Advanced Modern Physics
Advanced Modern Physics
Preface vii
1. Introduction 1
ix
ADVANCED MODERN PHYSICS - Theoretical Foundations
© World Scientific Publishing Co. Pte. Ltd.
http://www.worldscibooks.com/physics/7555.html
x Advanced Modern Physics
3. Angular Momentum 35
3.1 Translations . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2 Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.3 Angular Momentum Operator . . . . . . . . . . . . . . . . 39
3.4 Eigenvalue Spectrum . . . . . . . . . . . . . . . . . . . . . . 40
3.5 Coupling of Angular Momenta . . . . . . . . . . . . . . . . 46
3.6 Recoupling . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.7 Irreducible Tensor Operators . . . . . . . . . . . . . . . . . 55
3.8 The Wigner-Eckart Theorem . . . . . . . . . . . . . . . . . 57
3.9 Finite Rotations . . . . . . . . . . . . . . . . . . . . . . . . 59
3.9.1 Properties . . . . . . . . . . . . . . . . . . . . . . . 61
3.9.2 Tensor Operators . . . . . . . . . . . . . . . . . . . 63
3.9.3 Wigner-Eckart Theorem (Completed) . . . . . . . . 64
3.10 Tensor Products . . . . . . . . . . . . . . . . . . . . . . . . 65
3.11 Vector Model . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4. Scattering Theory 69
4.1 Interaction Picture . . . . . . . . . . . . . . . . . . . . . . . 69
4.2 Adiabatic Approach . . . . . . . . . . . . . . . . . . . . . . 70
4.3 Û -Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.4 Û -Operator for Finite Times . . . . . . . . . . . . . . . . . 76
4.5 The S-Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.6 Time-Independent Analysis . . . . . . . . . . . . . . . . . . 79
4.7 Scattering State . . . . . . . . . . . . . . . . . . . . . . . . 83
4.8 Transition Rate . . . . . . . . . . . . . . . . . . . . . . . . . 86
4.9 Unitarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
4.10 Example: Potential Scattering . . . . . . . . . . . . . . . . 90
4.10.1 Green’s Function (Propagator) . . . . . . . . . . . . 90
4.10.2 Scattering Wave Function . . . . . . . . . . . . . . . 92
4.10.3 T-matrix . . . . . . . . . . . . . . . . . . . . . . . . 93
4.10.4 Cross Section . . . . . . . . . . . . . . . . . . . . . . 93
4.10.5 Unitarity . . . . . . . . . . . . . . . . . . . . . . . . 94
6. Symmetries 125
6.1 Lorentz Invariance . . . . . . . . . . . . . . . . . . . . . . . 125
6.2 Rotational Invariance . . . . . . . . . . . . . . . . . . . . . 126
6.3 Internal Symmetries . . . . . . . . . . . . . . . . . . . . . . 127
6.3.1 Isospin–SU(2) . . . . . . . . . . . . . . . . . . . . . 127
6.3.1.1 Isovector . . . . . . . . . . . . . . . . . . . 127
6.3.1.2 Isospinor . . . . . . . . . . . . . . . . . . . 131
6.3.1.3 Transformation Law . . . . . . . . . . . . . 133
6.3.2 Lie Groups . . . . . . . . . . . . . . . . . . . . . . . 134
6.3.3 Sakata Model–SU(3) . . . . . . . . . . . . . . . . . . 138
Bibliography 463
Index 467
vii
ADVANCED MODERN PHYSICS - Theoretical Foundations
© World Scientific Publishing Co. Pte. Ltd.
http://www.worldscibooks.com/physics/7555.html
viii Preface
present book is the outcome of those musings. All of the material in this
book is taken from course lectures given over the years by the author at
either Stanford University or the College of William and Mary.
Quantum mechanics is first reformulated in abstract Hilbert space,
which allows one to focus on the general structure of the theory. The book
then covers the following topics: angular momentum, scattering theory, la-
grangian field theory, symmetries, Feynman rules, quantum electrodynam-
ics, path integrals, and canonical transformations for quantum systems.
Several appendices are included with important details. When finished,
the reader should have an elementary working knowledge in the principal
areas of theoretical physics of the twentieth century. With this overview in
hand, development in depth and reach in these areas can then be obtained
from more advanced physics courses.
I was again delighted when World Scientific Publishing Company, which
had done an exceptional job with four of my previous books, showed enthu-
siasm for publishing this new one. I would like to thank Dr. K. K. Phua,
Executive Chairman of World Scientific Publishing Company, and my ed-
itor Ms. Lakshmi Narayanan, for their help and support on this project.
I am greatly indebted to my colleagues Paolo Amore and Alfredo Aranda
for their reading of the manuscript.
Introduction
The goal of this book is to provide an extension of the previous book Intro-
duction to Modern Physics: Theoretical Foundations, refered to as Vol. I.
That volume develops the underlying concepts in twentieth-century physics:
quantum mechanics, special relativity, and general relativity. Included in it
are applications in atomic, nuclear, particle, and condensed matter physics.
It is assumed in Vol. I that readers have had a good calculus-based intro-
ductory physics course together with a good course in calculus. Several
appendices then provide sufficient background so that, with very few ex-
ceptions, the presentation is self-contained. Many of the topics covered in
that work are more advanced than in the usual introductory modern physics
books. It was the author’s intention to provide the best students with an
overview of the subject, so that they are aware of the overall picture and
can see how things fit together as they progress.
As projected in Vol. I, it is now assumed that mathematical skills have
continued to develop. In this volume, readers are expected to be familiar
with multi-variable calculus, in particular, with multiple integrals. It is
also assumed that readers have some familiarity with the essentials of linear
algebra. An appendix is included here on functions of a complex variable,
since complex integration plays a key role in the analysis. The ground
rules now are that anything covered in the text and appendices in Vol. I is
assumed to be mastered, while anything covered in the problems in Vol. I
will be re-summarized. Within this framework, readers should again find
Vol. II to be self-contained.
There are over 175 problems in this book, some after each chapter and
appendix. The problems are not meant to baffle the reader, but rather
to enhance the coverage and to provide exercises on working skills. The
problems for the most part are not difficult, and in most cases the steps
1
ADVANCED MODERN PHYSICS - Theoretical Foundations
© World Scientific Publishing Co. Pte. Ltd.
http://www.worldscibooks.com/physics/7555.html
2 Advanced Modern Physics
are clearly laid out. Those problems that may involve somewhat more
algebra are so noted. The reader is urged to attempt as many problems as
possible in order to obtain some confidence in his or her understanding of
the framework of modern theoretical physics.
In chapter 2 we revisit quantum mechanics and reformulate the the-
ory in terms of linear hermitian operators acting in an abstract Hilbert
space. Once we know how to compute inner products, and have the com-
pleteness relation, we understand the essentials of operating in this space.
The basic elements of measurement theory are also covered. We are then
able to present quantum mechanics in terms of a set of postulates within
this framework. The quantum fields of Vol. I are operators acting in the
abstract many-particle Hilbert space.
Chapter 3 is devoted to the quantum theory of angular momentum, and
this subject is covered in some depth. There are a variety of motivations
here: this theory governs the behavior of any isolated, finite quantum me-
chanical system and lies at the heart of most of the applications in Vol. I;1
it provides a detailed illustration of the consequences of a continuous sym-
metry in quantum mechanics, in this case the very deep symmetry of the
isotropy of space; furthermore, it provides an extensive introduction to the
theory of Lie groups, here the special unitary group in two dimensions
SU(2), which finds wide applicability in internal symmetries. An appendix
explores the use of angular momentum theory in the multipole analysis of
the radiation field, which is applicable to transitions in any finite quantum
mechanical system.
Chapter 4 is devoted to scattering theory. The Schrödinger equation
is solved in terms of a time-development operator in the abstract Hilbert
space, and the scattering operator is identified. The interaction is turned
on and off “adiabatically”, which allows a simple construction of initial and
final states, and the S-matrix elements then follow immediately. Although
inappropriate for developing a covariant scattering analysis, the time inte-
grations in the scattering operator can be explicitly performed and contact
made with time-independent scattering theory. It is shown how adiabatic
damping puts the correct boundary conditions into the propagators. A
general expression is derived for the quantum mechanical transition rate.
Non-relativistic scattering from a static potential provides a nice example
of the time-independent analysis. If the time is left in the scattering op-
erator, one has a basis for the subsequent analysis in terms of Feynman
diagrams and Feynman rules. The tools developed in this chapter allow
one to analyze any scattering or reaction process in quantum mechanics.
Lagrangian field theory provides the dynamical framework for a consis-
tent, covariant, quantum mechanical description of many interacting parti-
cles, and this is the topic in chapter 4. We first review classical lagrangian
particle mechanics, and then classical lagrangian continuum mechanics, us-
ing our paradigm of the transverse planar oscillations of a string. The string
mechanics can be expressed in terms of “two-vectors” (x, ict) where c is the
sound velocity in the string. We then discuss the quantization of these
classical mechanical systems obtained by imposing canonical quantization
relations on the operators in the abstract Hilbert space.
The appending of two additional spatial dimensions to obtain four-
vectors (x, ict), where c is now the speed of light, leads immediately to a
covariant, continuum lagrangian mechanics for a scalar field in Minkowski
space, which is then quantized with the same procedure used for the string.
We develop a covariant, continuum lagrangian mechanics for the Dirac field,
and discuss how anticommutation relations must be imposed when quantiz-
ing in this case. A general expression is derived for the energy-momentum
tensor, and Noether’s theorem is proven, which states that for every contin-
uous symmetry of the lagrangian density there is an associated conserved
current. A full appendix is dedicated to the lagrangian field theory of the
electromagnetic field.
Symmetries play a central role in developing covariant lagrangian densi-
ties for various interacting systems, and chapter 6 is devoted to symmetries.
The discussion starts with spatial rotations and the internal symmetry of
isospin, and it builds on the analysis of SU(2) in chapter 3. Here isospin
is developed in terms of global SU(2) transformations of the nucleon field
ψ = (ψp , ψn ). The internal symmetry is generalized to SU(3) within the
framework of the Sakata model with a baryon field ψ = (ψp , ψn , ψΛ ).2
It is also shown in chapter 6 how the imposition of invariance under local
phase transformations of the charged Dirac field, where the transformation
parameter depends on the space-time point x, necessitates the introduction
of a photon (gauge) field Aµ (x) and leads to quantum electrodynamics
(QED), the most accurate theory known. Yang-Mills theory, which extends
this idea to invariance under local internal symmetry transformations of
the Dirac field, and necessitates the introduction of corresponding gauge
bosons, is developed in detail. These gauge bosons must be massless, and to
2 Wigner’s supermultiplet theory based on internal SU(4) transformations of the nu-
3 Both quantum chromodynamics (QCD) and the Standard Model of electroweak in-
teractions are Yang-Mills theories built on internal symmetry groups, the former on an
N
internal color SU (3)C symmetry and the latter on an internal weak SU (2)W U (1)W .
4 In the Standard Model, it provides the basis for the “Higgs mechanism” (see, for
source and the second a vacuum-vacuum amplitude without it. The con-
nected Green’s functions can then be determined from the generating func-
tional by functional differentiation with respect to the source, as detailed
here. The generating functional is calculated for the free scalar field using
gaussian integration, and it is shown how the Feynman propagator and
Wick’s theorem are reproduced in this case. The treatment of the Dirac
field necessitates the introduction of Grassmann variables, which are anti-
commuting c-numbers. The generating functional is computed for the free
Dirac field, and the Feynman propagator and Wick’s theorem again recov-
ered. It is shown how to include interactions and express the full generating
functional in terms of those already computed.
An appendix describes how one uses the Faddeev-Popov method in a
gauge theory, at least for QED, to factor the measure in the path integral
into one part that is an integral over all gauge functions and a second part
that is gauge invariant. With a gauge-invariant action, the path integral
over the gauge functions then factors and cancels in the generating func-
tional ratio. It is shown how the accompanying Faddeev-Popov determinant
can be expressed in terms of ghost fields, which also factor and disappear
from the generating functional in the case of QED. The generating func-
tional for the free electromagnetic field is calculated here.
Although abbreviated, the discussion in chapter 10 should allow one to
use path integrals with some facility, and to read with some understanding
material that starts from path integrals.
The final chapter 11 deals with canonical transformations for quantum
systems. Chapter 11 of Vol. I provides an introduction to the properties
of superfluid Bose systems and superconducting Fermi systems. In both
cases, in order to obtain a theoretical description of the properties of the
quantum fluids, it is necessary to include interactions. A technique that
has proven invaluable for the treatment of such systems is that of canonical
transformations. Here one makes use of the fact that the properties of
the creation and destruction operators follow entirely from the canonical
(anti)commutation relations in the abstract Hilbert space. By introducing
new “quasiparticle” operators that are linear combinations of the original
operators, and that preserve these (anti)commutation relations, one is able
to obtain exact descriptions of some interacting systems, both in model
problems and in a starting hamiltonian.
The problem of a weakly interacting Bose gas with a repulsive interac-
tion between the particles is solved with the Bogoliubov transformation. A
phonon spectrum is obtained for the many-body system, which, as shown
7 The
author considered also including in Vol. II a chapter on solutions to the Einstein
field equations in general relativity; however, given the existence of [Walecka (2007)], it
was deemed sufficient to simply refer readers to that book.
3
X
v= vi ei ; vi = ei · v (2.2)
i=1
a + b : (a1 + b1 , a2 + b2 , a3 + b3 )
γa : (γa1 , γa2 , γa3 ) ; linear space (2.3)
3
X 3
X
v= vi ei = v̄i αi
i=1 i=1
X3 3
X
⇒ vi = v̄j (ei · αj ) = v̄j [αj ]i (2.6)
j=1 j=1
2.1.2 n-Dimensions
These arguments are readily extended to n-dimensions by simply increasing
the number of components
2.2.1 Example
Recall the set of plane waves in one spatial dimension in an interval of
length L satisfying periodic boundary conditions
1 2πn
φn (x) = √ eikn x ; kn = ; n = 0, ±1, ±2, · · · ;
L L
basis vectors (2.9)
These will be referred to as the basis vectors. They are orthonormal and
satisfy
Z L
dx φ⋆m (x)φn (x) = δmn ; orthonormal
0
≡ hφm |φn i ; inner product (2.10)
This relation allows us to define the inner product of two basis vectors,
denoted in the second line by hφm |φn i,2 and the positive-definite inner-
product norm of the basis vectors is then given by
Z L
2
|φn | = hφn |φn i = dx |φn (x)|2 ; (“length”)2 (2.11)
0
This is, after all, just a complex Fourier series. The orthonormality of the
basis vectors allows one to solve for the coefficients cn
Z L
cn = hφn |ψi = dx φ⋆n (x)ψ(x) (2.13)
0
2 The notation, and most of the analysis in this chapter, is due to Dirac [Dirac (1947)].
Just as with an ordinary vector, the function ψ(x) can now be characterized
by the expansion coefficients cn
2.2.2 Definition
The function ψ(x) in Eq. (2.17) is said to be square-integrable. The set of
all square-integrable functions (L2 ) forms a Hilbert space. Mathematicians
define a Hilbert space as follows:
The sum over the continuous index has here been appropriately defined
through a familiar integral. With this notation, the starting expansion in
Eq. (2.12) takes the form
∞
X
ψx = cn [φn ]x (2.20)
n=−∞
X
|ψi = cn |φn i ; abstract vector relation (2.21)
n
As before, one solves for the expansion coefficients cn by simply using the
orthonormality of the basis vectors in Eq. (2.10)
X Z
cn = hφn |ψi = [φn ]⋆x ψx = dx φ⋆n (x)ψ(x) (2.22)
x
hermitian (2.24)
We now make the important observation that if one knows the matrix
elements of L
Z
Lmn ≡ dx φ⋆m (x)Lφn (x) ≡ hm|L|ni ; matrix elements (2.26)
in any complete basis, then one knows the operator L. Let us prove this
assertion. Let ψ(x) be an arbitrary state in the space. If one knows the
corresponding Lψ(x), then L is determined. Expand ψ(x) in the complete
basis
X
ψ(x) = cn φn (x) ; complete basis (2.27)
n
These coefficients are thus determined for any given ψ. Now compute5
X
Lψ(x) = cn [Lφn (x)] (2.29)
n
4 SeeProbI. 4.5—the notation “ProbI” refers to the problems in Vol. I.
5 It
is assumed here that there is enough convergence that one can operate on this
series term by term.
The expansion in a complete basis can again be invoked to write the state
Lφn (x) as
X
Lφn (x) = βmn φm (x) ; complete basis (2.30)
m
Hence
X
Lφn (x) = φm (x)Lmn
m
XX
⇒ Lψ(x) = φm (x)Lmn cn ; known (2.32)
n m
This is now a known quantity, and thus we have established the equivalence6
2.3.1 Eigenfunctions
The eigenfunctions of a linear operator are defined by the relation
For the present purposes, one can simply take two of the postulates of
quantum mechanics to be:
[Fetter and Walecka (2003)] . The use of ordinary riemannian integration in the definition
of the inner product in Eq. (2.19), and the notion of completeness expressed in Eq. (2.14),
represent the extent of the mathematical rigor in the present discussion.
On the interval [0, L], with periodic boundary conditions, the eigenvalues
ξ run continuously over this interval. As to the orthonormality of these
eigenfunctions, one can just compute
Z Z
dx ψξ′ (x)ψξ (x) = dx δ(x − ξ ′ )δ(x − ξ) = δ(ξ − ξ ′ )
⋆
(2.43)
Hence
Z
dx ψξ⋆′ (x)ψξ (x) = δ(ξ − ξ ′ ) ; orthonormality (2.44)
Recall Eqs. (2.12) and (2.20) from above, which represent an expansion in
a complete set,
X
ψ(x) = cn φn (x)
n
X
or ; ψx = cn [φn ]x (2.47)
n
This can be viewed as the component form of the abstract vector relation
X
|ψi = cn |φn i ; abstract vector relation (2.48)
n
Z
hk |ki = dx φ⋆k′ (x)φk (x) = δkk′
′
; with p.b.c.
1
Z
hξ|ki = dx ψξ⋆ (x)φk (x) = √ eikξ (2.51)
L
The last relation follows directly from the wave functions in Eqs. (2.36) and
(2.41).10
10 See also Eq. (2.9); note that the subscript n on kn = 2πn/L is suppressed.
2.4.2 Completeness
As established in Vol. I, the statement of completeness with the set of
coordinate space eigenfunctions φp (x), where p denotes the eigenvalues of
a linear hermitian operator, is
X
φp (x)φ⋆p (y) = δ(x − y) ; completeness (2.52)
p
Insert this relation in the definition of the inner product in Eq. (2.49)
Z Z
hψa |ψb i = dx ψa⋆ (x)ψb (x) ≡ dxdy ψa⋆ (x)δ(x − y)ψb (y)
XZ Z
= dx ψa⋆ (x)φp (x) dy φ⋆p (y)ψb (y)
p
X
= hψa |φp ihφp |ψb i (2.53)
p
Here Eq. (2.52) has been used in the second line, and the definition of
the inner product used in the third. This relation can be summarized by
writing the abstract vector relation
X
|φp ihφp | = 1op ; completeness (2.54)
p
This unit operator 1op can be inserted into any inner product, leaving that
inner product unchanged. This relation follows from the completeness of
the wave functions φp (x) providing the coordinate space components of the
abstract state vectors |φp i.
2.4.3.1 Eigenstates
A linear hermitian operator Lop takes one abstract vector |ψi into another
Lop |ψi. The eigenstates of Lop , as before, are defined by
For example:
The adjoint operator in the abstract Hilbert space is defined in exactly the
same manner
hψa |L†op |ψb i ≡ hLop ψa |ψb i = hψb |Lop |ψa i⋆ ; adjoint (2.58)
Note that it follows from this definition that if γ is some complex number,
then
With an hermitian operator, one can just let it act on the state on the left
when calculating matrix elements.
(3) Similarly, compute the matrix elements of the kinetic energy Top . This
is readily accomplished by invoking the completeness relation for the
eigenstates of momentum [see Eq. (2.54)]
X
|kihk| = 1op ; completeness (2.63)
k
1 1 XX
hξ|Top |ξ ′ i = hξ|p2op |ξ ′ i = hξ|kihk|p2op |k ′ ihk ′ |ξ ′ i
2m 2m ′k k
2
~ XX ~2 X k 2 ik(ξ−ξ′ )
= hξ|kik 2 δkk′ hk ′ |ξ ′ i = e
2m 2m L
′ k k k
2 2
~ ∂ X 1 ik(ξ−ξ′ ) ~2 ∂ 2
=− 2
e =− 2
δ(ξ − ξ ′ ) (2.64)
2m ∂ξ L 2m ∂ξ
k
The final relation follows from the completeness of the momentum wave
functions.
(4) Make use of the statement of completeness of the abstract eigenstates
of position, which is
Z
dξ |ξihξ| = 1op ; completeness (2.65)
Note that the sum here is actually an integral because the position
eigenvalues are continuous.11
11 See Prob. 2.2.
Now insert Eq. (2.65) in the expression on the l.h.s., and use the results
from Eqs. (2.62) and (2.64)
Z
hξ|Hop |ψi = dξ ′ hξ|Hop |ξ ′ ihξ ′ |ψi
~2 ∂ 2
Z
′
= dξ − + V (ξ) δ(ξ − ξ ′ )ψ(ξ ′ )
2m ∂ξ 2
~2 ∂ 2
Z
= − + V (ξ) dξ ′ δ(ξ − ξ ′ )ψ(ξ ′ )
2m ∂ξ 2
~2 ∂ 2
= − + V (ξ) ψ(ξ) (2.68)
2m ∂ξ 2
Thus, in summary,
~2 ∂ 2
− + V (ξ) ψ(ξ) = E ψ(ξ) ; S-equation (2.69)
2m ∂ξ 2
∂
i~ |Ψ(t)i = Ĥ|Ψ(t)i ; S-equation
∂t
~
[p̂, x̂] = ; C.C.R. (2.70)
i
We make several comments:
12 The time-independent Schrödinger equation in the momentum representation is ob-
tained by projecting Eq. (2.66) onto the states |ki. This gives the components of the
operator relation in a basis of eigenstates of momentum (see Prob. 2.9).
2.5 Measurements
n n′
XX
= a⋆fn (t)afn′ (t)fn δnn′ (2.76)
n n′
Hence
X
hF i = |afn (t)|2 fn ; expectation value (2.77)
n
then the wave function ψ(x) can be expanded in the ufn (x) with time-
independent coefficients afn
X
ψ(x) = afn ufn (x) ; completeness (2.79)
n
hF i = fn ; in eigenstate (2.81)
Equations (2.77) and (2.75) suggest that one should interpret the quan-
tity |afn (t)|2 as the probability of measuring the value fn at the time t if
a system is in the state Ψ(x, t). Based on this argument, we make the
following measurement postulates:
(1) If one makes a precise measurement of F , then one must observe one
of the eigenvalues fn ;
(2) If one is in an arbitrary state Ψ(x, t), then |afn (t)|2 is the probability
that one will observe the value fn for F at the time t, where17
Z
afn (t) = dx u⋆fn (x)Ψ(x, t) (2.82)
If one measures F at the time t0 and finds a value fn , then right after this
measurement, the wave function must be such as to again give the value
fn , and it must be normalized. Thus, with no degeneracy, the effect of this
measurement is to reduce the wave function to the form18
afn (t0 )
Ψ(x, t0 )′ = uf (x) (2.90)
|afn (t0 )| n
This result can be abstracted and extended to lead to an additional mea-
surement postulate:
18 We speak here of “pure pass measurements” that do not modify the coefficients
afn (t).
(3) If, at the time t0 , one observes a value f for the quantity F which lies
in the inteval f ′ ≤ f ≤ f ′′ , then the state vector is reduced to
P′
af (t0 )|fn i
|Ψ(t0 )i′ = P n n 1/2 ; where f ′ ≤ fn ≤ f ′′ (2.91)
′ 2
n |afn (t0 )|
P′
Here n implies f ′ ≤ fn ≤ f ′′ .
Although this postulate may at first seem very mysterious, a little re-
flection will convince the reader that a measurement does indeed provide
a great deal of information about a system, in particular, this type of in-
formation. We briefly discuss, as an example, the classic Stern-Gerlach
experiment.
If the center-of-mass of the atom goes through the top slit (this will
happen with probability |c+1 (t0 )|2 where t0 is the time it goes through
the magnet), then the internal wave function of the atom must be20
c+1 (t0 )
ψint (x, t) = Rnp (r)Y11 (θ, φ)e−iEnp (t−t0 )/~ (2.94)
|c+1 (t0 )|
B m=+1
z
m=0
oven m=-1
d Bz
inhomogeneous dz
magnet (A)
Fig. 2.1 Sketch of the Stern-Gerlach experiment. We will refer to the entire boxed unit
as detector (A).
m=+1
m=+1
(A)
Fig. 2.2 Detector (A) placed after the upper beam with m = +1 in Fig. 2.1.
• If one looks for a beam emerging from the middle and bottom slits of
the second detector, there will be none. This illustrates the reduction
of the wave packet by the first measurement.
(1) There is a state vector |Ψ(t)i that provides a complete dynamical de-
scription of a system;
(2) An observable F is represented by a linear hermitian operator F̂ ;
(3) The operators obey canonical commutation relations, in particular
~
[p̂, x̂] = (2.95)
i
(4) The dynamics is given by the Schrödinger equation
∂
i~ |Ψ(t)i = Ĥ|Ψ(t)i (2.96)
∂t
(5) The eigenstates of a linear hermitian operator form a complete set
X
F̂ |fn i = fn |fn i ; |fn ihfn | = 1̂ (2.97)
n
Through his many years in physics, the author has found this to be a
complete and essential set of postulates for the implementation of quantum
mechanics.
[a, a† ] = 1 (2.99)
N̂ ≡ a† a ; number operator
Ĥ = ~ω(N̂ + 1/2) ; hamiltonian (2.100)
The second line expresses the hamiltonian in terms of the number operator.
As demonstrated in ProbsI. 4.17–4.18, it follows entirely from the general
properties of the linear hermitian operators involved that the spectrum of
the number operator consists of the positive integers and zero
N̂ |ni = n|ni ; n = 0, 1, 2, · · · , ∞
N̂ |0i = 0 ; ground state (2.101)
The last relation defines the ground state. It further follows that
√
a|ni = n |n − 1i ; destruction operator
√
a† |ni = n + 1 |n + 1i ; creation operator (2.102)
21 The reader is again strongly urged to work through those problems (see Prob. 2.1).
22 We suppress the carets on the creation and destruction operators, since it will hence-
forth be obvious that they act in the abstract occupation-number Hilbert space.
2.7.2 Bosons
With many identical bosons, one introduces a set of creation and destruc-
tion operators satisfying
(1) The normal modes for the transverse oscillations of a continuous string
of length L with periodic boundary conditions are given by23
1 2πm
φk (x) = √ eikx ; k= ; m = 0, ±1, ±2, · · · (2.107)
L L
23 We again suppress the subscript m on km = 2πm/L.
The string energy, which plays the role of free-field hamiltonian, is then
found in terms of the quantum field of the string q̂(x, t), obtained from
its classical transverse displacement, and the corresponding quantum
momentum density π̂(x, t) = σ∂ q̂(x, t)/∂t, obtained from its classical
transverse motion
X ~ 1/2 h i
q̂(x, t) = ak ei(kx−ωk t) + a†k e−i(kx−ωk t) ; ωk = |k|c
2ωk σL
k
1/2 h
1 X ~ωk σ i
π̂(x, t) = ak ei(kx−ωk t) − a†k e−i(kx−ωk t) (2.108)
i 2L
k
Here σ is the mass density, and c is the sound velocity. These operators,
which here carry the free-field time dependence, satisfy the canonical
equal-time commutation relations
2.7.3 Fermions
In the case of fermions, in order to satisfy the Pauli exclusion principle, one
quantizes with anticommutation relations instead of commutation relations.
26 Thisis called “second quantization”, since what were previously single-particle wave
functions now become field operators in the abstract many-particle Hilbert space. The
formulation of the many-body problem in second quantization is carried out in detail in
chapter 1 of [Fetter and Walecka (2003a)] .
27 The spinors with the same coordinate label are to be paired in Eq. (2.120).
Scattering Theory
Given a hamiltonian Ĥ, the goal of this chapter is to solve the Schrödinger
equation for a scattering problem and derive general expressions for the
S-matrix, T -matrix, and transition rate, many of whose consequences have
already been examined in Vol. I.1 We work in the abstract Hilbert space.
Assume the hamiltonian can be split into two parts Ĥ = Ĥ0 + Ĥ1 , the
first part of which leads to an exactly solvable problem, for example, free
quanta with no interactions. Ĥ1 may, or may not, have an explicit time
dependence; that depends on the problem at hand.2 We then want to solve
the Schrödinger equation
Ĥ = Ĥ0 + Ĥ1
∂
i~ |Ψ(t)i = Ĥ |Ψ(t)i ; Schrödinger-equation (4.1)
∂t
i
|ΨI (t)i ≡ e ~ Ĥ0 t |Ψ(t)i ; interaction picture
|ΨI (0)i = |Ψ(0)i ; coincide at t = 0 (4.2)
69
What equation of motion does this new state satisfy? Just compute
∂ i i ∂
i~ |ΨI (t)i = −Ĥ0 e ~ Ĥ0 t |Ψ(t)i + e ~ Ĥ0 t i~ |Ψ(t)i
∂t ∂t
i i
Ĥ t
= −Ĥ0 |ΨI (t)i + e ~ 0
(Ĥ0 + Ĥ1 )e− ~ Ĥ0 t |ΨI (t)i (4.3)
The advantage of this new formulation is that in the limit Ĥ1 → 0, the
state |ΨI (t)i becomes time-independent; the free time variation, which can
be extremely rapid, has been explicitly dealt with. Equations (4.2) and
(4.4) are said to be a formulation of the problem in the interaction picture.
We will find that when we try to solve the resulting equations and generate
the S-matrix, there will be infinite time integrals to carry out over oscil-
lating integrands. In order to give the theory a well-defined mathematical
meaning, we introduce an adiabatic damping factor e−ǫ|t| with ǫ ≥ 0, and
use the following interaction in the interaction picture
the hamiltonian simply reduces to Ĥ0 , and we know how to solve the non-
interacting problem
Ĥ = Ĥ0 ; t → ±∞
∂
i~ |Ψ(t)i = Ĥ0 |Ψ(t)i
∂t
i
|Ψ(t)i = e− ~ E0 t |ψi (4.6)
The interaction-picture state vector in Eq. (4.2) is then given in this same
limit by
i
|ΨI (t)i = e ~ Ĥ0 t |Ψ(t)i = |ψi ; t → ±∞
∂
i~ |ΨI (t)i = 0 (4.8)
∂t
One starts with an initial state of this type, and then slowly turns on and
off the interaction. The transition amplitude into a final state of this type is
then calculated. The (transition probability)/(time interval the interaction
is on) gives the transition rate,5 and the path from the transition rate to a
cross section was detailed in Vol. I.
It is then necessary to determine what happens when the interaction
in Eq. (4.5) is turned on and off adiabatically. This is done through the
construction of the time-development operator for the problem.
5 We shall get more sophisticated here and actually derive a general expression for the
transition rate itself.
4.3 Û -Operator
If this is to hold for all |ΨI (t0 )i, then Ûǫ (t, t0 ) must satisfy the operator
relation
∂
i~ Ûǫ (t, t0 ) = ĤIǫ (t)Ûǫ (t, t0 )
∂t
Ûǫ (t0 , t0 ) = 1 (4.11)
t
i
Z
′
Ûǫ (t, t0 ) = 1 − e−ǫ|t | ĤI (t′ )Ûǫ (t′ , t0 ) dt′ (4.12)
~ t0
i t −ǫ|t′ |
Z
Ûǫ (t, t0 ) = 1 − e ĤI (t′ ) dt′ +
~ t0
2 Z t Z t′
i −ǫ|t′ | ′ ′ ′′
− e ĤI (t ) dt e−ǫ|t | ĤI (t′′ )Ûǫ (t′′ , t0 ) dt′′ (4.13)
~ t0 t0
∞ n Z t Z t1 Z tn−1
X i −ǫ|t1 | −ǫ|t2 |
Ûǫ (t, t0 ) = − e dt1 e dt2 · · · e−ǫ|tn | dtn ×
n=0
~ t0 t0 t0
Here
Rt
• All the integrals are now over the full range t0 ;
• The “T-product” carries the instruction that the operators are to be
time-ordered, with the operator at the latest time sitting to the left;
• Each term in the sum is divided by n!.
The proof that Eq. (4.15) reproduces Eq. (4.14) is quite simple. There
are n! possible orderings of the times in the multiple integral, pick one, say
t1 > t2 > t3 > · · · > tn . All possible time orderings of these integration
variables provides a complete enumeration of the region of integration in
the multiple integral. The operator in the integrand is time-ordered in each
case. But now all of these contributions are identical by a change of dummy
integration variables. Thus Eq. (4.14) is reproduced.7
The scattering operator Ŝ is now defined in the following manner
One lets the initial time t0 → −∞, the final time t → +∞, and then, at the
very end, the limit of the adiabatic damping factor ǫ → 0 is taken. Thus
Ŝ = Limǫ→0 Ŝǫ
∞ n Z ∞ Z ∞
X i 1 −ǫ|t1 |
= Limǫ→0 − e dt1 · · · e−ǫ|tn | dtn ×
n=0
~ n! −∞ −∞
h i
T ĤI (t1 )ĤI (t2 ) · · · ĤI (tn ) (4.17)
How is the above analysis modified? Write Eq. (4.12) in the following
fashion
i t0 −ǫ|t′ |
Z
Ûǫ (t, t0 ) = 1 + e ĤI (t′ )Ûǫ (t′ , t0 ) dt′ (4.18)
~ t
It is readily verified that this expression reproduces Eqs. (4.11), and it
is most convenient since the integral now runs in the positive direction if
t0 ≥ t. A repetition of the above arguments in this case then leads to the
following infinite series
∞ n
1 t0 −ǫ|t1 |
Z t0 Z t0
i
X Z
−ǫ|t2 |
Ûǫ (t, t0 ) = e dt1 e dt2 · · · e−ǫ|tn | dtn ×
n=0
~ n! t t t
h i
T ĤI (t1 )ĤI (t2 ) · · · ĤI (tn ) ; t ≤ t0 (4.19)
This follows from the series expansions, and the explicit demonstration of
this relation for n = 2 is left as Prob. 4.1.
(3) It follows from the results in (1) and (2) that
hΨI (t)|ΨI (t)i = hΨI (t0 )|Ûǫ (t, t0 )† Ûǫ (t, t0 )|ΨI (t0 )i
= hΨI (t0 )|Ûǫ (t, t0 )−1 Ûǫ (t, t0 )|ΨI (t0 )i
= hΨI (t0 )|ΨI (t0 )i (4.24)
Let us demonstrate this result for t2 > t1 > t0 . The result in (2) can then
be used to extend it to any relative times. For example, if t1 > t2 , just
write
Ûǫ (t2 , t1 )Ûǫ (t1 , t0 ) = Ûǫ (t2 , t1 )Ûǫ (t1 , t2 )Ûǫ (t2 , t0 )
= Ûǫ (t2 , t0 ) ; t1 > t2 (4.26)
Write out the νth term in the sum on the r.h.s. of Eq. (4.25)
ν
1 t2 −ǫ|t′1 | ′
Z t2
i
Z
′
Ûǫ(ν) (t2 , t0 ) = − e dt1 · · · e−ǫ|tν | dt′ν ×
~ ν! t0 t0
h i
T ĤI (t′1 )ĤI (t′2 ) · · · ĤI (t′ν ) (4.27)
Now note:
• There are ν!/n!m! ways to partition the times t′1 · · · t′ν so that n times
are greater than the intermediate time t1 , and m times are less than t1
— pick one;
• Now integrate over all possible relative orderings of the times within
this particular partition;
• Then sum over all possible choices of the times within this particu-
lar partition. This provides a complete enumeration of the regions of
integration for a given (n, m) ;
• The contributions in the sum are identical by a change of dummy in-
tegration variables, giving ν!/n!m! equal contributions;
• Then sum over all values of (n, m) for which m + n = ν. This provides
a complete evaluation of the multiple integral in Eq. (4.27)
n+m
1 X i ν!
Ûǫ(ν) (t2 , t0 )= − × (4.28)
ν! n+m=ν ~ n!m!
Z t2 Z t2 h i
−ǫ|t′1 | ′ ′
e dt1 · · · e−ǫ|tn| dt′n T ĤI (t′1 ) · · · ĤI (t′n ) ×
t1 t1
Z t1 Z t1 h i
−ǫ|t′n+1 | ′
e dt′n+1 ··· e−ǫ|tn+m| dt′n+m T ĤI (t′n+1 ) · · · ĤI (t′n+m )
t0 t0
P P P P
• Finally, use ν n+m=ν = n m. This establishes Eq. (4.25).
In the end, we are to take the limit ǫ → 0, which restores the proper
hamiltonian. Let us assume that we have used the preceding analysis to
propagate the system from its initial state at t0 → −∞ to a finite time such
that
In this case, we can write a formal solution to the full Schrödinger equation
as9
i
|Ψi (t)i = e− ~ Ĥt |Ψi (0)i (4.32)
Here |Ψi (0)i = |ΨiI (0)i is the state that has propagated up to the time t = 0
from the initial state |ψi i prepared at t0 → −∞ [see Eqs. (4.2)]. With the
aid of the previous time-evolution operator, one can write this state as
(+)
|Ψi (0)i = |ΨiI (0)i = Ûǫ (0, −∞)|ψi i ≡ |ψi i (4.33)
9 We assume here and henceforth that Ĥ1 now has no explicit time dependence.
(+)
This relation defines |ψi i. A combination of Eqs. (4.32) and (4.33) allows
the solution to the Schrödinger equation at a finite time, which satisfies
Eq. (4.30), to be expressed as
i (+)
|Ψi (t)i = e− ~ Ĥt |ψi i (4.34)
Some comments:
• This is the full Schrödinger state vector that develops from the state
|ψi i at t0 → −∞ ;
• One needs the adiabatic damping factor to bring that state vector up
(+)
to finite time with |Ψi (0)i = Ûǫ (0, −∞)|ψi i ≡ |ψi i ;
• From there, one can use the formal solution to the full Schrödinger
equation in Eq. (4.34).
We note that under the conditions that one can indeed use the formal
solution to the full Schrödinger equation, it follows that the interaction-
picture state vector at the time t is given by
i i i
|ΨI (t)i = e ~ Ĥ0 t |Ψ(t)i = e ~ Ĥ0 t e− ~ Ĥ(t−t0 ) |Ψ(t0 )i
i i i
= e ~ Ĥ0 t e− ~ Ĥ(t−t0 ) e− ~ Ĥ0 t0 |ΨI (t0 )i (4.35)
Here |ΨI (t0 )i is the interaction-picture state vector at the time t0 . But now
we can immediately identify the time development operator Û (t, t0 ) from
the first of Eqs. (4.10)!
i i i
Û(t, t0 ) = e ~ Ĥ0 t e− ~ Ĥ(t−t0 ) e− ~ Ĥ0 t0 ; |t|, |t0 | ≪ 1/ǫ (4.36)
Û (t, t0 )† = Û(t0 , t)
Û (t, t0 )† = Û(t, t0 )−1 ; unitary
Û (t1 , t2 )Û (t2 , t3 ) = Û(t1 , t3 ) ; group property (4.37)
The interaction-picture state vector in the infinite future |ΨI (+∞)i that
develops from the interaction-picture state vector in the infinite past
Now, with the adiabatic damping factor, the interaction state vectors in
the infinite past and infinite future are simple, they are just the individual
non-interacting state vectors in Eq. (4.9), or linear combinations of them.
Thus, if one starts with one such prepared state |ΨiI (−∞)i = |ψi i, and
asks for the probability for finding a particular state |ψf i in the final state
|ΨiI (+∞)i that evolves, in the presence of all the interactions, from that
initial prepared state, one has
Pf i = |hψf |ΨiI (+∞)i|2 = |hψf |Ŝ |ΨiI (−∞)i|2 = |hψf |Ŝ|ψi i|2 (4.39)
This is the probability of finding the initial state |ψi i in the final state |ψf i
after the scattering has taken place. Here |ψi i and |ψf i are eigenstates of
the free hamiltonian Ĥ0 . The amplitude for this process to take place is
given by the S-matrix
It was argued in Vol. I that the general form of the S-matrix for a
scattering process is
where T → ∞ is the total time the interaction is turned on. The transition
rate is then given by
Pf i
ωf i =
T
2π
ωf i = δ(Ef − Ei )|T̃f i |2 ; transition rate (f 6= i) (4.44)
~
This is the transition rate into one final state in the continuum. To get the
transition rate into the group of states that actually get into our detectors
when the states are spaced very close together, one must multiply this
expression by the appropriate number of states dnf . To get a cross section,
one divides by the incident flux
2π dnf
dσ = δ(Ef − Ei )|T̃f i |2 ; cross section (4.45)
~ Iinc
Some comments:
Sf i = δf i − 2πiδ(Ef − Ei ) T̃f i
2 2π
ωf i = δf i Im T̃ii + δ(Ef − Ei )|T̃f i |2 ; transition rate (4.46)
~ ~
We will now perform some formal manipulations on the above results. Let
us try to explicitly carry out the time integrations in the general term in the
S-matrix in Eq. (4.17), which we rewrite in its initial time-ordered form
n Z ∞ Z t1 Z tn−1
(n) i −ǫ|t1 | −ǫ|t2 |
hψf |Ŝǫ |ψi i = − e dt1 e dt2 · · · e−ǫ|tn | dtn
~ −∞ −∞ −∞
i i i i
×hψf |e ~ Ĥ0 t1 Ĥ1 e− ~ Ĥ0 t1 e ~ Ĥ0 t2 Ĥ1 e− ~ Ĥ0 t2 · · ·
i i i
· · · Ĥ1 e− ~ Ĥ0 tn−1 e ~ Ĥ0 tn Ĥ1 e− ~ Ĥ0 tn |ψi i (4.47)
Here we have simply written out hψf |ĤI (t1 ) · · · ĤI (tn )|ψi i in detail.
We will change variables in the integrals as follows
x1 = t1 ; t1 = x1
x2 = t2 − t1 ; t2 = x1 + x2
x3 = t3 − t2 ; t3 = x1 + x2 + x3
.. ..
. .
xn = tn − tn−1 ; tn = x1 + x2 + · · · + xn (4.48)
First, let the hamiltonians Ĥ0 on either end of the operator in Eq. (4.47)
act on |ψi i and |ψf i, which are eigenstates of Ĥ0 with eigenvalues E0 and
Ef respectively. Equation (4.47) then can be written as
n Z ∞ Z t1 Z tn−1
(n) i −ǫ|t1 | −ǫ|t2 |
hψf |Ŝǫ |ψi i = − e dt1 e dt2 · · · e−ǫ|tn | dtn
~ −∞ −∞ −∞
i i i i i
×hψf |e ~ (Ef −E0 )t1 Ĥ1 e− ~ Ĥ0 (t1 −t2 ) e ~ E0 (t1 −t2 ) Ĥ1 e− ~ Ĥ0 (t2 −t3 ) e ~ E0 (t2 −t3 ) · · ·
i i
· · · e− ~ Ĥ0 (tn−1 −tn ) e ~ E0 (tn−1 −tn ) Ĥ1 |ψi i (4.49)
Next, introduce the change in variables in Eqs. (4.48), starting from the
right
n Z ∞
(n) i i
hψf |Ŝǫ |ψi i = − e ~ (Ef −E0 )x1 e−ǫ|x1 | dx1 ×
~ −∞
Z 0
i
hψf |Ĥ1 dx2 e{− ~ (E0 −Ĥ0 )x2 −ǫ|x1 +x2 |} Ĥ1 ×
−∞
Z 0
i
dx3 e{− ~ (E0 −Ĥ0 )x3 −ǫ|x1 +x2 +x3 |} Ĥ1 × · · ·
−∞
Z 0
i
· · · Ĥ1 dxn e{− ~ (E0 −Ĥ0 )xn −ǫ|x1 +···+xn |} Ĥ1 |ψi i (4.50)
−∞
Now do all the integrals starting on the right, keeping all the other variables
fixed while so doing.
Consider the first integral over dxn at fixed (x1 , · · · , xn−1 ). What we
(n)
really need is Limǫ→0 hψf |Ŝǫ |ψi i. Since the damping factors are just there
to cut off the oscillating exponentials, we should get the same results no
matter how we go to that limit, if the theory is to make sense. We claim
.
that in the limit, we can replace e−ǫ|x1 +···+xn | = eǫxn in the integral over
xn , since it is only important for very large negative xn . Repetition of this
The integrals now factor, and they can all be immediately carried out
n
(n) i
hψf |Ŝǫ |ψi i = − 2π~ δ(Ef − E0 ) ×
~
1 1
hψf |Ĥ1 Ĥ1 Ĥ1 · · ·
−i(E0 − Ĥ0 )/~ + ǫ −i(E0 − Ĥ0 )/~ + ǫ
1
··· Ĥ1 |ψi i
−i(E0 − Ĥ0 )/~ + ǫ
(4.52)
The operator Ĥ1 appears n times in this expression. This equation has
meaning in terms of a complete set of eigenstates of Ĥ0 inserted be-
tween each term. With the redefinition ǫ~ ≡ ε, one arrives at the time-
independent power series expansion of the S-matrix
Several comments:
• The T -matrix can now be identified from Eqs. (4.41) and (4.53)
explicitly covariant. Infinities arise from various sources, which are not
interpretable in a non-covariant approach. We will find that by leaving
the time integrations in, and starting from Eq. (4.17), we are able to
maintain a covariant, gauge-invariant S-matrix, which proves essential
to developing a consistent renormalization scheme.13
It follows from Eq. (4.32) that the Heisenberg state vector is independent
of time14
∂
i~ |ΨH i = 0 (4.61)
∂t
The interaction-picture state vector is defined in Eq. (4.2). The state vec-
tors in all the different pictures coincide at t = 0
The nth order contribution to Ûε (0, −∞) explicitly contains n powers
of Ĥ1
n Z 0 Z t1 Z tn−1
(n) i ǫt1 ǫt2
Ûǫ (0, −∞)|ψi i = − e dt1 e dt2 · · · eǫtn dtn ×
~ −∞ −∞ −∞
i i i i i
e ~ Ĥ0 t1 Ĥ1 e− ~ Ĥ0 (t1 −t2 ) Ĥ1 e− ~ Ĥ0 (t2 −t3 ) · · · e− ~ Ĥ0 (tn−1 −tn ) Ĥ1 e− ~ Ĥ0 tn |ψi i
(4.64)
13 Seethe discussion in Vol. I.
14 We remind the reader of the assumption, at this point, that Ĥ has no explicit time
dependence.
In comparing with our starting point in Eq. (4.47) from which we proceeded
to explicitly carrying out the time integrations, we note two differences:
• All the times satisfy t ≤ 0, hence the adiabatic damping factors in all
cases become e−ǫ|t| = eǫt ;
• There is no eigenstate |ψf i on the left, and hence the operator Ĥ0 on
the left can no longer be replaced by its eigenvalue Ef .
All the integrals now explicitly factor, and they can immediately be done
just as before with the result
1 1
Ûε(n) (0, −∞)|ψi i = Ĥ1 Ĥ1 · · ·
E0 − Ĥ0 + inε E0 − Ĥ0 + i(n − 1)ε
1
··· Ĥ1 |ψi i (4.66)
E0 − Ĥ0 + iε
Again, we are interested in the limit as ε → 0. Each of the in̄ε in the
denominators, where n̄ = (1, 2, · · · , n), simply serves to define how one
treats the singularity in the individual Green’s functions.15 Hence, we can
simply replace them all by iε in the limit. Thus we indeed reproduce the
previously employed expression in Eq. (4.56)
(+)
|ψi i = Ûε (0, −∞)|ψi i
∞ n
X 1
= Ĥ1 |ψi i ; scattering state (4.67)
n=0 E0 − Ĥ0 + iε
Again, by separating out the first term in the second line, and then re-
15 They serve to define a contour in the evaluation of the Green’s functions (see later).
(+)
identifying the series for |ψi i, this can be rewritten as an integral equation
(+) 1 (+)
|ψi i = |ψi i + Ĥ1 |ψi i (4.68)
E0 − Ĥ0 + iε
and the integral equation has a meaning, even when the power series solu-
tion to it does not.
(−)
Let us also consider the fully interacting state |ψf i ≡ |ΨfI (0)i that as
(−)
t → +∞ reduces to the state |ψf i, so that |ψf i = Ûǫ (0, +∞)|ψf i. If we go
back to Eq. (4.19), and go through the arguments leading from Eq. (4.64)
to (4.67), we see that the only changes are the replacements E0 → Ef and
ε → −ε (see Prob. 4.3). Thus
(−)
|ψf i ≡ Ûε (0, +∞)|ψf i
∞
!n
X 1
= Ĥ1 |ψf i ; scattering state (4.69)
n=0 Ef − Ĥ0 − iε
This can again be written as an integral equation, which has meaning even
when the power series solution for it does not
(−) 1 (−)
|ψf i = |ψf i + Ĥ1 |ψf i (4.70)
Ef − Ĥ0 − iε
The state |ψ (+) i is known as the outgoing scattering state, and |ψ (−) i as
the incoming scattering state.16
There are some important properties of these scattering states that fol-
low immediately:
Similarly17
(−) (−)
hψf ′ |ψf i = δf ′ f (4.72)
16 The Green’s function in the former case has outgoing scattered waves, while in the
latter case they are incoming [compare Eq. (4.107) and Prob. 4.9].
17 The completeness relation can also be written
P (−) (−) P
f |ψf ihψf | + bnd states
|ψb ihψb | = 1̂.
(2) Furthermore, from Eq. (4.21) and the group property of Ûε , it follows
that
(−) (+)
hψf |ψi i = hψf |Ûε (0, +∞)† Ûε (0, −∞)|ψi i
= hψf |Ûε (+∞, 0)Ûε (0, −∞)|ψi i
= hψf |Ûε (+∞, −∞)|ψi i (4.73)
(−) (+)
Thus the inner product of |ψf i and |ψi i is just the S-matrix!
(−) (+)
hψf |ψi i = hψf |Ŝ|ψi i ; S-matrix (4.74)
(3) Since taking the adjoint merely reverses the order of the operators and
changes the sign of the iε, the T -matrix in Eq. (4.55) can also be written
in the case Ef = E0 as
∞ n
X 1
hψf |T̂ |ψi i = hψf |Ĥ1 Ĥ1 |ψi i
n=0 E0 − Ĥ0 + iε
"∞ n #†
X 1
= hψf | Ĥ1 Ĥ1 |ψi i
n=0 E 0 − Ĥ 0 − iε
(−)
= hψf |Ĥ1 |ψi i ; Ef = E0 (4.75)
We now calculate the transition rate directly, in the presence of the adia-
batic switching. The derivation is from [Gell-Mann and Goldberger (1953)],
in their classic paper on scattering theory. The only subtlety in the cal-
culation is identifying those expressions that are well-defined in the limit
ε → 0, and knowing when to take that limit. This takes a little experience.
The Schrödinger state vector at the finite time t for a system that started
as |ψi i at t → −∞ is
i i i (+)
|Ψi (t)i = e− ~ Ĥt |Ψi (0)i = e− ~ Ĥt Ûε (0, −∞)|ψi i = e− ~ Ĥt |ψi i (4.77)
The states that one observes experimentally in scattering, decays, etc. are
the free-particle states
i
|Φf (t)i = e− ~ Ef t |ψf i (4.78)
This is the probability of having made a transition to the state |Φf (t)i at
the time t. The transition rate is the time derivative of this quantity
d d
ωf i = Pf i (t) = Mf⋆i (t) Mf i (t) + c.c. (4.80)
dt dt
We will show that this transition rate is independent of time for times such
that |t| ≪ 1/ε. In the end, we will again let Ω → ∞, and ε → 0, where Ω
is the quantization volume. Let us proceed to calculate the transition rate.
From Eqs. (4.77)–(4.79) one has
d i i i (+)
Mf i (t) = − e ~ Ef t hψf |Ĥ1 e− ~ Ĥt |ψi i (4.83)
dt ~
Now Eq. (4.58) states that in the above limit
(+)
(E0 − Ĥ)|ψi i=0 ; Ω→∞
ε→0 (4.84)
Substitution of these relations into Eq. (4.80) then expresses the tran-
sition rate as
2 (+) (+)
ωf i = Im hψf |Ĥ1 |ψi ihψf |ψi i⋆ (4.86)
~
This expression now has the following properties:
• It is independent of time;
• It is well-defined in the limit Ω → ∞, ε → 0.18
(+) 1
hψf |ψi i = hψf |ψi i + T̃f i (4.88)
E0 − Ef + iε
Substitution of this relation and the first of Eqs. (4.87) into Eq. (4.86) then
gives
2 2 1
ωf i = δf i Im T̃ii + Im |T̃f i |2 (4.89)
~ ~ E0 − Ef − iε
Finally, we make use of the relation
1 1
=P + iπδ(E0 − Ef ) (4.90)
E0 − Ef − iε E0 − Ef
Here P denotes the Cauchy principal value, defined by deleting an infinites-
imal symmetric region of integration through the singularity, and then let-
ting the size of that region go to zero. Equation (4.90) is a statement on
18 Here we will simply justify this observation a posteriori, through the many applica-
tions of the final expression. Note that by taking this limit too early in the derivation,
one can arrive at spurious results [for example, try substituting the second of Eqs. (4.85)
into Eq. (4.79)].
contour integration; it is derived in Prob. B.4. With the use of this relation,
Eqs. (4.89) and (4.76) become
2 2π
ωf i = δf i Im T̃ii + δ(E0 − Ef )|T̃f i |2 ; transition rate
~ ~
Sf i = δf i − 2πiδ(E0 − Ef )T̃f i ; S-matrix (4.91)
These expressions are exact. They are the results quoted in Eqs. (4.46) and
used extensively in Vol. I.
4.9 Unitarity
d X X d X
Pf i (t) = Pf i (t) = ωf i = 0 (4.93)
dt dt
f f f
Here the transition rate has been identified from Eq. (4.80). A substitution
of the expression for the transition rate in Eq. (4.91) into this relation then
gives
2 X 2π
− Im T̃ii = δ(Ef − E0 )|T̃f i |2 ; unitarity (4.94)
~ ~
f
This relation for the imaginary part of the elastic T -matrix, reflecting con-
servation of probability and depletion of the initial state, is known as uni-
tarity.
19 This sum now includes the state f = i; the reader should note that there is no sum
over the repeated index i implied in Eqs. (4.91) and (4.94).
p̂ |ti = ~t |ti
1
hx|ti = φt (x) = √ eit·x ; p.b.c. (4.97)
Ω
The eigenstates of momentum satisfy the completeness relation
X
|tiht| = 1̂ (4.98)
t
Insert this expression in Eq. (4.95), and use Eqs. (4.96) and (4.97)
2m X 1
G0 (x − y) = 2 hx|ti 2 ht|yi
~ t t − k 2 − iε
2m 1 X it·(x−y) 1
= 2 e (4.99)
~ Ω t t − k 2 − iε
2
In this limit
2m 1 1
Z
G0 (x − y) = 2 3
d3 t eit·(x−y) 2 (4.100)
~ (2π) t − k 2 − iε
t - plane
R C
k+i
-k-i
Fig. 4.1 Contour for the evaluation of the Green’s function G0 (x − y) in the complex
t-plane, together with the singularity structure arrived at with adiabatic damping. Here
R → ∞.
4.10.3 T-matrix
The T -matrix can also be expressed in the coordinate representation as
Z
(+)
T̃f i = d3 y hkf |yi V (y) hy|ψi i (4.110)
4.10.5 Unitarity
The scattering amplitude f (k, θ) and the T -matrix are related through
Eqs. (4.117) and (4.111), and thus
2 4π ~
− Im T̃ii = Im f (k, 0) (4.119)
~ Ω m
The unitarity relation in Eq. (4.94) states that
2 X 2π
− Im T̃ii = δ(Ef − E0 )|T̃f i |2 (4.120)
~ ~
f
Within a factor of the incident flux, the r.h.s. of this relation is just the
total cross section σtot . Thus Eq. (4.120) can be rewritten as
2 1 ~k
− Im T̃ii = Iinc σtot = σtot (4.121)
~ Ωm
A comparison of Eqs. (4.119) and (4.121) then leads to the optical theorem
relating the imaginary part of the forward elastic scattering amplitude and
the total cross section22
k
Im f (k, 0) = σtot ; optical theorem (4.122)
4π
The analysis of potential scattering in this section provides the under-
lying basis for the study of scattering in quantum mechanics, as presented,
for example, in [Schiff (1968)].23
22 So far, there is only elastic scattering in this potential model, but the optical theorem
is more general and holds in the presence of additional inelastic processes.
23 Problems 1.1–1.5 in [Walecka (2004)] take the reader through the essentials of the