tops or gyroscopes!
We shall see that quantum mechanically only the square of the total angular
momentum and one component can be constants of the motion. That is to say, they have
commuting operators. Thus we will look at the QM operators, their commutation relations
and their representation in spherical polar coordinates.
Looking at the quantum mechanical operator,
i j k
Lop = x y z = Lx i + L y j + Lz k
pˆ x pˆ y pˆ z
where
⎡ ∂ ∂ ⎤
Lx = − i ⎢ y −z
⎣ ∂z ∂ y ⎥⎦
⎡ ∂ ∂ ⎤
Ly = − i ⎢ z −x ⎥
⎣ ∂x ∂z⎦
⎡ ∂ ∂ ⎤
Lz = − i ⎢ x −y
⎣ ∂y ∂ x ⎥⎦
Also,
L2 = L ⋅ L = L2x + L2y + L2z

62
Let us look at commutation relations, e.g., ⎡⎣ L x , L y ⎤⎦
⎛ ∂f ∂f ⎞
First L x L y f = L x ⎜ − i ⎡⎢ z − x ⎤⎥ ⎟
⎝ ⎣ ∂x ∂z⎦ ⎠
Then evaluate L y L x f and subtract. After a bunch of algebra, we obtain the commutator
⎡Lx , L y ⎤ = i Lz
⎣ ⎦
By cyclic permutation of variables (x→y, y→z, z→x), we also get
⎡Lx , L y ⎤ = i Lz
⎣ ⎦
⎡L y , Lz ⎤ = i Lx
⎣ ⎦
⎡Lz , Lx ⎤ = i Ly
⎣ ⎦
Since none of these commutators are zero, we have that
No two components of angular momentum commute.
Thus we cannot specify two components of angular momentum at the same time!
We also find that [ L2 , L x ] = [ L2 , L y ] = [ L2 , L z ] = 0

So we can find functions that are simultaneously eigenfunctions of L2 and any one component
of the angular momentum.
We switch to spherical polar coordinates (r, θ, φ) and find that we can write the operators L2
and L z in the following form:

⎡ 1 ∂ ⎛ ∂ ⎞ 1 ∂2 ⎤
Li L ≡ L = − ⎢ ⎜ sin θ
2 2
⎟+ 2 2⎥
⎣ sin θ ∂θ ⎝ ∂θ ⎠ sin θ ∂φ ⎦


Lz = − i
∂φ

L x and L y are given in the text, and are more complex.

63
So finally we obtain
2
[ L , Lx ] = 0 [ H cent , L y ] = 0 [ Lx , Ly ] = i Lz

[ L2 , L y ] = 0 [ H cent , L z ] = 0 [Ly, L z ] = i L x
[ L2 , L z ] = 0 [ H cent , L x ] = 0
[ L z , L x ] = i Ly
[ H cent , L ] = 0
2

So we can find eigenfunctions of the central field Hamiltonian, H cent , that are
simultaneously eigenfunctions of H cent , L2 and one component of the total angular momentum.
Conventionally, and for mathematical simplicity this is taken to be the unique axis (z). Thus
we add L z to our list of eigenfunctions. Since there are only three possible quantum numbers
for this system (WHY??), our list is complete. We now go on our way about solving the three
eigenvalue equations for L z , L2 and H cent .
Note that our solutions for the first two must be valid for any central field problem.

I. Z - Component
We can readily solve the L z eigenvalue equation,
∂Φ
L z Φ (φ ) = − i = CΦ (φ ) ,
∂φ
to obtain Φ( φ ) = e imφ for 0 ≤ φ ≤ 2π with the eigenvalues C being m .
In order for Φ(φ) to be single valued, we must satisfy the boundary condition Φ(0) =
Φ(2π). This restriction requires that m = 0, ±1, ±2, . . . . Thus the z-component of total angular
momentum is quantized. This quantum number will turn out to be the m quantum number we
have seen before in the hydrogen atom.
Show that the corresponding eigenvalues are m . Note that this must mean that has
units of angular momentum.

64
II. Square of Total Angular Momentum
The solution to the L2 eigenvalue equation is not so straightforward, and we will not go
through it. However, the solution is a standard function of mathematical physics called the
spherical harmonics, Y ,m(θ,φ). These are tabulated in the text (and in any standard quantum
mechanics text). The key properties of these functions are summarized by their eigenvalues:
2
L2 Y ,m(θ,φ)= ( )( +1) Y ,m(θ,φ) = 0, 1, 2 , . . .
and
L z Y ,m(θ,φ)= m Y ,m(θ,φ) m= - , - +1 , . . . , 0 , . . . ,

What is the physical significance of this result??


Now L2 and L z must both be sharp quantities. That obviously means that L must also be exact.
With a little more thought, we see that L must be precessing about Lz and the m-quantum

Lz
L

number represents a spatial quantization. °


As an example, consider = 1. Then L = 2 and there are three possible m-values,
m= 1 m= 0 m = -1

z z z

√2 √2
√2

65
Some of the spherical harmonics are listed below:
2 1
Y0,0 (θ, φ ) =
4 2π
3 1
Y1,0 (θ, φ ) = cos θ
2 2π
3 1 ± iφ
Y1, ±1 (θ, φ ) = sin θ e
4 2π

Y2,0 (θ, φ ) =
5
8
c
3 cos2 θ − 1 h 1

15 1 ± iφ
Y2, ±1 (θ, φ ) = sin θ cos θ e
4 2π

15 2 1 ±2iφ
Y2, ±2 (θ, φ ) = sin θ e
16 2π
What about nodes in these functions?
III. Radial Part
Now back to the central potential problem. We had previously obtained the Schrödinger
equation for a central force as
2
∂ ⎛ 2∂ ⎞ L2
− ⎜r ⎟ψ (r,θ , φ ) + ψ (r,θ , φ ) + V(r)ψ (r,θ , φ ) = Eψ (r,θ , φ )
2μ r 2 ∂ r ⎝ ∂ r ⎠ 2μ r 2
Let a solution be of the form ψn, ,m(r,θ,φ) = Rn, (r)Y ,m(θ,φ).
When we make this substitution, we can explicitly carry out the angular momentum operation,
and bring the Y ,m(θ,φ) terms through the r - derivatives and the V(r) Rn, (r) term through the
angular momentum operator. We then divide through by Y ,m(θ,φ).
The Rn, (r) must satisfy the following ordinary differential equation :
2
FG
d 2 dR n, IJ + LMV(r) + b + 1g OP R
2

2μr dr
2 r
dr H K N 2μr Q 2 n, (r) = E R n , ( r )

expanding this expression,


2
FG 2 dR 2
d R n, IJ LM 2
( + 1) OP
2μ H r dr K N
n,
+
dr 2
+ V(r) +
2μr 2 Q
R n, (r) = E R n , ( r )

As this is a one-dimensional equation, we can readily obtain a numerical solution. Thus


the central field problem is a numerically solvable system.
If we consider the nucleus of an atom to be infinitely heavy, and neglect the electron-
electron repulsion, we shall see that the n-electron atom problem separates and becomes n - one

66
electron atom problems. This is the starting point for ANY system, and is the reason that the
solutions of the hydrogen atom problem play such a key role in our understanding of atomic and
molecular structure!
Rigid Rotor Energies and Wavefunctions
The rigid rotor consists of a pair of particles of masses M1 and M2 (reduced mass μ)
separated by a fixed distance Ro. With this model, we can set the central potential to be V(r) =
0 for R =Ro, and infinite otherwise. The R(r) derivatives in the preceding equation can be can
be shown to be zero, basically for the same reasons as for the infinite square well solved earlier.
Thus at r = Ro, the radial Schrödinger equation becomes
2
⎛ ( + 1) ⎞
⎜ ⎟ Y ,m (θ , φ ) = EY ,m (θ ,φ )
2 μ ⎝ Ro2 ⎠
By convention, the quantum number in the rotor problem is replaced by j. Thus the allowed
energies are given by
2 2
E j,m = j(j+ 1) = j(j+ 1) = B j(j+ 1)
2μ R 2o 2I

with j = 0,1, 2,… and m = − j,− j+ 1,…−1, 0,1… j

where ψ (r, θ, φ ) = Y j,m ( θ, φ )


The quantity I is the classical moment of inertia, I = μ Ro2.
The quantity B is known as the rotational constant.
Clearly, if we can measure the positions of the rotational energy levels, we can then derive the
length of a bond. This is the essence of rotational spectroscopy.
Notice that Yj,m Yj',m ' = δ j, j'δ m ,m '
Note the 2j +1 energy degeneracy of level Ej.
What is the significance of the m quantum number and what is the nature of the
degeneracy associated with it?

67

Anda mungkin juga menyukai