You are on page 1of 5

Harmonic Oscillator in Quantum Mechanics

Given the potential energy in Equation Chapter 5.9Chapter 5.9, we can write down
the Schrödinger equation for the one-dimensional harmonic oscillator:

−ℏ22mψ′′(x)+12kx2ψ(x)=Eψ(x)(Chapter 5.10)(Chapter 5.10)−ℏ22mψ″(x)


+12kx2ψ(x)=Eψ(x)

For the first time we encounter a differential equation with non-constant coefficients,


which is a much greater challenge to solve. We can combine the constants in
Equation Chapter 5.10Chapter 5.10 to two parameters

α2=mkℏ2(Chapter 5.11)(Chapter 5.11)α2=mkℏ2

and

λ=2mEℏ2α(Chapter 5.12)(Chapter 5.12)λ=2mEℏ2α

and redefine the independent variable as

ξ=α1/2x(Chapter 5.13)(Chapter 5.13)ξ=α1/2x

This reduces the Schrödinger equation to

ψ′′(ξ)+(λ−ξ2)ψ(ξ)=0(Chapter 5.14)(Chapter 5.14)ψ″(ξ)+(λ−ξ2)ψ(ξ)=0

The range of the variable xx (also ξξ) must be taken from −∞−∞ to +∞+∞, there


being no finite cutoff as in the case of the particle in a box. A useful first step is to
determine the asymptotic solution to Equation Chapter 5.13Chapter 5.13, that is, the
form of ψ(ξ)ψ(ξ) as ξ→±∞ξ→±∞. For sufficiently large values of |ξ||
ξ|, ξ2≫λξ2≫λ and the differential equation is approximated by

ψ′′(ξ)−ξ2ψ(ξ)≈0(Chapter 5.15)(Chapter 5.15)ψ″(ξ)−ξ2ψ(ξ)≈0

This suggests the following manipulation:

(d2dξ2−ξ2)ψ(ξ)≈(ddξ−ξ)(ddξ+ξ)ψ(ξ)≈0(Chapter 5.16)(Chapter 5.16)


(d2dξ2−ξ2)ψ(ξ)≈(ddξ−ξ)(ddξ+ξ)ψ(ξ)≈0

The first-order differential equation


ψ′(ξ)+ξψ(ξ)=0(Chapter 5.17)(Chapter 5.17)ψ′(ξ)+ξψ(ξ)=0

can be solved exactly to give

ψ(ξ)=const.e−ξ2/2(Chapter 5.18)(Chapter 5.18)ψ(ξ)=const.e−ξ2/2

Remarkably, this turns out to be an exact solution of the Schrödinger equation


(Equation Chapter 5.14Chapter 5.14) with λ=1λ=1. Using Equation Chapter
5.12Chapter 5.12, this corresponds to an energy

E=λℏ2α2m=12ℏkm−−−√=12ℏω(Chapter 5.19)(Chapter
5.19)E=λℏ2α2m=12ℏkm=12ℏω

where ωω is the natural frequency of the oscillator according to classical mechanics.


The function in Equation Chapter 5.18Chapter 5.18 has the form of a Gaussian, the
bell-shaped curve so beloved in the social sciences. The function has no nodes, which
leads us to conclude that this represents the ground state of the system.The ground
state is usually designated with the quantum number n=0n=0 (the particle in a box is
a exception, with n=1n=1 labeling the ground state). Reverting to the original
variable xx, we write

ψ0(x)=conste−αx2/2(Chapter 5.20)(Chapter 5.20)ψ0(x)=conste−αx2/2

with

α=(mk/ℏ2)1/2(Chapter 5.21)(Chapter 5.21)α=(mk/ℏ2)1/2

With help of the well-known definite integral (Laplace 1778)

∫∞−∞e−αx2dx=πα−−√(Chapter 5.22)(Chapter 5.22)∫−∞∞e−αx2dx=πα

we find the normalized eigenfunction

ψ0(x)=(απ)1/4e−αx2/2(Chapter 5.23)(Chapter 5.23)ψ0(x)=(απ)1/4e−αx2/2

with the corresponding eigenvalue

E0=12ℏω(Chapter 5.24)(Chapter 5.24)E0=12ℏω


Drawing from our experience with the particle in a box, we might surmise that the first
excited state of the harmonic oscillator would be a function similar to
Equation Chapter 5.23Chapter 5.23, but with a node at x=0x=0, say,

ψ1(x)=constxe−αx2/2(Chapter 5.25)(Chapter 5.25)ψ1(x)=constxe−αx2/2

This is orthogonal to ψ0(x)ψ0(x) by symmetry and is indeed an eigenfunction with the


eigenvalue

E1=32ℏω(Chapter 5.26)(Chapter 5.26)E1=32ℏω

Continuing the process, we try a function with two nodes

ψ2=const(x2−a)e−αx2/2(Chapter 5.27)(Chapter 5.27)ψ2=const(x2−a)e−αx2/2

Using the integrals tabulated in the Supplement 5, on Gaussian Integrals, we determine


that with a=12a=12 makes ψ2(x)ψ2(x) orthogonal to ψ0(x)ψ0(x) and ψ1(x)ψ1(x).
We verify that this is another eigenfunction, corresponding to

E2=52ℏω(Chapter 5.28)(Chapter 5.28)E2=52ℏω

The general result, which follows from a more advanced mathematical analysis, gives
the following formula for the normalized eigenfunctions:

ψn(x)=(α−−√2nn!π−−√)1/2Hn(α−−√x)e−αx2/2(Chapter 5.29)(Chapter
5.29)ψn(x)=(α2nn!π)1/2Hn(αx)e−αx2/2

where Hn(ξ)Hn(ξ) represents the Hermite polynomial of degree nn. The first few


Hermite polynomials are

H0(ξ)=1(Chapter 5.30)(Chapter 5.30)H0(ξ)=1


H1(ξ)=2ξ(Chapter 5.31)(Chapter 5.31)H1(ξ)=2ξ
H2(ξ)=4ξ2−2(Chapter 5.32)(Chapter 5.32)H2(ξ)=4ξ2−2
H3(ξ)=8ξ3−12ξ(Chapter 5.33)(Chapter 5.33)H3(ξ)=8ξ3−12ξ

The four lowest harmonic-oscillator eigenfunctions are plotted in


Figure Chapter5.3Chapter5.3. Note the topological resemblance to the corresponding
particle-in-a-box eigenfunctions.
Figure Chapter5.3Chapter5.3: Harmonic oscillator
eigenfunctions for n=0, 1, 2, 3.

The eigenvalues are given by the simple formula

En=(n+12)ℏω(Chapter 5.34)(Chapter 5.34)En=(n+12)ℏω

These are drawn in Figure Chapter5.2Chapter5.2, on the same scale as the potential


energy. The ground-state energy E0=12ℏωE0=12ℏω is greater than the classical
value of zero, again a consequence of the uncertainty principle. This means that the
oscillator is always oscillating.

It is remarkable that the difference between successive energy eigenvalues has a


constant value

ΔE=En+1−En=ℏω=hν(Chapter 5.35)(Chapter 5.35)ΔE=En+1−En=ℏω=hν

This is reminiscent of Planck’s formula for the energy of a photon. It comes as no


surprise then that the quantum theory of radiation has the structure of an assembly of
oscillators, with each oscillator representing a mode of electromagnetic waves of a
specified frequency.

You might also like