Quantum Mechanics — Lecture notes for PHYS223

XI Perturbation theory

Perturbation theory is a general method to analyse complex quantum systems in terms of simpler variants. The method relies on the expectation values, matrix elements and overlap integrals just introduced, which we now use to break down complex quantum processes into simpler parts. We focus on the simplest version of the method, which results in simple systematic approximations of energy levels. For completeness we also present details of the derivation, which illustrates the power of the Dirac notation.

XI.1 Objective

We want to find approximations for the energies En and eigenstates |ψn of a Hamiltonian H^,

En|ψn=H^|ψn, (139)

assuming that the Hamiltonian is of the form H^=H^(0)+w^, where H^(0) represents a simplified system with energies En(0) and eigenstates |ψn(0) solving the Schrödinger equation

En(0)|ψn(0)=H^(0)|ψn(0). (140)

The difference w^=H^-H^(0) between the real and the simplified system is called the perturbation, and the approximation scheme is called perturbation theory.

XI.2 Method

The idea of perturbation theory is to assume that the perturbation w^=λW^ is the product of an operator W^ and a small number λ (determining the strength of the perturbation).

Since the Hamiltonian

H^=H^(0)+λW^ (141)

now depends on the parameter λ, the energies En and eigenstates |ψn also depend on this parameter.

Consequently, we can expand the energies and eigenstates into a series:

En(λ) = En(0)+λEn(1)+λ2En(2)+ (142)
= l=0λlEn(l),
|ψn(λ) = |ψn(0)+λ|ψn(1)+λ2|ψn(2)+ (143)
= l=0λl|ψn(l).

Perturbation theory provides a systematic scheme to express the quantities En(l) and |ψn(l) in terms of the unperturbed energies En(0), the unperturbed eigenstates |ψn(0), and the matrix elements of the perturbation

Wmn=ψm(0)|W^|ψn(0).

The scheme commences by introducing the series expansions (142) and (143) into the Schrödinger equation (139) and sorting the expressions order by order in λ:

En(0)|ψn(0)+
λ(En(0)|ψn(1)+En(1)|ψn(0))+
λ2(En(0)|ψn(2)+En(1)|ψn(1)+En(2)|ψn(0))+
=
H^(0)|ψn(0)+
λ(H^(0)|ψn(1)+W^|ψn(0))+
λ2(H^(0)|ψn(2)+W^|ψn(1))+
(144)

In order to fulfill this equation, the expressions in front of the powers λl have to be identical on both sides of the equation.

Since we assume that the perturbation λ is small, we concentrate on the first few orders.

Zeroth order perturbation theory:
In zeroth order of λ (collecting all terms that are independent of λ) one finds

λ0: En(0)|ψn(0)=H^(0)|ψn(0). (145)

This equation is identical to the unperturbed Schrödinger equation (140) and hence verifies that we chose the correct leading coefficient in each series. This result had to be expected since for λ=0 the Hamiltonian H^=H^(0) equals the unperturbed Hamiltonian. Therefore the energies En(λ=0)=En(0) are equal to the unperturbed energies, and the eigenstates |ψn(λ=0)=|ψn(0) are identical with the unperturbed eigenstates.

First order perturbation theory:
In first order of λ (collecting all terms that are proportional to λ) one finds

λ1:En(0)|ψn(1)+En(1)|ψn(0)=H^(0)|ψn(1)+W^|ψn(0). (146)

The coefficient En(1) and the state |ψn(1) can now be determined using the orthonormality of the unperturbed eigenstates |ψn(0), i.e., ψm(0)|ψn(0)=δnm, where δnm=1 if n=m and δnm=0 if nm.

In order to use this property, take the scalar product of an unperturbed eigenstate |ψm(0) with both sides of Eq. (146):

En(0)ψm(0)|ψn(1)+En(1)δnm=Em(0)ψm(0)|ψn(1)+Wmn. (147)

Here we used ψm(0)|H^(0)ψn(1)=H^(0)ψm(0)|ψn(1)=Em(0)ψm(0)|ψn(1) and the definition of the matrix element Wmn.

For n=m, Eq. (147) reduces to

En(1)=Wnn, (148)

while for nm we find

ψm(0)|ψn(1)=WmnEn(0)-Em(0). (149)

Since the unperturbed eigenstates form a complete basis this gives

|ψn(1)=mnWmnEn(0)-Em(0)|ψm(0), (150)

where the sum is over all indices m, with the exception of m=n.

Second order perturbation theory:
By similar steps, one derives in the second order of λ

En(2)=mn|Wmn|2En(0)-Em(0). (151)

In this course, we will not need |ψn(2), En(3), or higher terms of the perturbation series.

XI.3 Summary of the result

  • Perturbed energy in first order of λ:

    EnEn(0)+λWnn. (152)

    Hence, the energy shift En-En(0)=λψn(0)|W^|ψn(0)=w^ is given by the expectation value of the perturbation w^.

  • Perturbed eigenstate in first order of λ:

    |ψn|ψn(0)+λmnWmnEn(0)-Em(0)|ψm(0). (153)
  • Perturbed energy in second order of λ:

    EnEn(0)+λWnn+λ2mn|Wmn|2En(0)-Em(0). (154)

In the following we only need these results, not any details of their derivation.

XI.4 Example I: worn out harmonic oscillator

We now discuss various simple examples of perturbed systems, including cases where we can compare to exact solutions. The first example is a worn-out oscillator, described by a reduced restoring force F=-mω02x+λx. This corresponds to a potential V=12mω02x2-12λx2, which is still parabolic, and hence still describes a harmonic oscillator.

Exact solution: The new oscillation frequency can be calculated from the curvature of the potential, mω2=mω02-λ, hence ω=ω02-λ/m. This determines the exact perturbed energies

En=ω(n+12). (155)

For small λ, the perturbed frequency can be approximated as

ωω0-λ2mω0. (156)

We focus on the ground state, with n=0. The perturbed ground state energy can be approximated as

E012ω0-λ4mω0. (157)

The first term is the unperturbed ground-state energy E0(0)=ω0/2. Consequently, the energy shift of the ground state is

E0-E0(0)-λ4mω0. (158)

Perturbation theory: Let us see whether we can recover this result in perturbation theory.

The Hamiltonian

H^=12mp^2+12mω02x2-12λx2 (159)

differs from the unperturbed harmonic oscillator by the perturbation w^=-12λx2. According to first-order perturbation theory, the energy shift of the states is given by the expectation value of this perturbation, calculated with the unperturbed states. Hence, the ground-state energy shift is

E0-E0(0)ψ0(0)|w^|ψ0(0)=-12λψ0(0)|x^2|ψ0(0), (160)

where ψ0(0)(x)=(mω0π)1/4exp(-x2mω02) is the ground state wave function of the unperturbed harmonic oscillator (see section VI). This state has the form of a minimal-uncertainty wave packet (see section VII.5), for which we already obtained ψ0(0)|x^2|ψ0(0)=2mω0. Consequently, first-order perturbation theory predicts

E0-E0(0)-12λ2mω0, (161)

which indeed agrees with Eq. (158).

XI.5 Example II: harmonic oscillator exposed to a constant force

Another exactly solvable perturbed problem describes a harmonic oscillator subjected to an additional constant force of strength λ, F=-mω2x+λ. A constant force simply shifts the equilibrium position (the position where the force vanishes) to

x0=λ/(mω2). (162)

The potential energy can be written as

V(x)=12mω2x2-λx, (163)

hence, the perturbation is w^=-λx.

Exact solution: The potential energy can be written so that the new equilibrium position becomes apparent,

V(x)=12mω2(x-x0)2-12mω2x02. (164)

This describes again a harmonic oscillator with the same frequency ω as before, but oscillating around x=x0. The states ψn(x)=ψn(0)(x-x0) hence are also simply shifted by x0. We can now introduce a new coordinate y=x-x0, and recover the potential energy of the unperturbed harmonic oscillator. The only difference the term

-12mω2x02=-λ22mω2 (165)

which shifts the bound state energies to a lower value:

En-En(0)=-λ22mω2. (166)

This result is exact. Notice that the shift is proportional to the square λ2.

Perturbation theory: Let us see again whether we can recover this result in perturbation theory. Since the exact energy shift is proportional to the square of the perturbation strength we have to resort to second-order perturbation theory.

We again concentrate on the ground-state energy E0. The first-order energy shift E0-E0(0)λW00 is given by the expectation value of the perturbation W^=-x^,

W00=-ψ0(0)|x^|ψ0(0). (167)

Now notice that xψ0(0)(x)=/(2mω)ψ1(0)(x) is proportional to the wave function

ψ1(0)(x)=c1xexp(-x2mω2),

of the first excited state of the unperturbed system, calculated in section VI.

Hence

W00=-ψ0(0)|x|ψ0(0)=-/(2mω)ψ0(0)|ψ1(0)=0 (168)

due to the orthogonality of the eigenstates |ψ0(0) and |ψ1(0). This agrees with the absence of a term linear in λ in the exact expression (166) of the energy shift.

In second-order perturbation theory,

E0E0(0)+λW00+λ2mn|Wm0|2E0(0)-Em(0), (169)

we have to calculate the matrix elements

Wm0 = ψm(0)|W^|ψ0(0) (170)
= -ψm(0)|x^|ψ0(0) (171)
= -/(2mω)ψm(0)|ψ1(0) (172)

Again because of the orthogonality of the eigenstates |ψn(0), all of these matrix elements vanish, with the exception of the element

W10=-/(2mω). (173)

Hence the energy shift reduces to

E0-E0(0)|W10|2E0(0)-E1(0). (174)

Moreover, since En(0)=ω(n+1/2), we have E0(0)-E1(0)=-ω. Collecting all results we obtain

E0-E0(0)λ2/(2mω)-ω=-λ22mω2, (175)

which agrees with the exact result (166).

XI.6 Example III: anharmonic oscillator

The parabolic potential Vosc(x)=12mω2x2 is often used as an approximation of the motion in around an equilibrium position x-x0, such as in the vibration of molecules or solids. The exact potential V(x) can be expanded in a Taylor series around the equilibrium position,

V(x) = V0+12mω2(x-x0)2
+terms of order (x-x0)3 and higher.

In this expansion there are no terms linear in (x-x0) since the force F=-V vanishes at the equilibrium position x=x0.

As we have seen in example II, the constant V0 simply shifts the energy, and the displacement by x0 does not change the energies at all. In the following we set x0=0, V0=0.

The terms of order x3 and higher are called the anharmonicity of the potential. Classically, they perturb the motion of the oscillator so that the oscillation period T depends on the energy of the oscillator (recall the period T of a harmonic oscillator is independent of the oscillation amplitude Δx). Quantum mechanical, this results in a perturbation of the bound-state energies. For small energies, the classical oscillation amplitude Δx is small, and hence the effect of the anharmonic terms xnΔxn should be small and rapidly decrease with n. We hence only account for the anharmonic term of lowest order in the Taylor series,

V(x)=12mω2x2+λxnn! (177)

where typically n=3, but sometimes also n=4 or larger due to symmetries of the problem.

Now we apply perturbation theory in order to estimate the resulting energy shift of the ground state.

In first-order perturbation theory, the energy shift is given by the expectation value of the perturbation λxn/n!,

E0-E0(0)=λn!ψ0(0)|xn|ψ0(0). (178)

We hence have to calculate the overlap integral

ψ0(0)|xn|ψ0(0) = -xn|ψ0(0)(x)|2𝑑x
= -xn(mωπ)1/2exp(-x2mω)𝑑x.

This integral vanishes if n is an odd integer, since the integrand is antisymmetric (the integral from - to 0 exactly cancels the integral from 0 to ). Hence, in first-order perturbation theory

E0-E0(0)0(n odd). (181)

An energy shift is only found in second-order perturbation theory, which we however do not pursue for the present problem.

For n and even integer we use the standard integral

-xnexp(-ax2)𝑑x=πa-(n+1)/22-nn!(n/2)!,

where we set a=mω. This gives the first order energy shift

E0-E0(0)λ(mω)n/22-n(n/2)!. (182)

It is convenient to express this result in terms of the uncertainty Δx=2mω of position in the ground state, which is the equivalent of the classical oscillation amplitude. One then obtains

E0-E0(0)λ(Δx)n2-n/2(n/2)!,(n even). (183)

Note that the numerical coefficient rapidly falls of with increasing n.