The development of time-dependent perturbation
theory was initiated by Paul Dirac’s early work on the
semi-classical description of atoms interacting with
electromagnetic fields1. Dirac, Wheeler, Heisenberg, Feynman
and Dyson developed it into a powerful set of techniques for
studying interactions and time evolution in quantum mechanical
systems which cannot be solved exactly. It is used for the
quantitative description of phenomena as diverse as proton-proton
scattering, photo-ionization of materials, scattering of electrons
off lattice defects in a conductor, scattering of neutrons off
nuclei, electric susceptibilities of materials, neutron absorption
cross sections in a nuclear reactor etc. The list is infinitely
long. Time-dependent perturbation theory is an extremely important
tool for calculating properties of any physical system.
So far all the Hamiltonians which we had studied
were time-independent. This property was particularly important for
the time-energy Fourier transformation from the time-dependent
Schrödinger equation to a time-independent Schrödinger equation.
Time-independence of H also
ensures conservation of energy, as will be discussed in detail in
Chapter 16 Time-dependent perturbation
theory, on the other hand, is naturally also concerned with
time-dependent Hamiltonians H(t) (although it provides very useful
results also for time-independent Hamiltonians, and we will see
later that most of its applications in quantum field theory concern
systems with time-independent Hamiltonians). We will therefore
formulate all results in this chapter for time-dependent
Hamiltonians, and only specify to time-independent cases where it
is particularly useful for applications.
13.1 Pictures of quantum dynamics
As a preparation for the
discussion of time-dependent perturbation theory (and of field
quantization later on), we now enter the discussion of different
pictures of quantum
dynamics.
The picture which we have used so far is the
Schrödinger picture of
quantum dynamics: The time evolution of a system is encoded in its
states | ψ(t)〉 which have to satisfy a Schrödinger
equation iℏ
d | ψ
S (t)〉∕dt = H(t) | ψ S (t)〉. However, every transformation on
states and operators | ψ〉 → U | ψ〉, A → U ⋅ A ⋅ U + with a unitary operator
U leaves the matrix
elements 〈ϕ | A | ψ〉 and therefore the observables of a
system invariant.
If U is
in particular a time-dependent unitary operator, then this changes
the time-evolution of the states and operators without changing the
time-evolution of the observables. Application of a time-dependent
U(t) corresponds to a change of the
picture of quantum dynamics, and two important cases besides the
Schrödinger picture are the Heisenberg picture and the interaction (or Dirac) picture. In the
Heisenberg picture all time dependence is cast from the states onto
the operators, whereas in the Dirac picture the operators follow a
“free” (or better: exactly solvable) time evolution, while the
interaction (non-solvable) part of the Hamiltonian determines the
time evolution of the states.
There are essentially two reasons for introducing
the Heisenberg picture. The less important of these reasons is that
the Hamilton-Poisson formulation of the classical limit of quantum
systems is related to the Heisenberg picture. The really important
reason is that quantum field theory in Chapter 17 appears first in the Heisenberg
picture.
The rationale for introducing the Dirac picture
is that time-dependent perturbation theory automatically leads to
the calculation of matrix elements of the time evolution operator
in the Dirac picture. As soon as we want to calculate transition
probabilities in a quantum system under the influence of
time-dependent perturbations, we automatically encounter the Dirac
picture.
Before immersing ourselves into the discussion of
the Heisenberg and Dirac pictures, we have to take a closer look at
time evolution in the Schrödinger picture.
Time evolution in the Schrödinger
picture
In the Schrödinger picture the basic operators
(like x or p) are time-independent, , and all the time evolution from
the dynamics is carried by the states. The differential equation
yields an equivalent integral equation
and iteration of this equation yields
with the time evolution
operator 2
Taking the adjoint switches t with t 0 in the argument of the
time evolution operator,
This and the composition law (13.7) below imply
unitarity of the time evolution operator.
(13.1)
(13.2)
Please note that the time ordering operator T in
equations (13.1) and (13.2) always ensures that
the Hamiltonians are ordered from right to left such that their
time arguments go from closer to the lower integration boundary
(t 0 in
equation (13.1), t
in equation (13.2)) to the upper integration boundary
(t in
equation (13.1), t
0 in equation (13.2)), irrespective of whether the upper integration
boundary is larger or smaller than the lower integration
boundary, e.g. if t > t 0 in
equation (13.1) then of course t 0 < t in equation (13.2). Apparently, the
identification of “lower” and “upper” integration boundary in the
previous statement implies the convention that the integrand in the
exponent is − iH(t)∕ℏ. Otherwise the statement would be
ambiguous.
The re-ordering of integrations in the second
lines of equations (13.1, 13.2) is trivial for the 0th and 1st order
terms. For the higher order terms e.g. in
equation (13.1) we can recursively use for any consecutive
pair of integrations
which proves the re-ordering for n = 2, see also
Figure 13.1. For higher n we can perform an induction step,
which concludes the proof.
(13.3)
(13.4)
Fig. 13.1
Time evolution operators satisfy several
important properties which include Schrödinger type operator
equations, unitarity, and a simple composition property. We begin
with the Schrödinger type differential equations satisfied by
U(t, t′).
The derivative with respect to the first time
argument of the time evolution operator is most easily calculated
using the representation in the first line of (13.1),
while the derivative with respect to the second argument follows
most easily from the second line in (13.1),
Taking the adjoint of (13.5) or using (13.2) yields
The time evolution operator is the unique solution of these
differential equations with initial condition U(t 0, t 0) = 1. The differential
equations together with the initial condition also imply the
integral equations
(13.5)
(13.6)
Another important property of the time evolution
operator is the composition law
Proving this through multiplication of the left hand side and
sorting out the n-th order
term is clumsy, due to the need to prove that the sum over
n + 1 n-fold integrals on the left hand side
really produces the n-th
order term on the right hand side. However, we can find a much more
elegant proof by observing that U(t′, t)U(t, t 0) is actually independent
of t due to
equations (13.5, 13.6),
and therefore
The composition law yields in particular
and combined with (13.2) this implies unitarity of the time
evolution operator,
i.e. time evolution preserves the norm of states.
(13.7)
(13.8)
The time evolution operator for the harmonic
oscillator
The time evolution operator for time-independent
Hamiltonians H is invariant
under time translations,
The matrix elements in x
space can then be written in terms of the wave functions of energy
eigenstates H | E, ν〉 = E | E, ν〉,
where ν is a set of
degeneracy indices. There are no degeneracy indices in one
dimension and the expansion takes the form
E.g. the time evolution operator of the harmonic
oscillator
has matrix elements
Use of the Mehler formula (D.8) yields
(13.9)
To use the Mehler formula we should take
ω → ω − iε for t > 0. This complies with the shifts
E′ → E′ − iε,
which define retarded Green’s functions in the energy
representation, see e.g. (11.8, 20.14). The time-dependent
retarded Green’s function for the oscillator is related to the
propagator (13.9) in the standard way
The Heisenberg picture
In the Heisenberg picture we use the unitary time
evolution operator U(t, t 0) to cast the time
dependence from the states onto the operators,
For the time evolution of the operators in the Heisenberg picture
we observe that
and the Heisenberg evolution equation
In the last equation, H
H (t) is the Hamiltonian written in terms
of operators
in the Heisenberg picture.
(13.10)
For time-dependent we have
13.2 The Dirac picture
For the Dirac or interaction picture we split the
Schrödinger picture Hamiltonian H(t) into a “free” (or rather: solvable)
part H
0(t) and an
“interaction” (or rather: perturbation) part V (t),
and define the “free” time evolution operator
(13.11)
The common terminology of denoting H 0(t) and U 0(t, t 0) as “free” Hamiltonian
and time evolution operator while V (t) is the “interaction” part is
motivated from scattering theory, which is one of the most common
applications of time-dependent perturbation theory. However, we
should always keep in mind that H 0(t) does not really need to be a free
particle Hamiltonian. E.g. for a hydrogen atom under the influence
of an external electromagnetic field with wavelength λ ≫ a 0, the “free” part
H 0 would
actually be the hydrogen Hamiltonian including the Coulomb
interaction between the proton and the electron, while V (t) would describe the effective
coupling of the electromagnetic field to the quasiparticle which
describes relative motion in the hydrogen atom. We will discuss
this case in detail in Chapter 18, and in particular in
Section 18.4
The interaction picture splits off the solvable
part of the time evolution from the states,
where the last line identifies the time evolution operator
U D (t, t′) acting on the states in the
interaction picture.
(13.12)
The solvable part of the time evolution is cast
onto the operators
to preserve the time evolution of matrix elements and expectation
values in the interaction picture.
(13.13)
Substituting the composition law for time
evolution operators confirms that evolves freely between different
times,
and substituting
for
shows that
is related to the operator in the Heisenberg picture through the
particular variant U
D (t, t 0) of the interaction
picture evolution operator U D (t, t′) (13.12),
(13.14)
(13.15)
The differential equation for time evolution of
the operators is
where in the last equation (similar to the previous remark for the
Heisenberg picture) H
0, D
(t) is written in terms of
operators
in the Dirac picture.
The interactions are encoded in the time
evolution of the states,
where again U 0
+(t, t
0)V
(t)U 0(t, t 0) = V D (t) ≡ H D (t) due to the operator transition
in the
Hamiltonians.
(13.16)
Conversion of equation (13.16) into the
equivalent integral equation gives us another equation for the time
evolution operator U
D (t, t′) for the states in the Dirac
picture,
This evolution operator apparently satisfies
(13.17)
We have split the time evolution asymmetrically
between states and operators, and therefore there are two
Hamiltonians and related time evolution operators in the
interaction picture: the “free” Hamiltonian H 0(t) for the evolution of the operators
and the interaction Hamiltonian H D (t) for the evolution of the states (and
then there is the third Hamiltonian H(t) and its time evolution operator
appearing in the derivation of the interaction picture).
If we substitute3
into equation (13.17) and use the composition property for
time evolution operators
we find
(13.18)
The n-th
term in the sum can be interpreted as n scatterings at the perturbation
V (t), with “free” time evolution under
the Hamiltonian H
0(t) between any
two scattering events, see Figure 13.2. In the end
everything is evolved again to the fiducial time t 0.
Equation (13.21) below will show that this as a
consequence of the fact that we will express transition probability
amplitudes in terms of states at some fixed time t 0.
Fig. 13.2
Scattering off a time-dependent
perturbation
Dirac picture for constant H 0
We have H
0 = H
0, D if
H 0 is a
time-independent operator in the Schrödinger picture, because
H 0 and
U 0(t, t 0) = exp[−iH 0(t − t 0)∕ℏ] commute.
The Hamiltonian H D (t) acting on the states in the
interaction picture is related to the Hamiltonian with the ordinary
operators p, x,… of the Schrödinger picture via
The time evolution operator for the states in the interaction
picture is then
(13.19)
The case of time-independent unperturbed
operators H 0 is
the most common case in applications of time-dependent perturbation
theory. Equation (13.19) therefore shows the most commonly
employed form of U
D (t, t′) for the evaluation of the
transition amplitudes or scattering matrix elements which will be
introduced in Section 13.3.
13.3 Transitions between discrete states
We are now in a position to discuss transitions
in a quantum system under the influence of time-dependent
perturbations. We are still operating in the framework of
“ordinary” quantum mechanics (“first quantized theory”), and at
this stage time-dependent perturbations of a quantum system arise
from time dependence of the parameters in the Schrödinger
equation.
We will denote states as discrete states if they can be
characterized by a set of discrete quantum numbers, e.g. the bound
energy eigenstates | n, ℓ, m ℓ , m s 〉 of hydrogen or the
states | n
1, n
2, n
3〉 of a three-dimensional harmonic oscillator are
discrete. States which require at least one continuous quantum
number for their labeling are denoted as continuous states. Momentum eigenstates
are examples of
continuous states. Quantum mechanical transitions involving
continuous states require special considerations. Therefore we will
first discuss transitions between discrete states, e.g. transitions
between atomic or molecular bound states.
We consider a system with an unperturbed
Hamiltonian H 0
under the influence of a perturbation V (t):
The perturbation operator will in general be a function of the
operators p and x, V (t) ≡ V (p, x, t). We will see later that in many
applications V
(t) has the form
(13.20)
In this section we assume that all states under
consideration can be normalized to 1.
For the calculation of transition probabilities
in the system, recall that the expansion of a general
state | ϕ〉 in terms of an
orthonormal complete set of states | ψ n 〉 is
and therefore the probability P n (ϕ) of finding the state | ψ n 〉 in a measurement performed on
the state | ϕ〉 is
P n (ϕ) = | 〈ψ n | ϕ〉 | 2. We can also
understand this as the expectation value of the projection
operator | ψ
n 〉〈ψ n | in the state | ϕ〉.
Now assume that the state | ϕ〉 is a state | ψ in (t)〉, where the state at an earlier time
t′ < t was an unperturbed
state | ψ in (0)(t′)〉, typically an eigenstate of
H 0. Then we
know that the state at time t is
and since the state now evolved with the full Hamiltonian including
the perturbation V
(t), it will not be an
unperturbed state any more, but a superposition of unperturbed
states. If at time t a
measurement is performed on the state | ψ in (t)〉, the probability to measure a
certain unperturbed state | ψ out (0)(t)〉 will be | 〈ψ out (0)(t) | ψ in (t)〉 | 2.
Therefore the probability amplitude for
transition from an unperturbed state | ψ in (0)(t′)〉 to an unperturbed
state | ψ out (0)(t)〉 between times t′ and t is
The transition probability
amplitudes between unperturbed states are matrix elements of the
time evolution operator in the interaction picture, where
the unperturbed states are taken at some arbitrary fixed
time.
(13.21)
The Schrödinger equations for the unperturbed
states | ψ
(0)(t
0)〉 and the free evolution operators U 0(t′, t 0) and U 0
+(t, t
0) imply
i.e. the choice of the parameter t 0 is (of course)
irrelevant for the transition matrix element. We set t 0 = 0 in the
following.
(13.22)
If we substitute the expansion (13.18) for the time
evolution operator in the interaction picture we get a series
(13.23)
Now we assume that our unperturbed states are
energy eigenstates
of the unperturbed Hamiltonian. Equation (13.23) then yields for
the transition probability amplitude between eigenstates of
H 0 (see also
equation (13.19)),
with the transition frequencies ω nm = (E n − E m )∕ℏ.
(13.24)
The transition probability from a discrete
state | m〉 to a discrete
state | n〉 is then
(13.25)
Equation (13.24) assumes that we use eigenstates of
H 0 for the
initial and final states, but equation (13.25) holds for
arbitrary discrete initial and final states, and we even do not
have to require the same basis for the decomposition of the initial
and the final state, i.e. equation (13.25) also holds if
m and n are discrete quantum numbers
referring to different bases of states.
P
m → n (t, t′) is a dimensionless positive number
if both the initial and final states are discrete states, i.e.
dimensionless states (see the discussion of dimensions of states in
Section 5.3), and due to the unitarity of
U D (t, t′) it is also properly normalized as a
probability,
As a corollary, this observation also implies ,
as required for a probability.
We will denote the transition probability amplitude
〈n | U D (t, t′) | m〉 also as a scattering matrix element or
S matrix element,
(13.26)
In the literature this definition is more
commonly employed with default values t → ∞, t′ → −∞ for the initial and final times,
S nm ≡ S nm (∞, −∞). It is also usually reserved for
transitions with two particles in the initial state (to be
discussed in Chapter 17 and following chapters), but here
we are still dealing with a single particle perturbed by a
potential V (t), or an effective single particle
description of relative motion of two particles. The connection
with many particle scattering theory later on is easier if we
introduce the scattering matrix already for single particle
problems, and it is also useful to have this notion available for
arbitrary initial and final times.
Møller operators
At this point it is also interesting to note a
factorized representation of the time evolution operator in the
interaction picture, which is applicable if both H and H 0 do not depend on time.
In this case we have with t
0 = 0,
with the Møller operator
Let us repeat the basic equation (13.21) and substitute
this definition,
Here we have introduced states
For the interpretation of these states we notice
i.e. is the fictitious
interacting state at time t
0 = 0 which yields the unperturbed state | ψ (0)(t)〉 at time t under full time evolution from t 0 = 0 to t.
(13.27)
In the framework or quantum mechanics, the case
that both H and
H 0 are
time-independent would often be dealt with in the framework of
time-independent perturbation theory or potential scattering
theory. However, we will see later that in the framework of quantum
field theory, time-independent H and H 0 is very common in
applications of time-dependent perturbation theory.
First order transition probability between
discrete energy eigenstates
For n ≠
m, the first order result
for S nm is the matrix element of the
Fourier component V
(ω nm ),
(13.28)
If the time dependence of the perturbation
V (t) is such that the Fourier transform
V (ω) exists in the sense of standard
Fourier theory (i.e. if V
(ω) is a sufficiently well
behaved function, which is the case e.g. if V (t) is absolutely integrable or square
integrable with respect to t), then the first order scattering
matrix (13.28) provides us with finite first order
approximations for transition probabilities
(13.29)
Note that the Fourier transform
of a potential V
(t) has the dimension
energy×time. Therefore P
m → n is a dimensionless number, as
it should be. Furthermore, the probability interpretation and the
use of first order perturbation theory entail that we should have
.
Otherwise first order perturbation theory is not applicable and
higher order terms must be included to estimate transition
probabilities.
The first order transition probability between
discrete states requires existence of a regular Fourier transform
V (ω) of the perturbation V (t). This condition is not satisfied in
the important case of monochromatic perturbations like V (t) = Wexp(−iω t), which have a δ function as Fourier transform,
Consistent treatment of this case requires that at least one of the
states involved is part of a continuum of states, as discussed in
Sections 13.4 and 13.5. If both the initial and final atomic or
molecular state are discrete, then the perturbation V (t) = Wexp(−iω t) must be treated as arising from a
quantized field which comes with its own continuum of states.
Monochromatic perturbations V (t) = Wexp(±iω t) typically arise from photon
absorption or emission, and the previous statement simply means
that the consistent treatment of transitions between bound states
due to monochromatic perturbations requires the full quantum theory
of the photon, see Section 18.6 See also
Problem 13.6 for an explanation why the Golden Rule of
first order perturbation theory, which is discussed in the next
section, cannot be used for transitions between discrete
states.
13.4 Transitions from discrete states into continuous states: Ionization or decay rates
Ionization of atoms or molecules, transitions
from discrete donor states into conduction bands in n-doped
semiconductors, or disintegration of nuclei are processes where
particles make a transition from discrete states into states in a
continuum.
We assume that the unperturbed Hamiltonian
H 0 contains an
attractive radially symmetric potential which generates bound
states | n, ℓ, m〉, where ℓ and m are the usual angular momentum
quantum numbers for the bound states and the quantum number
n labels the energy levels.
The free states for H
0 are usually given in terms of hypergeometric
functions, e.g. the Coulomb waves | k, ℓ, m〉 from Section 7.9
Here we initially use plane wave states instead
and ask for the probability for the system to go from a bound
state | n, ℓ, m〉 into a plane wave state under the influence of
a perturbation V
(t). This is a
simplification, but the prize that we pay is that the transition
matrix elements from a bound state into plane waves do not
necessarily tell us something about ionization or decay of a bound
system, because those transition matrix elements will also not
vanish for perturbations which primarily generate another bound
state since the bound states can also be written as superpositions
of plane waves, see e.g. Problem 13.7. Therefore the
transition matrix elements into plane wave states generically
correspond to a mixture of transitions into bound states and free
states. However, the focus in this preliminary discussion is not
the calculation of actual ionization or decay rates, but to explain
how continuous final states affect the interpretation of transition
matrix elements.
For continuous final states like , the appropriate
projection of U
D (t, t′) | ψ in (0)〉 is onto the
dimensionless combination
(recall from
that the plane wave states in three dimensions
have length dimension length3∕2, see
Section 5.3). This means that in a
transition from a discrete state | n, ℓ, m〉 into a momentum eigenstate
, the dimensionless quantity
is a differential transition
probability amplitude, in the sense that
is a differential transition
probability for the transition from the discrete state into
a volume element around the vector in momentum space. The meaning of
this statement is that
is the transition
probability from the discrete state | n, ℓ, m〉 into a volume
in -space. Another way to say this is to
denote the quantity with the dimension length3
as the transition probability
density per -space volume. The S matrix element
is then a transition probability
density amplitude (just like a wave function is a
probability density
amplitude rather than a probability amplitude, but for
obvious reasons neither of these designations are ever used).
(13.30)
With this interpretation, the transition
amplitudes into continuous states yield correctly normalized
probabilities, e.g. for plane waves,
The important conclusion from this is that transition matrix
elements of U
D (t, t′) from discrete states into
continuous final states yield transition probability densities, which have to be integrated
to yield transition probabilities. We will also rediscover this in
the framework of the spherical Coulomb waves in the following
subsection.
Ionization probabilities for hydrogen
Now that we have clarified the meaning of
transition amplitudes from discrete states into continuous states
with the familiar basis of plane wave states, let us come back to
the ionization or decay problems, i.e. transitions from the
discrete bound spectrum of an unperturbed Hamiltonian H 0 into the continuum of
unbound states. We will use hydrogen states as an example, but the
derivations go through in the same way for any Hamiltonian
H 0 with
discrete and continuous states.
The unperturbed Hamiltonian for hydrogen is
and the ionization problem concerns transitions from bound
states | n, ℓ, m〉 into Coulomb waves | k, ℓ, m〉 under the influence of a
time-dependent perturbation4 V (t). The contribution from Coulomb waves
to the decomposition of unity in terms of hydrogen states came with
a measure k 2
dk (7.75),
where we also introduced an energy representation for the spherical
Coulomb waves, | E, ℓ, m〉 = | k, ℓ, m〉,
and the corresponding density of spherical Coulomb waves in the
energy scale,
This differs from (12.13) for d = 3 by missing a factor g∕2π 2 = g4π∕8π 3. The spin factor is
g = 1, because spin flips
can usually be neglected in ionization transitions. Inclusion of
spin quantum numbers m
s and
m′ s for the initial and final
states would therefore result in a factor . There is no factor
4π because the angular
directions in space have been discretized in terms
of angular momentum quantum numbers (ℓ, m), and there is no factor
(2π)−3 because
the density ϱ(E) in equation (13.32) is a number of
states per unit of energy, but it is not a number of states per energy and
volume (remember
V → (2π)3 in the continuum limit).
It comes in units cm−3eV−1 because the
projector | k, ℓ, m〉〈k, ℓ, m | for spherical Coulomb waves has
dimension length3, and therefore scattering matrix
elements
from bound states into ionized states come in units of
cm3. Please also recall the remark after
equation (12.8).
(13.31)
(13.32)
Suppose we start with an unperturbed bound
state | n, ℓ, m〉. We can calculate two kinds of
scattering matrix elements, viz. for transitions into bound states,
and into ionized states
For the sums of the absolute squares of these scattering matrix
elements, we observe from the completeness relation for hydrogen
states, the unitarity of U
D (∞, −∞), and 〈n, ℓ, m | n, ℓ, m〉 = 1 that
This confirms ,
as is required for transition probabilities between bound states,
but it also tells us that
must be the ionization probability due to the perturbation
V (t), since the sum over all transition
probabilities into bound states is
This confirms again that absolute squares of scattering matrix
elements into continuous final states must be integrated against
final state densities to yield transition probabilities, where the
appropriate density of final states follows from the completeness
relation of the unperturbed system.
(13.33)
If we want to know the probability for the
hydrogen atom to ionize into a state with energy
0 < E
1 ≤ E ≤ E 2 for the relative motion
between proton and electron, we have to calculate
On the other hand, if we only know the energy level E n of the initial bound state, we
would calculate the ionization probability of the atom as a
weighted average
(13.34)
The first order results for the ionization
probabilities follow from the first order scattering matrix
elements
with the transition frequency ω kn = (E k − E n )∕ℏ. This assumes that the Fourier
transformed operator V
(ω kn ) exists in the sense of
standard Fourier theory. The case of a monochromatic perturbation,
for which the Fourier transform is a δ function in frequency space, requires
special treatment and is discussed in the following
subsection.
(13.35)
Even for well behaved Fourier transform
V (ω kn ), use of the first order
result (13.35) to estimate the ionization probability,
can only make sense for P
n, ℓ, m → E > 0
(1) ≤ 1.
The Golden Rule for transitions from discrete
states into a continuum of states
Now assume that we perturb a hydrogen atom in the
initial bound state | n, ℓ, m〉 with a monochromatic
perturbation5
The corresponding scattering matrix element for transition into the
ionized state | E, ℓ′, m′〉 is in first order
The cross multiplication terms in | S E, ℓ′, m′; n, ℓ, m | 2 cancel for
ω ≠ 0 due to the
incompatibility of the δ
functions, and therefore we can focus in the following discussion
only on the first terms in equations (13.36–13.38), i.e. we
continue with
The square of this S matrix
element yields a factor
in dP n, ℓ, m → E, ℓ′, m′∕dE = ϱ(E) | S E, ℓ′, m′; n, ℓ, m | 2. Dividing by
the factor T provides us
with a differential transition rate into a final state energy
interval [E, E + dE],
Integration over the final state energy E then yields an expression for the
transition rate,
which is commonly referred to as the Golden Rule.
(13.36)
(13.37)
(13.38)
(13.39)
The total first order ionization rate of the
state | n, ℓ, m〉 under the
perturbation (13.36) is then
(13.40)
The standard expression for the Golden Rule for
the transition rate from a discrete state | m〉 into a continuous
state | n〉 due to the
perturbation V = Wexp(−iω t) is
(13.41)
This is also particularly popular for
time-independent V,
Quantum systems can have degeneracy between states | m〉 which are labelled by discrete
quantum numbers and states | n〉 with continuous quantum numbers.
Metastable states, or excited bound states in many-electron atoms
provide examples for this, and equation (13.42) would be the first
order expression for the decay rate of these states. An example for
this is the Auger effect, which is electron emission from atoms due
to Coulomb repulsion. The perturbation operator6 V = e 2∕4π | x 1 −x 2 | is time-independent, and
energy conservation is fulfilled because the discrete bound state
of two electrons in an excited atom can exceed the sum of ground
state energy and ionization energy, see Figure 13.3.
(13.42)
Fig. 13.3
Energy schematics for an Auger process. The
initial bound state of the two electrons has the same energy as the
final continuous state of an ion and a free electron
Time-dependent perturbation theory in second
order and the Golden Rule #1
We will discuss a time-independent perturbation
V,
and transition from a discrete state | m〉 into a continuous
state | n〉.
The completeness relation for the eigenstates of H 0 is
We will also write this symbolically as
If 〈n | V | m〉 = 0, the leading order term for the
scattering matrix element 〈n | U D (∞, −∞) | m〉 is the second order term
To make the τ′ integral
convergent, we add a small negative imaginary part to ω lm → ω lm − iε, so that the time integrals yield
This yields the second order scattering matrix element
and the differential transition rate
Integration yields the second order expression for the transition
rate,
(13.43)
(13.44)
(13.45)
Equation (13.45) tells us how transitions through virtual
intermediate states can generate the transition from | m〉 to | n〉 even if the direct transition is
forbidden due to a selection rule 〈n | V | m〉 = 0.
In his famous lectures on nuclear physics at the
University of Chicago in 1949, Fermi coined the phrase “Golden Rule
#2” for the first order transition rate (13.41, 13.42). He denoted the
corresponding second order expression for transition rates as
“Golden Rule #1”, because it is important for nuclear reactions
through intermediate compound nuclei [30].
13.5 Transitions from continuous states into discrete states: Capture cross sections
Transitions from continuous to discrete states
arise e.g. in the capture of electrons by ions, in the absorption
of an electron from a valence band into an acceptor state in a
p-doped semiconductor, in neutron capture by nuclei etc. Consider
e.g. the process | k, ℓ, m〉 → | n, ℓ′, m′〉 of absorption of an electron by an
H+ ion, where we still assume that the hydrogen
Hamiltonian (13.31) for relative motion is perturbed by
addition of an operator V
(t). From our previous
experience, we expect that the transition matrix element
yields a measure of probability for the absorption in the form of a
transition probability density
Indeed, the dimensionless number
is the probability that the state | n, ℓ′, m′〉 emerged from some capture
(p+ + e− → H) event rather than from an
internal transition in the hydrogen atom. This assumes again that
the perturbation V
(t) has a well behaved
Fourier transform V
(ω) such that the time
integrals in the perturbation series can be defined as classical
functions. However, a more common use of transition matrix elements
from continuous initial states is the calculation of cross sections
due to monochromatic perturbations. One possibility to calculate
capture or absorption due to a Coulomb potential is to use
parabolic coordinates because the incoming asymptotic plane wave
can be described in parabolic coordinates, just like in Rutherford
scattering [3]. However, radial
coordinates are just as convenient for this problem.
(13.46)
Calculation of the capture cross section
We will outline how to calculate the first order
cross section for the reaction p + + e − → H due to a monochromatic
perturbation7
V (t) = Wexp(iω t). For a judicious choice of the
operator W ≡ W(p, x) this describes electron-proton
recombination due to emission of a photon with energy ℏ ω. We will discuss these operators in
Chapter 18, but here we do not specify the
operator W further. Our
present focus is rather to develop the formalism for calculating
the capture or recombination cross section for a general
perturbation W(p, x)exp(iω
t). We should also mention that perturbations V (t) = V (x, t) due to interactions with additional
nearby electrons or ions are much more efficient and therefore more
important for electron capture than direct radiative recombination
due to photon emission.
The wave function for the approach between a free
electron and a proton in the effective single particle description
for relative motion is given by the wave function which
was constructed by Mott and Gordon in 1928 (7.76). The normalization factor is
irrelevant because it cancels in the cross section. For
convenience, it was chosen in equation (7.76) such that the asymptotic
incoming current density is
where μ is the reduced mass
of the two-particle system. This current density has units of
cm∕s = cm−2s−1∕cm−3 because it is
actually a current density per unit of volume in
space, which is a consequence of the
use of an asymptotic plane wave state in its calculation. A current
density per space volume is the correct notion
for the calculation of the electron-proton recombination cross
section, because the S
matrix element
yields a transition probability density per space volume
which comes in units of cm3, again due to the use of an
asymptotic plane wave state as incoming state.
(13.47)
contains the
factor
We can use this to calculate a transition rate density per
space volume
(13.48)
The transition rate is certainly proportional to
the asymptotic current density j in , and therefore we divide the
transition rate density by this current density to get a measure
for the probability of the absorption process | k〉 MG → | n, ℓ, m〉. This yields the absorption cross
section
with units of cm2. The total absorption cross section
due to the perturbation operator W(p, x)exp(iω
t) is then
(13.49)
The capture cross section enters into the
calculations of rate coefficients (σ v) av , where the notation indicates
averaging over the distribution of relative particle velocities in
a plasma of ions and electrons. The rate coefficients go into the
balance equations for electron and ion densities,
where in general additional terms due to collisional relaxation and
ionization have to be included. Due to (13.47) the rate
coefficients are directly related to the transition rates per
space volume calculated in the
state (7.76),
Calculations of radiative capture cross sections
for electron-proton recombination into arbitrary hydrogen shells
were performed in parabolic coordinates by Oppenheimer8 and by Bethe and Salpeter
[3]. Calculations in polar
coordinates had been performed by Wessel, Stückelberg and Morse,
and Stobbe9. All
these authors had noticed that the electron capture cross sections
for ions from radiative recombination were much too small to
explain the experimental values, and it was eventually recognized
that collisional relaxation due to interactions with spectator
particles dominated the observed recombination rates. Therefore
modern calculations of electron-ion recombination rates focus on
collisional relaxation, which means that the relevant perturbation
operators V are not
determined by photon emission but by Coulomb interactions in a
plasma, and the spectator particles also have to be taken into
account in the initial and final states. Electron-ion recombination
rates are particularly important for plasma physics and
astrophysics.
13.6 Transitions between continuous states: Scattering
For transitions between continuous states, e.g.
,
the S matrix element
is a quantity with the dimension length3, because both
external states have dimension length3∕2. We know from
Section 13.4 how to make sense of transition matrix
elements with continuous final states, viz. as transition probability
densities
in the final state space. We also know from the discussion in
Section 13.5 that a continuous initial state in the
scattering matrix will yield a transition probability density in
the space of initial states if V (ω) is a classical function. In that
case
will tell us the probability for transitions between states in
space volumes
and due to the perturbation V (t).
However, just like in Section 13.5, the most important
applications of scattering matrix elements with continuous initial
states concern the calculation of cross sections due to
monochromatic perturbations. We know from
Sections 13.4 and 13.5 that monochromatic perturbations call for
normalization of
by the reaction time T, and
we have learned in Section 13.5 that continuous initial states under the
influence of a monochromatic perturbation require normalization of
the transition rate with the current density j in of incident particles to
calculate a cross section for the quantum mechanical reaction
described by the S matrix
element, see equation (13.49). Our previous experience with initial or
final continuous states therefore motivates the definition of the
differential scattering cross
section
This has again the dimension length2, because the
incident current density j
in for plane
waves has units of cm∕s, see equation (13.47) and the following
discussion.
(13.50)
The notion of a differential scattering cross
section is sufficiently important to warrant rederivation of
equation (13.50) in simple steps in the next
paragraph.
Cross section for scattering off a periodic
perturbation
We apply the transition probability between
continuous states to calculate the scattering cross section for a
monochromatic perturbation
Our Hamiltonian is
and our unperturbed states are plane waves .
The first order result for the scattering matrix
is
The factor in the
scattering matrix element is also denoted as a scattering amplitude.
(13.51)
The transition probability density
contains the factor
and we can calculate a transition rate density
The corresponding differential transition rate into the final state
volume is
However, this still comes in units of cm3∕s instead of
s−1, due to the initial plane wave state. For initial
continuous states, we do not apply a volume measure (here
) in the space of initial
states, but normalize by the current density of the incident
particles. This yields a differential cross section for scattering
of momentum eigenstates,
The motivation for dividing out the current density j in of incoming particles from the
scattering rate is the trivial dependence of the scattering rate on
this parameter: if we double the number of incoming particles per
second or per cm2, we will trivially double the number
of scattering events per second. Therefore all the interesting
physics is in the proportionality factor d σ between j in and dw. This proportionality factor has the
dimension of an area, and in classical mechanics, integration of
d σ over for scattering of classical
particles off a hard sphere of radius r yields the cross section area of the
sphere σ = ∫d
σ = π r
2. Therefore the name differential scattering cross section
for d σ.
(13.52)
The current density
for a plane wave, ,
is actually a current density per unit of volume in -space. This is the correct current
density to be used in (13.52), because is a
transition rate per unit of volume in -space, and the ratio yields a
bona fide differential
cross section10.
Expressed in terms of continuum plane wave matrix elements, the
differential scattering cross section is
We can use the δ-function
in (13.53) to integrate over k′. This leaves us with a differential
cross section per unit of solid angle,
The corresponding result for ω = 0 (scattering off a static
potential) can also be derived within the framework of the
time-independent Schrödinger equation, see
Chapter 11 For the comparison note that we
can write the differential scattering cross
section (13.54) as
with the scattering amplitude
cf. (11.23), i.e.
equation (13.55) reduces to (11.27) for scattering off a static
potential if ω = 0. The
potential scattering formalism could be extended to time-dependent
perturbations by using the asymptotic expansion of the
time-dependent retarded Green’s function (11.46). However, the equivalent
scattering matrix formalism is more convenient.
(13.53)
(13.54)
(13.55)
(13.56)
Scattering theory in second order
We will discuss scattering off the
time-independent potential V in second order. The Hamiltonian is
If and ,
the leading order term for the S-matrix is the second order term
To make the τ′ integral convergent, we add a term
ε τ′ in the exponent, so
that the time integrals yield
and
The corresponding differential transition rate is
and the differential cross section for scattering of momentum
eigenstates in second order is
Integration over k′ yields
the differential scattering cross section per unit of solid angle
in second order,
(13.57)
(13.58)
13.7 Expansion of the scattering matrix to higher orders
For time-independent perturbation V we can write the expansion of the
scattering matrix in the form
Taking the limits t′ → −∞ and t → ∞, we find the equation
(13.59)
However, we can also use
This yields a form which resembles expressions for the shifts of
wave functions in time-independent perturbation theory,
(13.60)
If the initial state is
continuous, | S
fi
| 2 will enter into the calculation of cross sections.
If only the final state is continuous, | S fi | 2 will enter
into the calculation of decay rates. If both external states are
discrete, the perturbation V should be treated as arising from a
quantum field, see the remarks at the end of
Section 13.3.
This state satisfies the Lippmann-Schwinger
equation (11.5)
and therefore also
Indeed, one of the objectives of the original work of Lippmann and
Schwinger was to relate states of the form (13.62) to the scattering
matrix, and it was thought that they relate to the Møller
states (13.27). However, we now see that they instead
appear as stated in equation (13.61).
(13.63)
We can also write the result (13.61) in even neater
form
with the state
This state satisfies the equation
and therefore also
(13.64)
(13.65)
(13.66)
(13.67)
13.8 Energy-time uncertainty
We are now finally in a position to address the
origin of energy-time uncertainty in a more formal way. Energy
conservation in each term of the scattering
matrix (13.59) came from the final time integral over
τ 1, which in
symmetric form for the initial and final time limits can be written
as
However, this tells us that if we allocate only a finite time
window to
observe the evolution of the system, or if the system is forced to
make the transition within a time window ,
then we will observe violations of energy conservation of order
Here we used that the sinc function sin(x)∕x is rather broad with half maximum
near x = ±2.
(13.68)
(13.69)
How can that be? The theorem of energy
conservation for time-independent Hamiltonian H = H 0 + V in a static spacetime holds in
quantum mechanics just as in any other physical field theory. We
will see this in Section 16.2 However, by allocating a finite
time window
for our measurement device to observe the system, or by
constraining the system to make the transition within the fixed
finite time window, we apparently introduce a time-dependent
perturbation into the system that results in an energy uncertainty
in excess of in the final state.
13.9 Problems
Hints: The equation for n = 0 and n = 1 is trivial, and for n = 2 it can easily be demonstrated
from equation (13.3). This motivates a proof by induction with
respect to n, which can
easily be accomplished using the property
A more direct way to prove (13.70) is to express the ordering of operators
through appropriate functions under the assumption t > t 0 (forward evolution) or
t < t 0 (backward
evolution).
13.2. Use Fourier transformation to
calculate the matrix elements 〈x | U(t) | x′〉 for the free time evolution
operator in one dimension. Compare with the
result (13.9) for the harmonic oscillator.
13.3. Calculate the annihilation and
creation operators a(t) and a +(t) of the harmonic oscillator in the
Heisenberg picture.
Use the previous results to calculate the
operators x(t) and
p(t) for the harmonic
oscillator in the Heisenberg picture.
13.4. Start from the definition
of the time evolution operator of states in the interaction picture
to prove that
13.5. Calculate the first order
transition probability for the transition 1s → 2p for a hydrogen
atom which is perturbed by a potential
P and τ are constants. What is the meaning of
P in the limit τ → 0?
13.6. The Golden Rule #2 for the first
order transition rate (13.41) is often abused for the discussion of
transitions between discrete states. In this problem you will be
asked to figure out where the derivation of the Golden Rule #2 for
transitions between discrete states breaks down.
13.6a. Calculate the first order
transition probability for transitions between discrete energy
eigenstates | m〉 → | n〉 under the influence of a
monochromatic perturbation V (t) = Wexp(−iω t) which only acts between times
t′ and t. Which consistency requirements do
you find from the condition that the first order result describes a
transition probability P
m → n (t, t′)? Calculate also the transition rate
w m → n (t, t′) = dP m → n (t, t′)∕dt.
13.6b. Try to take the limit t − t′ → ∞ to derive the Golden Rule #2. Does
this comply with the consistency requirements
from 13.6a?
13.6c. Why do the inconsistencies
of 13.6b
not appear if the final state | n〉 is a continuous state?
Solution for Problem 13.6.
For a periodic perturbation V (t) = Wexp (−iω t) the first order transition
amplitude between times t′
and t, and between
different eigenstates of H
0 becomes
The resulting transition probability is
and the rate of change of the transition probability follows as
Equations (13.71) and (13.72) yield perfectly
well behaved, dimensionally correct expressions for the first order
transition probability and transition rate between discrete states.
Consistency with the probability interpretation for the extreme
case ω nm −ω = 0 requires
or alternatively, consistency of (13.71) with the
probability interpretation for arbitrary
requires
(13.71)
(13.72)
(13.73)
(13.74)
The problem arises with the limit , which would transform
the transition rate from an ordinary function of frequencies into a
δ function,
Here we used
Taking the limit violates either the
condition (13.73), or the condition (13.74) through its result
ω nm −ω → 0 for a transition. From this point
of view (and ignoring the fact that we should have at least one
continuous external state when properly taking into account
photons, see Section 18.6 and Problem 18.11), the
resolution of the paradox of emergence of a δ function between discrete states in
the limit is that in the region
of frequencies (13.74) where the first order result might be
applicable, the first order result becomes subdominant for large
and (at the very least) higher order terms would have to be
included to get estimates of transition probabilities and
transition rates, or perturbation theory is just not suitable any
more to get reliable estimates for those parameters.
(13.75)
These problems do not arise for continuous final
states, because in these cases P m → n (t, t′) → dP m → n (t, t′) = dE n ϱ(E n ) | S n, m (t, t′) | 2 are not transition
probabilities any more (which would be bounded by 1), but only
transition probability densities for which only the integral over
the energy scale with measure factor ϱ(E n ) is bounded.
13.7. Calculate the representation of the
ground state of hydrogen as a superposition of plane waves.
Solution.
Fourier transformation to space yields
The representation in terms of plane waves is therefore
i.e. the ground state is an isotropic superposition of plane waves
which is dominated by small wave numbers k ≲ 1∕a or large wavelengths λ ≳ 2π
a. This problem was included in this chapter to drive home
the point that calculation of transition rates into plane wave
states does not necessarily tell us something about scattering or
ionization in a system with bound states, unless the energy of the
final plane wave state is large compared to the binding energies of
the bound states.
13.8. Calculate the first order
ionization rate for particles which are trapped in a
one-dimensional δ-function
potential (Section 3.3), if the particles are perturbed
by a potential V
(t) = F 0xexp(−iω t). What is the meaning of the
constant F
0?
13.9. Calculate the first order capture
cross section for free particles with wave
function (3.18) which can become trapped in a
one-dimensional δ-function
potential, if the particles are perturbed by a potential
V (t) = F 0xexp(iω t). Recall that the normalization of
initial states does not matter in the calculation of cross sections
since it cancels in the ratio of capture rate to current
density.
13.10. Calculate the cross section for
recombination of an electron and a proton with energy ℏ 2 k 2∕2μ (in their relative motion) into the
ground state of hydrogen. Perform the calculation both in parabolic
and in polar coordinates.
13.11. Calculate the differential and
total scattering cross sections for particles with initial momentum
which are scattered off the
time-dependent potential
13.12. In 1984 Michael Berry published a
paper studying (among other things) the following interesting
question: Suppose H(t) is a time-dependent Hamiltonian with
the property that for each value of t there is a discrete spectrum
such that
We can relate the eigenstates of H(t) to the states | ψ n (t)〉 of the physical system described by
H(t) simply through the completeness
relation (13.77),
However, assume that we start with an eigenstate at time
t = 0, i.e. we are seeking
a solution of the initial value problem
Can we directly relate | ψ
n (t)〉 to without invoking a
superposition (13.78) of all the eigenstates ?
(13.76)
(13.77)
(13.78)
(13.79)
13.12a. Can you give an example of a
time-dependent Hamiltonian satisfying the
requirements (13.76, 13.77)?
13.12b. Since the states | ψ n (t)〉 and are both normalized
they could only differ by a time-dependent phase if they are
directly related,
Which conditions would the phase have to fulfill for | ψ n (t)〉 to satisfy the initial value
problem (13.79)?
(13.80)
13.12c. What would be the solution for
if a solution exists?
Which condition does have to satisfy for
existence of ?
Solution for 13.12b
and 13.12c.
Substitution of (13.80) into the
time-dependent Schrödinger equation (13.79) and taking into
account (13.76) yields
i.e. we would need to satisfy the conditions
and to ensure that the Ansatz (13.80) satisfies the
initial value problem (13.79). If the condition (13.81) is consistent, it
is equivalent to
with solution
However, the condition (13.81) will usually not be consistent and will not exist in many cases. The
problem is that condition (13.81) requires that ,
(13.81)
(13.82)
(13.83)
(13.84)
which would yield with (13.83)
with the Berry phase
11
The constraints on the existence of physical states of the
form (13.80) and on the usefulness of the Berry
phase can most easily be seen from the fact that (13.85) is equivalent to
whereas generically time-dependence of the Hamiltonian mixes its
eigenstates under time-evolution,
Stated differently, the condition for existence of is
or the condition for (13.80, 13.83) as an approximate solution of the
time-dependent Schrödinger equation is that for every m ≠ n
Note that if exists, the
solution (13.80, 13.83) can also be written as
Of course, the Berry phase always exists in the sense of the
definition (13.86), and in the same manner one might
simply adopt (13.83) as a definition of . The problem is whether or not they
are related to the evolution of the states | ψ n (t)〉 of the system described by the
Hamiltonian H(t). We have found (13.90) as a condition
for the usefulness of the Berry phase. Comparison
of (13.91) with the exact evolution formula
shows that the condition for usefulness of the Berry phase for the
approximate description of the evolution of the system between
times 0 and t can also be
expressed in the form
for 0 ≤ τ ≤ t or t ≤ τ ≤ 0.
(13.85)
(13.86)
(13.87)
(13.88)
(13.89)
(13.90)
(13.91)
(13.92)
(13.93)
Bibliography
3.
H.A. Bethe, E.E. Salpeter,
Quantum Mechanics of One- and
Two-Electron Atoms (Springer, Berlin, 1957)
30.
J. Orear, A.H. Rosenfeld,
R.A. Schluter, Nuclear Physics: A
Course Given by Enrico Fermi at the University of Chicago
(University of Chicago Press, Chicago, 1950)
Footnotes
3
The transformation law for operators from the
Schrödinger picture into the interaction picture implies
H D (t) ≡ V D (t). The notation V D (t) is therefore also often used for
H D (t).
4
If the perturbation V (t) contains directional information
(e.g. polarization of an incoming photon or the direction of an
electric field), then we might also like to calculate probabilities
for the direction of dissociation of the hydrogen atom. This
direction would be given by the vector of relative motion between
the electron and the proton after separation. For the calculation
of directional information we would have to combine the spherical
Coulomb waves | k, ℓ, m〉 into states which approximate plane
wave states at infinity, similar to
the construction of incoming approximate plane wave states in
Section 13.5, see also the discussion of the photoeffect
in [3].
5
Recall that the notation tacitly implies
dependence of the operators V and W on x and p (just like we usually write
H instead of H(x, p) for a Hamilton operator).
7
See the discussion after
equation (13.38) for an explanation why we can deal with
monochromatic perturbations as abridged non-hermitian
operators.
9
W. Wessel, Annalen Phys. 397, 611 (1930); E.C.G.
Stückelberg, P.M. Morse, Phys. Rev. 36, 16 (1930); M. Stobbe,
Annalen Phys. 399, 661 (1930).
10
Alternatively, we could have used box
normalization for the incoming plane waves,
both in and
in ( ⇒ ),
or we could have rescaled both and
with the conversion factor
8π
3∕V to make both
quantities separately dimensionally correct, ,
.
All three methods yield the same result for the scattering cross
section, of course.