You are on page 1of 46

1

1
Quantum Mechanics

Joachim Burgdörfer and Stefan Rotter

1.1 Introduction 3
1.2 Particle-Wave Duality and the Uncertainty Principle 4
1.3 Schrödinger Equation 6
1.4 Boundary Conditions and Quantization 8
1.5 Angular Momentum in Quantum Mechanics 9
1.6 Formalism of Quantum Mechanics 12
1.7 Solution of the Schrödinger Equation 16
1.7.1 Methods for Solving the Time-Dependent Schrödinger Equation 16
1.7.1.1 Time-Independent Hamiltonian 16
1.7.1.2 Time-Dependent Hamiltonian 17
1.7.2 Methods for Solving the Time-Independent Schrödinger Equation 19
1.7.2.1 Separation of Variables 19
1.7.2.2 Variational Methods 23
1.7.3 Perturbation Theory 24
1.7.3.1 Stationary Perturbation Theory 25
1.7.3.2 Time-Dependent Perturbation Theory 25
1.8 Quantum Scattering Theory 27
1.8.1 Born Approximation 28
1.8.2 Partial-Wave Method 29
1.8.3 Resonances 30
1.9 Semiclassical Mechanics 31
1.9.1 The WKB Approximation 31
1.9.2 The EBK Quantization 33
1.9.3 Gutzwiller Trace Formula 34
1.10 Conceptual Aspects of Quantum Mechanics 35
1.10.1 Quantum Mechanics and Physical Reality 36
1.10.2 Quantum Information 38
1.10.3 Decoherence and Measurement Process 39
1.11 Relativistic Wave Equations 41
1.11.1 The Klein–Gordon Equation 41
1.11.2 The Dirac Equation 42

Encyclopedia of Applied High Energy and Particle Physics. Edited by Reinhard Stock
Copyright  2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 978-3-527-40691-3
2 1 Quantum Mechanics

Glossary 43
References 44
Further Reading 45
3

1.1 characteristic speed v of the constituents


Introduction of the system under study. The domain
of nonrelativistic QM is characterized by
Quantum mechanics (QM), also called speeds that are small compared with the
wave mechanics, is the modern theory of speed of light:
matter, of atoms, molecules, solids, and of
the interaction of electromagnetic fields vc (1)
with matter. It supersedes the classical
mechanics embodied in Newton’s laws where c = 3 × 108 m s−1 is the speed of
and contains them as a limiting case (the light. For v approaching c, relativistic exten-
so-called classical limit of QM). Its develop- sions are necessary. They are referred to
ment was stimulated in the early part of the as relativistic wave equations. Because of the
twentieth century by the failure of classical possibility of production and destruction of
mechanics and classical electrodynamics to particles by conversion of relativistic energy
explain important discoveries, such as the E into mass M (or vice versa) according to
photoelectric effect, the spectral density of the relation (see chapter 2)
blackbody radiation, and the electromag-
netic absorption and emission spectra of E = Mc2 (2)
atoms. QM has proven to be an extremely
successful theory, tested and confirmed a satisfactory formulation of relativistic
to an outstanding degree of accuracy for quantum mechanics can only be devel-
physical systems ranging in size from sub- oped within the framework of quantum
atomic particles to macroscopic samples field theory. This chapter focuses primar-
of matter, such as superconductors and ily on nonrelativistic quantum mechanics,
superfluids. Despite its undisputed success its formalism and techniques as well as
in the description of physical phenomena, applications to atomic, molecular, optical,
QM has continued to raise many concep- and condensed-matter physics. A brief dis-
tual and philosophical questions, in part cussion of extensions to relativistic wave
due to its counterintuitive character, defy- equations is given at the end of the chapter.
ing common-sense interpretation. The new ‘‘quantum world’’ will be
One distinguishes between nonrelativis- contrasted with the old world of classical
tic quantum mechanics and relativistic mechanics. The bridge between the latter
quantum mechanics depending on the and the former is provided by the

Encyclopedia of Applied High Energy and Particle Physics. Edited by Reinhard Stock
Copyright  2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 978-3-527-40691-3
4 1 Quantum Mechanics

semiclassical mechanics, whose rigorous unifies the description of distinct entities


formulation and application have only of classical physics: mechanics and elec-
recently been developed. tromagnetism. Particles, the fundamental
constituents of matter, possess, in addition
to mechanical properties such as momen-
1.2 tum and angular momentum, wave charac-
Particle-Wave Duality and the Uncertainty ter. Conversely, electromagnetic radiation,
Principle successfully described by Maxwell’s the-
ory of electromagnetic waves, takes on,
The starting point of modern quantum under appropriate conditions, the prop-
mechanics is the hypothesis of the particle- erty of quasiparticles, the photons, carrying
wave duality put forward by de Broglie mechanical energy and momentum. The
(1924): the motion of each particle of mass particle and wave character are comple-
M is associated with a wave (‘‘matter wave’’) mentary to each other. An object behaves
of wavelength either like a particle or like a wave.
The hypothesis of particle-wave duality
h has been experimentally verified in remark-
λ= (3a)
p able detail. Interference of matter waves
due to the indistinguishability of paths
where h = 6.63 × 10−34 J s is Planck’s con-
was demonstrated first for electrons (Davis-
stant, discovered by M. Planck in 1900 in
son and Germer, 1927), later for neutrons,
his investigation of blackbody radiation,
and, most recently, for neutral atoms. The
and p = Mv is the classical momentum of
particle-wave duality of electromagnetic
the particle. Similarly, the frequency ν of
radiation and its complementarity have
the wave is associated with the mechan-
been explored through modern versions
ical energy of the particle through the
of Young’s double-slit experiment using
expression
state-of-the-art laser technology and fast
E switches. Hellmuth and colleagues (1987)
ν= (3b) succeeded, for example, in switching from
h
particle to wave character and vice versa
These relations, ascribing wave properties even after the photon has passed through
to particles, are the logical complements the slit (in the experiment, a beam split-
to postulates put forward 20 years earlier ter) by activating a switch (a Pockels cell)
by Einstein (1905), assigning the charac- in one of the two optical paths of the
teristics of particles (‘‘photons,’’ particles interferometer. By use of the switch, the
of zero rest mass) to electromagnetic radi- interference patterns are made to appear
ation. In fact, Einstein employed, in his or disappear, as expected for a wavelike or
explanation of the ‘‘photoelectric effect,’’ particlelike behavior of the photon, respec-
the expression E = hν (Eq. 3b) to find the tively. This ‘‘delayed choice’’ experiment
energy E of each photon associated with not only demonstrates the particle-wave
electromagnetic radiation of frequency ν. duality for photons but also elucidates
Photons can eject electrons from the atoms important conceptual aspects of the nonlo-
of a material provided that the energy cality of quantum mechanics. The decision
exceeds the binding energy of the elec- as to whether the photon will behave
tron. This particle-wave duality of QM like a particle or a wave need not be
1.2 Particle-Wave Duality and the Uncertainty Principle 5

made at the double slit itself but can be propagation. The physical interpretation of
spatially separated and temporarily post- the amplitude A was provided by Born
poned. (1926) in terms of a probability density
The consequences of the wave-particle
duality for the description of physical phe- P(x) = |ψ(x)|2 = |A|2 (5)
nomena are far reaching and multifaceted.
One immediate consequence that repre-
sents the most radical departure from for finding the particle at the coordinate x.
classical physics is the uncertainty princi- The matter wave (Eq. 4) is further charac-
ple (Heisenberg, 1927). One deeply rooted terized by the wave number k = 2π/λ = p/
notion of classical physics is that all dynam- and the angular frequency ω = 2πv = E/.
ical observables can be measured, at least The ‘‘rationalized’’ Planck’s constant  =
in principle, with arbitrary accuracy. Uncer- h/2π = 1.05 × 10−34 J s is frequently used
tainties in the measurement are a matter in QM.
of experimental imperfection that, in prin- The uncertainty principle is inherent to
ciple, can be overcome. The wave nature all wave phenomena. Piano tuners have
of particles in QM imposes, however, fun- exploited it for centuries. They sound a
damental limitations on the simultaneous vibrating tuning fork of standard frequency
accuracy of measurements of dynamical in unison with a piano note of the same
variables that cannot be overcome, no mat- nominal frequency and listen to a beat tone
ter how much the measurement process between the struck tune and the tuning
can be improved. This uncertainty princi- fork. For a fork frequency of ν = 440 Hz
ple is an immediate consequence of the and a string frequency of ν  = 441 Hz,
wave behavior of matter. Consider a wave one beat tone will be heard per second
of the form (δν = ν  − ν = 1 Hz). The goal of the tuner
is to reduce the number of beats as much
ψ(x) = Aei(kx−ωt) (4) as possible. To achieve an accuracy of δν =
0.010.01 Hz, the tuner has to wait for at
traveling along the x coordinate least about δt = 100 s (in fact, only a fraction
(Figure 1.1). Equation (4) is called a of this time since the frequency mismatch
plane wave because the wave fronts form can be detected prior to completion of
planes perpendicular to the direction of one beat period) to be sure no beat had

Traveling matter wave


Reψ(x, t)

0
0

2
t x Fig. 1.1 Matter wave traveling
3 along the x direction; the figure
10
shows the real part of ψ(x),
4 0
Eq. (4) (after Brandt and Dah-
−10 men, 1989).
6 1 Quantum Mechanics

occurred. Piano tuning therefore relies on implications become obvious only in the
the frequency–time uncertainty microscopic world of atoms, molecules,
electrons, nuclei, and elementary parti-
ν t  1 (6) cles. More generally, whenever the size
of an object or mean distance between con-
Measurement of the frequency with infi- stituents becomes comparable to the de
nite accuracy (ν → 0) requires an infinite Broglie wavelength λ, the wave nature of
time period of measurement (t → ∞), or matter comes into play. The size of λ also
equivalently, it is impossible within any provides the key for quantum effects in
finite period of time to determine the fre- macroscopic systems such as superfluids,
quency of the string exactly since time superconductors, and solids.
and frequency are complementary variables.
Similarly, the variables characterizing the
matter wave (Eq. 4) are complementary: 1.3
Schrödinger Equation
k x ≥ 1
ω t ≥ 1 (7) The Schrödinger equation for matter waves
ψ describes the dynamics of quantum
These relations, well known from classical particles. In the realm of quantum physics,
theory of waves and vibrations, have it plays a role of similar importance to that
unexpected and profound consequences which Newton’s equation of motion does
when combined with the de Broglie for classical particles. As with Newton’s
hypothesis (Eq. 3) for quantum objects: laws, the Schrödinger equation cannot be
rigorously derived from some underlying,
more fundamental principles. Its form can
p x ≥  be made plausible, however, by combining
E t ≥  (8) the Hamiltonian function of classical
mechanics,
The momentum and the position of
a particle, its energy and the time, H =T +V =E (9)
and more generally, any other pair of
complementary variables, such as angular which equals the mechanical energy, with
momentum and angle, are no longer the de Broglie hypothesis of matter waves
simultaneously measurable with infinite (Eq. 4). The potential energy is denoted by
precision. The challenge to the classical V and the kinetic energy by T = p2 /2M.
laws of physics becomes obvious when Formally multiplying the wave function
one contemplates the fact that Newton’s ψ(x,t) by Eq. (9) yields
equation of motion uses the simultaneous
knowledge of the position and of the p2
change of momentum. Hψ(x, t) = ψ(x, t) + Vψ(x, t) (10)
2M
Since  is extremely small, the uncer-
tainty principle went unnoticed during In order to connect the de Broglie relations
the era of classical physics for hun- for energy and momentum appearing in
dreds of years and is of little conse- the arguments of the plane wave (Eq. 4)
quence for everyday life. Its far-reaching with the energy and momentum in the
1.3 Schrödinger Equation 7

Hamiltonian function, H and p in Eq. (10) classical Hamiltonian function (Eq. 9),
must be taken as differential operators, through the replacement of the classi-
cal observable by the operator equivalents
∂ (Eq. 11).
H ←→ i
∂t The replacement of dynamical variables
 ∂
p ←→ (11) by differential operators leads to the
i ∂x noncommutativity of pairs of conjugate
This substitution into Eq. (10) leads to the variables in QM, for example,
time-dependent Schrödinger equation
 ∂  ∂
[x, px ] ≡ x − x = i (14)
∂ i ∂x i ∂x
i ψ(x, t) = Hψ(x, t)
∂t
 
2 ∂ 2 with similar relations for other comple-
= − + V(x)
2M ∂x2 mentary pairs, such as Cartesian coordi-
× ψ(x, t) (12) nates of position and momentum (y, py ),
(z, pz ), or energy and time. It is the noncom-
which describes the time evolution of the mutativity that provides the basis for a for-
matter wave. For physical systems that mal proof of the uncertainty principle, Eq.
are not explicitly time dependent, i.e., that (8). The Schrödinger equation (Eqs. 12 and
have time-independent potentials V(x), the 13) is a linear, homogeneous, and (in gen-
energy is conserved and i ∂/∂t can be eral) partial differential equation. One con-
replaced by E, giving the time-independent sequence of this fact is that matter waves
Schrödinger equation ψ satisfy the superposition principle. If ψ 1
 
and ψ 2 are two acceptable solutions of the
2 ∂ 2 Schrödinger equation, called states, a linear
Eψ(x) = − + V(x) ψ(x) (13)
2M ∂x2 combination

If no interaction potential is present


ψ = c1 ψ1 + c2 ψ2 (15)
(V(x) = 0), Eq. (12) is referred to as the
free-particle Schrödinger equation. The plane
waves of Eq. (4) are solutions of the is an acceptable solution, as well. The phys-
free-particle Schrödinger equation with ical implication of Eq. (15) is that the quan-
eigenenergies E = (k)2 /2M. tum particle can be in a ‘‘coherent superpo-
The correspondence principle, first for- sition of states’’ or ‘‘coherent state,’’ a fact
mulated by Bohr, provides a guide to that leads to many strange features of QM,
quantizing a mechanical system. It states such as destructive and constructive inter-
that in the limit of large quantum num- ference, the collapse of the wave function,
bers or small de Broglie wavelength, the and the revival of wave packets.
quantum-mechanical result should be the A solution of a linear, homogeneous
same, or nearly the same, as the classi- differential equation is determined only
cal result. One consequence of this pos- up to an overall amplitude factor, which
tulate is that the quantum-mechanical is arbitrary. The probability interpreta-
Hamiltonian operator in the Schrödinger tion of the matter wave (Eq. 5), however,
equation (Eq. 12) should be derived from removes this arbitrariness. The probabil-
its corresponding classical analog, the ity for finding the particle somewhere
8 1 Quantum Mechanics

is unity, time-independent Schrödinger equation:


 
2 ∂
dxP(x) = dx|ψ(x)|2 = 1 (16) − ψ(φ) = Eψ(φ) (18)
2I ∂φ 2

Equation (16) and its generalization to where φ is the angle of rotation about the z
three dimensions are referred to as the axis. Among all solutions of the form
normalization condition on the wave func-
tion and eliminate the arbitrariness of the ψ(φ) = Aei(mφ+φ0 ) (19)
modulus of the wave function. A wave
function that can be normalized according only those satisfying the periodic boundary
to Eq. (16) is called square integrable. The conditions
only arbitrariness remaining is the overall
‘‘phase factor’’ of the wave function, eiφ ,
ψ(φ + 2π ) = ψ(φ) (20)
i.e., a complex number of modulus 1.

are admissible in order to yield a uniquely


1.4 defined single-valued function ψ(φ). (The
Boundary Conditions and Quantization implicit assumption is that the rotator
is spinless, as discussed below.) The
In order to constitute a physically accept- magnetic quantum number m therefore
able solution of the Schrödinger equation, takes only integer values m = −∞, . . . , −1,
i.e., to represent a state, ψ must satisfy 0, 1, . . . ∞. Accordingly, the z component
appropriate boundary conditions. While of the angular momentum
their detailed forms depend on the coor-
dinates and symmetry of the problem Lz ψ(φ) = mψ(φ) (21)
at hand, their choice is always dictated
by the requirement that the wave func- takes on only discrete values of multiples
tion be normalizable and unique and that of the rationalized Planck’s constant .
physical observables take on only real The real number on the right-hand side
values (more precisely, real expectation that results from the operation of a
values) despite the fact that the wave func- differential operator on the wave function
tion itself can be complex. Take, as an is called the eigenvalue of the operator.
example, the Schrödinger equation for a The wave functions satisfying equations
rotator with a moment of inertia I, con- that are of the form of Eqs (18) or (21)
strained to rotate about the z axis. The are called eigenfunctions. The eigenvalues
classical Hamiltonian function H = L2z /2I, and eigenfunctions contain all relevant
with Lz the z component of the angular information about the physical system; in
momentum, particular, the eigenvalues represent the
measurable quantities of the systems.
 ∂ The amplitude of the eigenfunction
Lz = xpy − ypx = (17)
i ∂φ follows from the normalization condition
 2π
gives rise, according to the corre- dφ|ψ(φ)|2 = 1 (22)
spondence principle, to the following 0
1.5 Angular Momentum in Quantum Mechanics 9

as A = (2π)−1/2 . The eigenvalues of E in along the z axis becomes


Eq. (18) are E m = 2 m2 /2I. Just like for
a vibrating string, it is the imposition −γ BLz ψ(φ) = Em ψ(φ) (24)
of boundary conditions (in this case,
periodic boundary conditions) that leads to
‘‘quantization,’’ that is, the selection of an The eigenenergies
infinite but discrete set of modes of matter
waves, the eigenstates of matter, each of Em = −γ Bm (25)
which is characterized by certain quantum
numbers.
Different eigenstates may have the same are quantized and depend linearly on the
energy eigenvalues. One speaks then of magnetic quantum number. Depending
degeneracy. In the present case, eigenstates on the sign of γ , states with positive m
with positive and negative quantum num- for γ > 0, or negative m for γ < 0, corre-
bers ±m of Lz are degenerate since the spond to states of lower energy and are
Hamiltonian depends only on the square preferentially occupied when the atomic
of Lz . Degeneracies reflect the underlying magnetic moment is in thermal equilib-
symmetries of the Hamiltonian physical rium. This leads to an alignment of the
system. Reflection of coordinates at the magnetic moments of the atoms along one
x –z plane, (y → −y, py → −py and there- particular direction and to paramagnetism
fore Lz → −Lz ) is a symmetry operation, of matter when its constituents carry a net
i.e., it leaves the Hamiltonian invariant. In magnetic moment.
more general cases, the underlying symme- The force an atom experiences as
try may be dynamical rather than geometric it passes through an inhomogeneous
in origin. magnetic field depends on the gradient
of the magnetic field, ∇B. Because of
the directional quantization of the angular
1.5 momentum, a beam of particles passing
Angular Momentum in Quantum Mechanics through an inhomogeneous magnetic
field will be deflected into a set of
The quantization of the projection of the discrete directions determined by the
angular momentum, Lz , is called directional quantum number m. A deflection pattern
quantization. The Hamiltonian function consisting of a few spots rather than
for the interaction of an atom having a a continuous distribution of a beam
magnetic moment µ = γ L with an external of silver atoms was observed by Stern
magnetic field B reads and Gerlach (1921, 1922) and provided
the first direct evidence of directional
H = −µ · B = −γ L · B (23) quantization.
In addition to the projection of the angu-
The proportionality constant γ between lar momentum along one particular axis,
the magnetic moment and the angular for example, the z axis, the total angu-
momentum is called the gyromagnetic ratio. lar momentum L plays an important role
According to the correspondence principle, in QM. Here, one encounters a concep-
the Schrödinger equation for the magnetic tual difficulty, the noncommutativity of
moment in a field with field lines oriented different angular momentum components.
10 1 Quantum Mechanics

Application of Eq. (14) leads to − l, . . . , −1, 0, 1, . . . l. The total number of


    quantized projections of Lz is 2l + 1. The
Lx , Ly = iLz , Ly , Lz = iLx eigenvalue of the L2 operator is, however,
 
Lz , Lx = iLy (26) 2 l(l + 1) and not 2 l2 . The latter can be
understood in terms of the uncertainty
As noncommutativity is the formal expres- principle and can be illustrated with the
sion of the uncertainty principle, the help of a vector model (Figure 1.2). Because
implication of Eq. (26) is that different of the uncertainty principle, the projections
components of the angular momentum Ly and Lx do not take a well-defined value
cannot be simultaneously determined with and cannot be made to be exactly zero,
arbitrary accuracy. The maximum num- when Lz possesses a well-defined quantum
ber of independent quantities formed by number. Therefore, the z projection m
components of the angular momentum will always  be smaller than the magnitude
that have simultaneously ‘‘sharply’’ defined of L =  l(l + 1) in order to accommodate
eigenvalues is two. Specifically, the nonzero fluctuations in the remaining
components.
[L2 , Lz ] = 0 (27) The spherical harmonics are given in
terms of associated Legendre functions
with L2 = L2x + L2y + L2z being the square of Plm (cos θ ) by
the total angular momentum. Therefore,  1/2
one projection of the angular momentum (2l + 1) (l − m)!
Ylm (θ , φ) = (−1)m
vector (conventionally, one chooses the z 4π (l + m)!
projection) and the L2 operator can have ×Plm (cos θ )eimφ (29)
simultaneously well-defined eigenvalues,
unaffected by the uncertainty principle. The eigenfunctions of Lz (Eq. 19) can be
The differential operator of L2 in spherical recognized as one factor in the expression
coordinates gives rise to the following for Ylm . The probability densities |Ylm |2 of
eigenvalue equation: the first few eigenfunctions are represented
   by the polar plots in Figure 1.3.
1 ∂ ∂ In addition to the orbital angular
L Y(θ , φ) = −
2 2
sin θ
sin θ ∂θ ∂θ momentum operator L, which is the

1 ∂2 quantum-mechanical counterpart to the
+ Y(θ , φ)
sin2 θ ∂φ 2 classical angular momentum, another type
= 2 l(l + 1)Y(θ , φ) (28) of angular momentum, the spin angu-
lar momentum, plays an important role
The solutions of Eq. (28), Ylm (θ , ψ), are for which, however, no classical analog
called spherical harmonics. The superscript exists. The spin is referred to as an inter-
denotes the magnetic quantum number nal degree of freedom. The Stern–Gerlach
of the z component, as above, while experiment discussed above provided the
the quantum number l = 0, 1, . . . ∞ is major clue for the existence of the electron
called the angular momentum quantum spin (Goudsmit and Uhlenbeck, 1925). The
number. Since |Lz | ≤ | L|, we have the orbital angular momentum quantization
restriction |m | ≤ l, that is, for given angular allows only for an odd number of dis-
momentum quantum number l, the z crete values of Lz (2l + 1 = 1, 3, 5, . . .)
component varies between −l and l : m = and therefore an odd number of spots in
1.5 Angular Momentum in Quantum Mechanics 11

Lz Fig. 1.2 Vector model of angular momentum


√ L
for quantum number l = 2 and |L| = 6 .

2h

6h
h L

6h

0
6h

6h
−h

6h

−2h

the deflection pattern. Stern and Gerlach by relativistic quantum mechanics (see
observed, however, two discrete directions Section 1.11). The identification of the spin
of deflection (see schematic illustration in as an angular momentum rests on the
Figure 1.14). The spin hypothesis solved validity of the commutation rules (Eq. 26),
this mystery. The outermost electron of    
a silver atom has a total orbital angular Sx , Sy = iSz , Sy , Sz = iSx ,
 
momentum of zero but has a spin angular Sz , Sx = iSy (30)
momentum of s = 12 . The number of dif-
ferent projections of the spin is therefore identical to those of the orbital angular
2s + 1 = 2 in accordance with the num- momentum. Any set of the components
ber of components into which the beam of a vector operator satisfying Eq. (26)
was magnetically split. Within nonrela- has a spectrum of eigenvalues (quantum
tivistic quantum mechanics, the spin must numbers) of an angular momentum with
be introduced as an additional degree of quantum numbers j = 0, 12 , 1, . . . and m =
freedom on empirical grounds. A concep- −j, −j + 1 . . . , + j. The quantum number
tually satisfactory description is provided j is an integer in systems with an even
12 1 Quantum Mechanics

Fig. 1.3 Polar plots of the probability


θ
distributions |Ylm (θ, φ)|2 of angular momentum
eigenfunctions.
=0

m=0

=1

m=0 m = ±1

=2

m=0 m = ±1 m = ±2

number of spin 12 particles (including an infinite-dimensional vector space, the


zero) and half-integer in systems with Hilbert space,
an odd number of spin 12 particles. The
Hamiltonian for the interaction of an ψ −→ |ψ (32)
electron spin with an external magnetic
field is given in direct analogy to Eq. (23) with the vectors represented by a ‘‘ket.’’
by The ket notation was first introduced by
Dirac (1930). The coherent superposition
illustrated by Eq. (15) can be recognized as
H = −γ gS · B (31)
vector addition in Hilbert space. Physical
observables such as the Hamiltonian or the
The only difference is the ‘‘anomalous’’ angular momentum can be represented by
gyromagnetic ratio represented by the linear and Hermitian operators. Linearity
factor g. An approximate value of g can be of an operator B means
deduced from the relativistic wave equation
(Dirac’s equation) for the electron (see B(c1 |ψ1 + c2 |ψ2 ) = c1 B|ψ1 + c2 B|ψ2
Section 1.11). (33)
for all complex numbers c1 , c2 and all
vectors |ψ 1 , |ψ 2 . Hermitian operators are
1.6 defined as those for which
Formalism of Quantum Mechanics
B† = B (34)
The quantum-mechanical description out-
lined in the previous section with the help that is, the operator is equal to its Hermi-
of specific examples can be expanded to tian adjoint. The Hermiticity assures that
a general formalism of quantum mechan- the eigenvalue equation
ics. The starting point is the identifica-
tion of wave functions with vectors in B|ψ = λ|ψ (35)
1.6 Formalism of Quantum Mechanics 13

possesses real eigenvalues λ, which can be containing only its eigenstates {ψ i } and its
identified with measurable quantities. The eigenvalues λi . Two operators possess the
time-independent Schrödinger equation, same eigenbasis when they commute with
Eq. (13), is of the form in Eq. (35) with each other. Physical observables whose
B being the Hamiltonian operator H and operators commute with each other are
λ being the eigenenergy E. The result of called compatible observables. A state of a
a measurement of a physical observable physical system is therefore specified to
yields an eigenvalue λ of the corresponding the maximum extent possible, consistent
operator, B. with the uncertainty principle, by one
The noncommutativity of operators (Eq. common eigenstate of a maximum set
14) corresponds to the well-known non- of compatible observables. For example,
commutativity of matrices in linear alge- the set {L2 ,Lz } is the maximum set of
bra. The overlap of wave functions can be compatible observables for the angular
identified with the scalar product in Hilbert momentum with the common eigenstate
space: |lm .
The probability amplitude for finding the

system in a particular state |ψ 1 is given by
φ|ψ ≡ d3 r φ ∗ (r)ψ(r) (36)
the projection of the state vector onto this
state, i.e., by the scalar product ψ 1 | ψ .
The adjoint of a vector in Hilbert space The corresponding probability is
ψ| is called a bra and the scalar
product a bracket (Dirac, 1930). The
P(ψ1 ) = | ψ1 |ψ |2 (39)
normalization requirement imposed on
the wave function can be expressed in
terms of the scalar product as ψ | ψ = 1. Formally, Eq. (39) can be viewed as the
The requirement of normalizability (Eq. expectation value of the projection operator
22) implies that the Hilbert space is P =|ψ 1 ψ 1 | in the state |ψ . The process
spanned by square-integrable functions. of a measurement is therefore associated
Since eigenstates of Hermitian operators with a projection in Hilbert space. This
belonging to different eigenvalues are projection is sometimes referred to as
orthogonal to each other, each physical the collapse of the wave function. The
dynamical variable possesses a complete process of measurement thus influences
orthonormal set of eigenstates: the quantum system to be measured.
The state of a physical system is often
1 i=j not completely specified by a ‘‘pure’’
ψi |ψj = δij = (37) state described by a single Dirac ket
0 i = j
|ψ or a coherent superposition of kets
The operator B is a diagonal matrix in ∼c1 | ψ 1 +c2 | ψ 2 . The limited informa-
a basis consisting of its eigenstates, the tion available may only allow us to specify a
‘‘eigenbasis’’ |ψ i with i = 1, . . . , ∞. In its statistical mixture of states. Such a mixture
eigenbasis, the operator B possesses the is described by the statistical operator or
following ‘‘spectral representation’’: density operator

B= λi |ψi ψi | (38) ρ= Pi |ψi ψi | (40)


i i
14 1 Quantum Mechanics

with Pi denoting the probability for the with the strength of the singularity such
system to be found in a particular state that the area underneath the peak is
|ψ i . The result of the measurement of an 
observable B is then given by the statistical dxδ(x) = 1 (45)
expectation value


While the δ function is not an ordinary
B = Tr(ρB) = ψi |B|ψi Pi (41) (Riemann integrable) function, it can be
i approximated as a limit of Riemann
integrable functions, for example,
i.e., by the diagonal elements of the opera-
1 ε
tor B, each weighted with the probability to lim = δ(x) (46)
find the system in the corresponding state.
ε−→0 π x2 + ε2
The preferred set of states, |ψ i , for which Equation (46) satisfies Eq. (45) provided
the density matrix is ‘‘diagonal,’’ that is, of one evaluates the integral first and takes
the form in Eq. (40), is sometimes referred the limit ε → 0 afterward. As the order of
to as the set of pointer states. the operations of integration and taking the
In light of the expression for the limit is not interchangeable, Eq. (46) is said
probability as a projection, Eq. (39), we to converge ‘‘weakly’’ toward a δ function
can rewrite Eq. (5) as in the limit ε → 0. δ functions play an
important role in the quantum theory of
P(r) = |ψ(r)|2 = | r|ψ |2 (42) scattering.
The time evolution of the Hilbert-space
vectors is given by Schrödinger’s equation
where |r is formally treated as a vector
written in Dirac’s ket notation:
(‘‘ket’’) in Hilbert space. This definition
leads to the difficulty that |r cannot be ∂
i |ψ = H|ψ (47)
defined as a square-integrable function. ∂t
Nevertheless, it can be consistently incor-
For systems with a classical analog, the
porated into the formalism by employing
Hamiltonian operator is given, according
the concept of generalized functions (‘‘dis-
to the correspondence principle, by the
tributions’’; see Lighthill, 1958; Schwartz,
classical Hamiltonian function H(r, p)
1965). One defines a scalar product
upon treating the canonical variables (r, p)
through
as noncommuting operators (see Eq. 14).
There are, however, additional dynamical
r|r = δ(r − r ) variables in quantum mechanics, such as
= δ(x − x )δ(y − y )δ(z − z ) (43) spin, for which no classical counterparts
exist.
A vector in Hilbert space can be expanded
where the right-hand side is called Dirac’s
in terms of any basis (a ‘‘representation’’),
δ function. The δ function is zero almost
everywhere:

|ψ = ci |φi = φi |ψ |φi (48)


i i

0, x = 0
δ(x) = (44) where ci = φ i | ψ are the expansion coef-
∞, x=0 ficients of the ket |ψ in the basis
1.6 Formalism of Quantum Mechanics 15

{|φ i }. Switching from one representation motion depends on the initial conditions
to another |ψ → | ψ  = U | ψ amounts x(t = t0 ), p(t = t0 ), which are not subject
to a unitary transformation, U, since to any symmetry constraint. They may
the norm of a state vector must be ‘‘break’’ the symmetry and usually do so. In
conserved, quantum mechanics, however, the parity
of a state vector becomes a dynamical
ψ  |ψ  = ψ|U † U|ψ observable.
= ψ|ψ (49) We can define a parity operator p̂ through

p̂H(x, p)p̂† = H(−x, −p) (52)


or U † U = 1. Transformations among rep-
resentations play an important role in QM.
Since p̂2 is the unit operator (two successive
While the maximum number of compat-
reflections restore the original function),
ible observables is determined by com-
we have p̂ = p̂† = p̂−1 . Eigenfunctions of
mutation rules, the choice of observables
p̂, ψ(x) = x|ψ , can now be characterized
that are taken to be diagonal or to have
by their parity:
‘‘good’’ quantum numbers depends on
the specific problem under consideration.
x|p̂|ψ = −x|ψ = ± x|ψ (53)
Unitary transformations provide the tool
to switch from a basis within which one The ± sign in Eq. (53) is referred to
set of observables is diagonal to another.
as indicating positive or negative parity. A
The transformation law for state vectors
wave function cannot always be assigned a
|ψ  = U | ψ implies the transformation
well-defined parity quantum number ±1.
law for operators
However, when p̂ commutes with H,

A = UAU † (50)
p̂H(x, p)p̂ = H(x, p) (54)

The latter follows from the invariance there exists a common eigenbasis in
of the expectation value ψ  | A | ψ  = which both the Hamiltonian operator
ψ | U † A U | ψ under unitary transfor- and the parity operator are diagonal and,
mations. consequently, ψ is an eigenstate of p̂ with a
Classical Hamiltonian functions as well-defined parity. A similar argument can
well as quantum-mechanical Hamiltonian be developed for symmetry with respect
operators possess discrete symmetries. For to the time-reversal operator T: t → −t,
example, the Hamiltonian function of the p → −p, r → r.
harmonic oscillator, An additional discrete symmetry, which
is of fundamental importance for the
p2
H(x, p) = + 12 Mω2 x2 (51) quantum mechanics of identical particles,
2M
is the permutation symmetry. Consider
is invariant under reflection of all spatial a wave function describing the state
coordinates (x → −x, p → −p) since H of two identical particles ψ(r1 , r2 ). The
depends only on the square of each permutation operator p̂12 exchanges the
canonical variable. In classical mechanics, two particles in the wave function:
this symmetry is of little consequence
since the solution of Newton’s equation of p̂12 ψ(r1 , r2 ) = ψ(r2 , r1 ) (55)
16 1 Quantum Mechanics

The spin-statistics postulate of quantum that


mechanics makes a general statement  
concerning the permutation symmetry. A −iH(t − t0 )
U(t, t0 ) = exp (57)
many-body system consisting of identical 
particles of integer spin (‘‘bosons’’) pos-
sesses wave functions that are even under where the exponential function of the oper-
all interchanges of two particles, whereas ator H is defined through its power-series
the wave function is odd for half-integer expansion
spin particles (‘‘fermions’’). The eigenvalue
of p̂ij , the operator of permutation for any t − t0
U(t, t0 ) = 1 − i H
two particles of an N-body system, is +1 for   2
bosons and −1 for fermions. Wave func- 1 t − t0
− H · H + · · · (58)
tions are totally symmetric under particle 2 
exchange for bosons and totally antisym-
metric for fermions. An immediate con- If one is able to solve the time-independent
sequence is that two fermions cannot be Schrödinger equation
in identically the same state for which
all quantum numbers agree (the so-called H|ψn = En |ψn (59)
Pauli exclusion principle) since the anti-
symmetric wave function would vanish
identically. where n labels the eigenstate, one can cal-
culate the time evolution of the eigenstate:

1.7 U(t, t0 )|ψn = e−iH(t−t0 )/ |ψn


Solution of the Schrödinger Equation = e−iEn (t−t0 )/ |ψn (60)

1.7.1
Methods for Solving the Time-Dependent using the definition in Eq. (58). The phase
Schrödinger Equation factor in Eq. (60) is called the dynamical
phase. The explicit solution for an arbitrary
1.7.1.1 Time-Independent Hamiltonian state can then be found by expanding
If the Hamilton operator is not explic- |ψ(t0 ) in terms of eigenstates of the
itly time dependent, i.e., the total Hamilton operator:
energy is conserved, the time-dependent

Schrödinger equation can be reduced to the |ψ(t0 ) = |ψn ψn |ψ(t0 )


time-independent Schrödinger equation. n
Equation (47) can be formally solved by |ψ(t) = U(t, t0 )|ψ(t0 )

= e−iEn (t−t0 )/ |ψn


n

|ψ(t) = U(t, t0 )|ψ(t0 ) (56) × ψn |ψ(t0 ) (61)

where U(t, t0 ) is called the time-evolution Hence, for time-independent systems, the
operator, a unitary transformation describ- solution of the energy eigenvalue problem
ing the translation of the system in time. (Eq. 59) is sufficient to determine the time
Substitution of Eq. (56) into Eq. (47) shows evolution for any initial state.
1.7 Solution of the Schrödinger Equation 17

1.7.1.2 Time-Dependent Hamiltonian The additional phase γ n (t) is called


For a time-dependent Hamiltonian H(t), Berry’s phase (Berry, 1984) or the geometric
the solution of the time-evolution operator phase, since it records the information on
U in terms of the exponential function the excursion of the Hamiltonian H(t ) in
Eq. (57) must be redefined in order to parameter space in the case of adiabatic
take into account that the Hamiltonian evolution. The role of γ n (t) is to relate
operator at different instances of time may the phases of the adiabatic wave function
not commute, [H(t), H(t )] = 0. We have |ψ n (t) at different points in time. Insertion
of Eq. (63) into Schrödinger’s equation

i t  yields the equation that defines the real
U(t, t0 ) = 1 − dt H(t )
 t0 phase function γ n (t),
 t
1 d
− 2 dt H(t ) γ̇n (t) − i φn,t φn,t = 0 (64)
 t0 dt
 t
× dt H(t ) · · · The analogy to the parallel transport of vec-
t0
   tors on a sphere (Figure 1.4) can give a
i t
= T exp − H(t )dt (62) glimpse of the concept underlying Berry’s
 t0 phase. Consider a quantum-mechanical
state to be represented by a vector at point
Equation (62) is called the time-ordered A. By infinitely slowly (adiabatically) vary-
exponential function and T is called ing an external parameter, for example, the
the time-ordering operator. In two lim- direction of the magnetic field, one can
iting cases, the evolution operator for ‘‘transport’’ state vectors successively from
time-dependent Hamiltonian systems can point A to B, then to C, and finally back
be easily calculated by reducing the prob- to A. After completion of the closed loop,
lem to the time-independent Hamiltonian. the vector does not coincide with its orig-
One is the adiabatic limit. If the temporal inal orientation but is rotated by an angle
changes in H(t) are very slow compared γ that is proportional to the solid angle
with the changes in the dynamical phases subtended by the loop. This angle gives
of Eq. (60) for all t, the U matrix is diagonal rise to a geometric phase, exp(iγ ), which
in the slowly changing eigenbasis of H(t) depends on the geometry of a loop, but
at each instant t, unlike the dynamical phase, Eq. (60), it is
   independent of the time it takes to com-
i t  plete the transport of the state along the
|ψn (t) = eiγn (t) exp − dt En (t ) |φn,t
 t0 closed loop. If one is now able to form in a
(63) physical system a coherent superposition of
The dynamical phase factor is determined two states, one accumulating the geomet-
by E n (t ), the energy eigenvalue of the ric phase along the adiabatic change of the
stationary Schrödinger equation for the external field while the other is subject to a
Hamiltonian operator H(t ) at time t . In Eq. constant field, an interference pattern will
(63), the state |ψ n (t) denotes the evolved arise. The Berry phase has been observed
state and the state |φ n,t represents the in interference experiments involving neu-
stationary eigenstate for the Hamiltonian tron spin rotation, in molecules, in light
H = H(t) at the time t, where the time plays transmission through twisted optical-fiber
the role of a parameter. cables, and in the quantum Hall effect.
18 1 Quantum Mechanics

Fig. 1.4 Parallel transport of a vector on a sphere


along a closed loop leads to a rotation of the vector.
The vector’s orientation relative to its local environment
C
remains unchanged during parallel transport. The angle
of rotation is related to Berry’s phase.

γ B
A

In the opposite limit, where a sudden The projection onto a final state |f
change of the Hamiltonian from H(t) = representing, for example, the atom in
H0 for all t < 0 to another Hamiltonian the ground state and the emission of a
H(t) = H0 for all t > 0 occurs, no dynamical photon,
or geometric phase can be accumulated
during the instantaneous change of the
Pf (t) = | f |ψ(t) |2
Hamiltonian at t = 0. The initial eigenstate
of H0 , ψ n , in which the system is prepared = |c1 |2 | f |ψ1 |2 + |c2 |2 | f |ψ2 |2
 
at t = 0, evolves for t > 0 as it
+ c1 c2∗ exp (E2 − E1 )t


 
i ε n t × f |ψ1 f |ψ2 ∗
|ψn (t) = exp − |ψ n ψ n |ψn
 + complex conjugate (67)
n
(65)
In the sudden limit, the matrix U is non-
will display oscillations (‘‘quantum beats’’)
diagonal. The eigenstate belonging to the
due to the time dependence of the relative
‘‘old’’ Hamiltonian forms a coherent super-
phases of the two states. These quantum
position of eigenstates |ψ n of the new
beats due to dynamical phases have been
Hamiltonian. The ‘‘sudden’’ formation of
observed, for example, in the Lyman-α (n =
coherent states causes ‘‘quantum beats,’’
2 → n = 1) radiation of hydrogen excited in
observable oscillations in amplitudes and
the n = 2 level subsequent to fast collisions
probabilities.
with a foil (Figure 1.5).
Consider, for example, an atom that is
The standard computational method of
excited at the time t = 0 to a state that is a
solving the time-dependent Schrödinger
coherent superposition of a state |ψ 1 , with
equation for time-dependent interactions
energy E 1 , and a state |ψ 2 , with energy E 2 ,
consists of expanding a trial wave function
that is, |ψ = c1 | ψ 1 + c2 | ψ 2 . At a later
|ψ(t) in terms of a conveniently chosen
point in time, t, the state acquires a phase
truncated time-independent basis of finite
factor as a function of time according to
size:
Eq. (65):

N
|ψ(t) = c1 e−iE1 t/ |ψ1 + c2 e−iE2 t/ |ψ2 |ψ(t) = an (t)|ψn (68)
(66) n=1
1.7 Solution of the Schrödinger Equation 19

Fig. 1.5 Quantum beats in the Lyman-α photon


intensity of hydrogen after collision with a carbon
foil: •, experimental data from Andrä (1974); – ,
theory from Burgdörfer (1981). The theory curve is
shifted relative to the data for clarity.
Intensity

0 10 20 30 40 50 mm

with unknown, time-dependent coeffi- 1.7.2.1 Separation of Variables


cients an (t). Insertion of Eq. (68) into Eq. The energy eigenvalue problem, Eq. (59),
(47) leads to a system of coupled first-order requires the solution of a second-order par-
differential equations, which can be written tial differential equation. It can be reduced
in matrix notation as to an ordinary second-order differential
equation if the method of separation of
iȧ = H(t) a(t) (69) variables can be applied, i.e., if the wave
function can be written in a suitable coor-
using the vector notation a = (a1 , . . . an ) dinate system (q1 , q2 , q3 ) in factorized form
with H the matrix of the Hamiltonian
in the basis {|ψ n }, Hn, n = ψ n | H | ψ n .
The solution of Eq. (69) is thereby reduced ψ(r) = f (q1 )g(q2 )h(q3 ) (70)
to a standard problem of numerical mathe-
matics, the integration of a set of N coupled with obvious generalizations for an arbi-
first-order differential equations with vari- trary number of degrees of freedom.
able (time-dependent) coefficients. Equation (70) applies only to a very spe-
cial but important class of problems for
1.7.2 which the Hamiltonian possesses as many
Methods for Solving the Time-Independent constants of motion or compatible observ-
Schrödinger Equation ables as there are degrees of freedom (in
our example, three). In this case, each fac-
The solution of the stationary Schrödinger tor function in Eq. (70) is associated with
equation, Eq. (59), proceeds mostly in one ‘‘good’’ quantum number belonging
the {|r } representation, i.e., using wave to one compatible observable. This class
functions ψ(r). Standard methods include of Hamiltonians permits an exact solution
the method of separation of variables and of the eigenvalue problem and is called
variational methods. separable (or integrable).
20 1 Quantum Mechanics

Consider the example of the isotropic all contributions in the orthogonality inte-
three-dimensional harmonic oscillator. gral (Eq. 36) be made to cancel. The wave
The Hamiltonian is functions for excited states of a harmonic
oscillator are given in terms of Hermite
py 2
p2x p2 polynomials Hn by
H= + + z
2M 2M 2M
 1/4
+ 12 Mω2 (x2 + y2 + z2 ) (71) Mω
fn (x) = (2n n!)−1/2

 
In this case, one suitable set of coordinates Mω
−(Mω/2π )x 2
for the separation are the Cartesian ×e Hn x (75)

coordinates (q1 = x, q2 = y, q3 = z). Each
of the three factor functions of Eq. (70)
Examples of the first few eigenstates are
satisfies a one-dimensional Schrödinger
shown in Figure 1.6 With increasing n,
equation of the form
the probability density |f n (x) | 2 begins to
  resemble the classical probability density
2 ∂ 2
− + 2 Mω x f (x) = Ex f (x)
1 2 2
for finding the oscillating particle near the
2M ∂x2
coordinate x, Pcl (x) (Figure 1.7). The latter
(72)
and E = Ex + Ey + Ez . The simplest solu- is proportional to the time the particle
tion of Eq. (72) satisfying the boundary spends near x or inversely proportional
conditions f (x → ±∞) → 0 for a normal- to its speed:
izable wave function is a Gaussian wave ω −1/2
function of the form Pc1 (x) = 2Ex /M − ω2 x2 (76)
π
 1/4
Mω 2
f (x) = e−(Mω/2)x (73)
π The increasing similarity between the clas-
sical and quantal probability densities as
It corresponds to the lowest possible energy n → ∞ is at the core of the correspon-
eigenstate, with Ex = 12 ω. Ground-state dence principle. This example highlights
wave functions are, in general, node the nonuniformity of the convergence to
free, i.e., without a zero at any finite x. the classical limit: the quantum probabil-
The Schrödinger equation possesses an ity density oscillates increasingly rapidly
infinite number of admissible solutions around the classical value with decreasing
with an increasing number n = 0, 1, . . . of de Broglie wavelengths λ (compared with
nodes of the wave function and increasing the size of the classically allowed region).
eigenenergies Furthermore, the wave function penetrates
  the classically forbidden region (‘‘tunnel-
Ex = ω n + 12 (74) ing’’). While the amplitude decreases expo-
nentially, ψ(x) possesses significant non-
The fact that wave functions belonging to vanishing values outside the domain of
higher states of excitation have an increas- classically allowed trajectories of the same
ing number of zeros can be understood as a energy over a distance of the order of the
consequence of the orthogonality require- de Broglie wavelength λ. Only upon aver-
ment, Eq. (37). Only through the additional aging over regions of the size of λ can
change of sign of the wave function can the classical–quantum correspondence be
1.7 Solution of the Schrödinger Equation 21

E Fig. 1.6 Wave functions f n (x) for the first five


ψn(X) eigenstates of the harmonic oscillator.
5

0
x
−4 −2 0 2 4

0.4 Fig. 1.7 Probability density


n = 39 |f n (x)|2 of a harmonic oscillator
in n = 39 compared with the
0.3 classical probability density
Classical probability density Pcl (x) = (ω/π )(2E/M − ω2 x 2 )−1/2
indicated by the dashed line.
0.2 Classical turning points are
denoted by a and b.
Quantum probability density

0.1

0.0
Turning point a Turning point b

−0.1
−10 0 10
x

recovered. As λ tends to zero in the classi- The energy depends only on the sum of the
cal limit, classical dynamics emerges. It three quantum numbers,
is the smallness of λ that renders the
macroscopic world as being classical for n = nx + ny + nz (79)
all practical purposes.
The complete wave function for the
three-dimensional isotropic oscillator (Eq. but not on the individual quantum num-
71) consists of a product of three wave bers nx , ny , and nz , separately. Different
functions each of which is of the form of combinations of quantum numbers yield
Eq. (75), the same energy; the spectrum is therefore
degenerate. The degeneracy factor g n , the
ψnx ny nz (x, y, z) = fnx (x) fny (y) fnz (z) (77) number of different states having the same
energy E n , is
with eigenenergies
  (n + 1)(n + 2)
Enx ny nz = ω nx + ny + nz + 32 (78) gn = (80)
2
22 1 Quantum Mechanics

Only the ground state (n = 0) is nonde- equation is separable in spherical coordi-


generate, a feature generic to quantum nates and the wave function can be writ-
systems and the origin of the third law ten as a product ψ(r, θ , φ) = Rnl (r)Yem (θ , φ)
(Nernst’s theorem) of statistical mechan- with one factor being the spherical har-
ics. The degeneracy signals that alternative monics Ylm and the Rnl (r) being the radial
complete sets of commuting observables wave function satisfying the second-order
(instead of the energies in each of the three differential equation
Cartesian coordinates E x , E y , and E z ) and   
corresponding alternate sets of coordinates −2 ∂2 2 ∂ 2 l(l + 1)
2
+ +
should exist in which the problem is sep- 2me ∂r r ∂r 2me r 2
2 
arable. One such set of coordinates is the e
− Rnl (r) = ERnl (r) (83)
spherical coordinates. 4π ε0 r
Consider the example of the hydrogen
atom. One of the most profound early The negative-energy eigenvalues of bound
successes of quantum mechanics was the states can be found as
interpretation of the line spectrum emitted me e 4 EH
from the simplest atom, the hydrogen En = − =− 2 (84)
22 n2 (4π ε0 )2 2n
atom, consisting of an electron and a
proton. The Schrödinger equation (in the where E H = me e4 /2 (4πε0 )2 is called a
center-of-mass frame) for the electron Hartree or an atomic unit of energy. The
interacting with the proton through a energy depends only on the ‘‘principal’’
Coulomb potential V(r) = −e2 /4πε0 r reads quantum number n, being independent of
as the angular momentum
 
−2 ∇ 2 e2 l = 0, 1, . . . , n − 1 (85)
− ψ(r) = Eψ(r) (81)
2me 4π ε0 r
and of the magnetic quantum number m.
where me = 9.1 × 10−31 kg is the reduced The m independence can be found for any
mass of the electron, r is the vector that isotropic potential V(|r|) that depends only
points from the nucleus to the electron, on the distance of the interacting particles.
e = 1.6 × 10−19 C is the electric charge of The degeneracy in l is a characteristic
the nucleus, and −e is the charge of the feature of the Coulomb potential (and of the
electron. If we express the Laplace operator isotropic oscillator), and it reflects the fact
∇ 2 (defined as ∇ 2 = ∂ 2 /∂x2 + ∂ 2 /∂y2 + that the Schrödinger equation is separable
∂ 2 /∂z2 ) in polar spherical coordinates (r, not only in polar spherical coordinates but
θ , φ), the Schrödinger equation becomes also in parabolic coordinates. The Coulomb
   problem possesses a dynamical symmetry
−2 ∂ 2 2 ∂ L2 that allows the choice of alternative
2
+ +
2me ∂r r ∂r 2me r 2 complete sets of compatible observables.
2  The separability in spherical coordinates
e
− ψ(r, θ , φ) = Eψ(r, θ , φ) (82)
4π ε0 r is related to the set {H, L2 , Lz } while
the separability in parabolic coordinates
Since the angular coordinates enter only is related to the set {H, Lz , Az } where
through the square of the angular momen- A is the Runge–Lenz (or perihelion)
tum operator (see Eq. 28), the Schrödinger vector, which, as is well known from
1.7 Solution of the Schrödinger Equation 23

Fig. 1.8 The Runge–Lenz vector A and the angular


L momentum L for the classical orbit in an attractive
−1/r potential.

A
Nucleus

v L ×A

Electron v

astronomy, is a constant of motion for a nonlinearly on parameters (c1 , . . . ,ck ). The


1/r potential (Figure 1.8). The emission trial function is optimized by minimizing
spectrum follows from Eq. (84) as the energy in the k-dimensional parameter
space according to
(En − En0 )
ν=
d
h  E = 0 (1≤i≤k) (87)
EH 1 1 dci
= − (86)
2h n20 n2
Provided the trial function is sufficiently
with n0 = 1 and n = 2, 3, . . . giving rise to flexible to approximate a complete set of
the Lyman series of spectral lines, n0 = 2 states in Hilbert space, ψ becomes exact in
and n = 3, 4, . . . giving rise to the Balmer the limit as k → ∞. The ground state of the
series, and so on. The radial probability helium atom consisting of an α particle as
densities r 2 Rnl (r)2 are shown in Figure 1.9 a nucleus (nuclear charge number Z = 2)
and two electrons provides one of the
1.7.2.2 Variational Methods most striking illustrations of the success of
For all but the few separable Hamilto- the variational method. The Hamiltonian
nians, the time-independent Schrödinger (neglecting the motion of the nucleus)
equation is only approximately and numer-
ically solvable. The accuracy of approxi- 2 Ze2
H=− (∇12 + ∇22 ) −
mation can, however, be high. It may 2me 4π ε0 r1
only be limited by the machine preci- Ze2 e2
− + (88)
sion of modern computers. The most 4π ε0 r2 4π ε0 |r1 − r2 |
successful and widely used method in
generating accurate numerical solutions is nonseparable, an analytic solution is
is the variational method. The solution unknown, and no perturbation solution
of the Schrödinger equation can be rec- is possible since all interactions are
ognized as a solution to the variational of comparable strength. The variational
problem of finding an extremum (in prac- method pioneered by Hylleraas (1929)
tice, a minimum) of the energy functional and further developed by Pekeris (1958,
E = ψ | H | ψ / ψ | ψ by varying |ψ 1959) achieves, however, a virtually exact
until E reaches a stationary point at an result for the ground-state energy, E =
approximate eigenvalue. The wave function 2.903 724 377 03 . . . EH , for which the accu-
at the minimum is an approximate solution racy is no longer limited by the method
to the Schrödinger equation. In practice, of solution but rather by the approxi-
one chooses functional forms called trial mate nature of the Hamiltonian itself
functions ψ(r) that depend linearly and/or (for example, the neglect of the finite
24 1 Quantum Mechanics

E Fig. 1.9 Radial densities r 2 Rnl (r) for n = 1, 2,


r 2R2n,l(r), l = 0
3 and l = 0, 1, and 2 of the hydrogen atom.
0

−0.1

−0.2

−0.3

−0.4

−0.5

E r 2R2n,l(r), l = 1
0

−0.05

−0.1

E
r 2R2n,l(r), l = 2
0

−0.05

−0.1

r
0 5 10 15 20 25

size of the nucleus and of relativistic perturbation gV,


effects).
H = H0 + gV (89)

1.7.3 which is nonseparable, but ‘‘small’’ such


Perturbation Theory that it causes only a small modification
of the spectrum and of the eigenstates of
A less accurate method, which is, however, H0 . The control parameter g (0 ≤ g ≤ 1)
easier to implement, is perturbation theory. is used as a convenient formal device for
The starting point is the assumption that controlling the strength of the perturbation
the Hamiltonian H can be decomposed and keeping track of the order of the
into a ‘‘simple,’’ separable (or, at least, perturbation. At the end of the calculation,
numerically solvable) part H0 plus a g can be set equal to unity if V itself is small.
1.7 Solution of the Schrödinger Equation 25

Two cases can be distinguished: stationary Alternative perturbation series can be


perturbation theory and time-dependent devised. For example, in the Brillouin–
perturbation theory. Wigner expansion, the perturbed E 0 (g),
rather than the unperturbed energy E 0 ,
1.7.3.1 Stationary Perturbation Theory appears in the energy denominator of Eq.
The Rayleigh–Schrödinger perturbation (94). The energy is therefore only implicitly
theory involves the expansion of both the given by this equation.
perturbed energy eigenvalue E 0 (g) and the
perturbed eigenstate |ψ 0 (g) in powers of 1.7.3.2 Time-Dependent Perturbation
the order parameter g (or, equivalently, in Theory
powers of V), Time-dependent perturbation theory deals
with the iterative solution of the system of
E0 (g) = E0 + gE (1) + g 2 E (2) + · · · coupled equations, Eq. (69), representing
|ψ0 (g) = |ψ0 + g|ψ (1) + g 2 |ψ (2) + · · · the time-dependent Schrödinger equation.
One assumes again that the Hamiltonian
(90)
consists of an unperturbed part H0
Inserting Eq. (90) into the Schrödinger
and a time-dependent perturbation V(t).
equation
Expansion in the basis of eigenstates of the
unperturbed Hamiltonian (Eq. 68) yields a
H(g)|ψ0 (g) = E0 (g)|ψ0 (g) (91)
matrix H consisting of a diagonal matrix
i | H0 | j = E j δ ij and a time-dependent
yields recursion relations ordered in perturbation matrix V ij = i | V(t) | j .
ascending powers of g for the successive The dynamical phases associated with
corrections to E 0 (g) and |ψ 0 (g) . In prac- the unperturbed evolution can be easily
tice, the first few terms are most important: included in the perturbative solution of Eq.
Thus the first-order corrections for a non- (69) by making the phase transformation
degenerate eigenvalue E 0 are
aj (t) = cj (t)e−iEj t/ (95)
E (1) = ψ0 |V|ψ0 (92)
This transformation, often referred to
and as the transformation to the interaction

ψn |V|ψ0 representation, removes the diagonal matrix
|ψ (1) = |ψn (93) i | H0 | j from Eq. (69), resulting in the
E0 − En
n=0
new system of equations
where the sum over {|ψ n n=0 } extends d
over the complete set of eigenstates of i C(t) = VI (t) C(t) (96)
dt
the unperturbed Hamiltonian excluding
the state |ψ 0 for which the perturbation with matrix elements
expansion is performed. The second-order  
correction to the eigenenergy is i
i|VI (t)|j = exp t(Ei − Ej ) i|V(t)|j


| ψ0 |V|ψn |2 (97)
E (2) = ψ0 |V|ψ (1) = and C = (c1 ,c2 , . . . ,cN ). The perturbative
E0 − En
n=0
solution of Eq. (96) proceeds now by
(94) assuming that initially, say at t = 0, the
26 1 Quantum Mechanics

f (ω, t ) Fig. 1.10 The function f (ω, t) as a


function of ω.

t2

−6π −4π −2π 2π 4π 6π


0 ω
t t t t t t

system is prepared in the unperturbed state with ωij = (εi − εj )/. Plotted as a function
|j with unit amplitude, cj (0) = 1. Since of ω, the function f (ω,t) = 4 sin2 (ωt/2)/ω2
the perturbation is assumed to be weak, (Figure 1.10) is strongly peaked around
the system at a later time t will still be ω = 0. The peak height is t2 and the
approximately in this state. However, small width of the central peak is given by
admixtures of other states will develop with ω  2π/t. In other words, f describes
amplitudes the energy–time uncertainty governing
the transition: The longer the available

1 t  time, the more narrowly peaked is the
ci (t) = dt i|VI (t )|j cj (t )
i 0 transition probability around ωi − ωj = 0,

1 t  that is, for transition into states that are
≈ dt i|VI (t )|j (98)
i 0 almost degenerate with the initial state.
In the limit t → ∞, the time derivative of
Therefore, to first order, the time- Eq. (100), that is, the transition probability
dependent perturbation induces during the per unit time, or transition rate w, has the
time interval [0, t] transitions |j → | i with well-defined limit
probabilities
d 2π
 2 w= Pi (t)t−→∞ = | i|V|j |2 δ(εi − εj )
1 t dt 
Pi (t) = |ci (t)|2 = 2 dt i|VI (t )|j (101)
 0
where the representation of the δ function,
(99)
πδ(ω) = limt→∞ (sinωt/ω), has been used.
In the special case that V is time
Equation (101) is often called Fermi’s golden
independent, Eq. (99) gives
rule and provides a simple, versatile tool
iωij t to calculate transition rates to first order.
1 2 |e − 1|2
Pi (t) = | i|V|j | The physical meaning of the δ function is
2 ωij2
such that in the limit t → ∞ only those
1 4 sin2 (ωij t/2) transitions take place for which energy is
= | i|V|j |2 (100)
 2 ωij2 conserved, as we would have expected since
1.8 Quantum Scattering Theory 27

the interaction in Eq. (100) was assumed to function. Nevertheless, most of the time
be time-independent (after being switched one can manipulate scattering states using
on at time t = 0). For example, in scattering the same rules valid for Hilbert-space vec-
processes, many final states with energies tors (see Section 1.6) except for replacing
very close to that of the initial states are the ordinary orthonormalization condition,
accessible. The δ function is then evaluated Eq. (37), by the δ normalization condition,
within the sum (or integral) over these final Eq. (102).
states. The strategy for solving the Schrödinger
equation, which governs both bounded
and unbounded motion, is different for
1.8 scattering states. Rather than determining
Quantum Scattering Theory the energy eigenvalues or the entire
wave function (which is, in many cases,
Scattering theory plays, within the frame- of no interest), one aims primarily at
work of quantum mechanics, a special determining the asymptotic behavior of
and important role. Its special role is the wave function emerging from the
derived from the fact that scattering deals interaction region, which determines the
with unbound physical systems: two (or flux of scattered particles. The central
more) particles approach each other from quantity of both classical and quantum
infinity, scatter each other, and finally scattering is the cross section σ , defined as
separate to an infinite distance. Obvi-
ously, one cannot impose upon the wave
function the boundary condition ψ(r = number of scattered particles
|r2 − r2 | →∞) → 0 since the probability per unit time
σ =
number of incident particles
of finding the scattered particles at large
per unit area and per unit time
interparticle distances r is, in fact, nonva-
(103)
nishing. This has profound consequences: A cross section σ has the dimension of
energies are not quantized since the corre- an area and can be visualized as an effec-
sponding boundary condition is missing. tive area that the target presents to the
The energy spectrum consists of a con- incoming projectile. Using the probabilis-
tinuum instead of a set of discrete states. tic description of quantum mechanics, one
The states describing the unbound motion, can replace in the definition Eq. (103) all
the ‘‘scattering states,’’ are not square particle currents by corresponding prob-
integrable and do not, strictly speaking, ability currents. Accordingly, σ can be
belong to the Hilbert space. They can expressed in terms of the transition rate
only be orthonormalized in terms of a δ w as
function,
w
 σ = (104)
ψE |ψE = δ(E − E ) (102) j

Mathematically, they belong to the space where j = |ψ | 2 v represents the incoming


of distributions and can be associated with probability current density. For an incom-
states in Hilbert space only upon form- ing plane particle wave (Eq. 4), j = k/M
ing ‘‘wave packets,’’ that is, by convoluting (with M the reduced mass and the ampli-
the scattering state with a well localized tude A set equal to unity). Calculations of
28 1 Quantum Mechanics

total cross sections σ proceed typically by and vanishes sufficiently rapidly at large
first calculating differential cross sections distances for the matter wave to become a
for scattering into a particular group of final plane wave. In such cases, the matrix ele-
states and then summing over all possible ment in Eq. (107) can be expressed as the
final states. For elastic scattering of parti- Fourier component of the potential,
cles with incident wave vectors kin pointing 
in the z direction, a group of final states kout |V|kin = d3 r eir · (kout −kin ) V(r)
is represented by a wave vector kout of the
(108)
outgoing particles pointing in a solid angle
centered around the spherical angles (θ , φ),
For a Coulomb potential V(r) = Z1 Z2 e2 /
4πε0 r representing two charged particles
dσ M with charges Z1 e or Z2 e scattering each
= wk −→kout (105)
d kin in other, Eq. (107) yields – coincidentally –
the Rutherford cross section
Equation (105) gives the differential cross  
section per unit solid angle. dσ Z12 Z22 e4
= (109)
d R 16E (4π ε0 )2 sin4 (θ/2)
2

1.8.1
Born Approximation This expression provided essential clues
as to an internal structure and charge
The evaluation of dσ /d is most straight- distribution of atoms. Two remarkable
forward in the Born approximation. The features should be noted: Prior to the
first Born approximation is nothing but advent of quantum theory, Rutherford
the application of Fermi’s ‘‘golden rule,’’ (1911) used classical mechanics instead of
Eq. (101), to scattering: quantum mechanics to derive Eq. (109). He
could do so since, for Coulomb scattering,
 ∞
dσ 2π M k2out classical and quantum dynamics happen to
= 2 dkout (106)
d  kin 0 (2π )3 agree. Keeping in mind that the classical
| kout |V|kin |2 × δ(Ein − Eout ) limit entails the limit  → 0, the QM
and classical cross sections should agree,
where the integral is taken over the since (dσ /d)R does not depend on .
magnitude of kout but not over the The equivalence of the classical and the
solid angle and where k2out dkout d/(2π)3 QM Rutherford cross sections is due to
represents the number of final states the long range (in fact, infinite range)
available for the scattered particles in of the Coulomb potential. In this case,
the interval dkout d. Since k2out dkout = the de Broglie wavelength of the particle
(M/2 )kout dEout , we have wave is negligibly small compared to the
‘‘size’’ of the target at all energies. A
dσ M2 second coincidence lies in the fact that
= 4 | kout |V|kin |2 (107) the first Born approximation yields the
d  (2π )2
exact quantum result even though the
with |kout | = | kin |. The first Born approx- Born approximation is, strictly speaking,
imation to elastic scattering, Eq. (107), is not even applicable in the case of a
only valid if the potential V is sufficiently Coulomb potential. The infinite range of
weak to allow for a perturbative treatment the Coulomb potential distorts the matter
1.8 Quantum Scattering Theory 29

wave at arbitrarily large distances so that potential V(r), the wave function will be
it never becomes the plane wave that is modified within the region of the nonvan-
assumed in Eq. (108). It turns out that ishing potential. At large distances r → ∞,
all high-order corrections to the Born however, when V(r) → 0, the wave func-
approximation can be summed up to give tion will again become u(r) = sin(kr+δ 0 )
a phase factor that drops out from the since in this region it must satisfy Eq.
expression for the cross section. (111). The only effect that the potential
can have on the scattered wave function
1.8.2 at large distances is to introduce a phase
Partial-Wave Method shift δ 0 relative to the unperturbed
wave.
For elastic scattering by a spherically sym- Simple arguments give a clue as to the
metric potential, that is, a potential that origin and the sign of the phase shift. For
depends only on the distance between the an attractive potential  (negative V), the
particles, V(r) = V(|r|), phase-shift analy- wave number k(r) = 2M(E − V(r))/2
sis provides a versatile nonperturbative increases locally and the de Broglie
method for calculating differential and wavelength λ(r) = 2π/k(r) becomes shorter
integral cross sections. For each angular compared with that of a free particle of
momentum l (which is a good quantum the same energy. The phase of the matter
number for a spherically symmetric poten- wave is therefore advanced compared with
tial), the phase shift for a radial wave the wave of a free particle and δ 0 is
function u(r) = rR(r) can be determined positive (Figure 1.11). Similarly, if V
from the radial Schrödinger equation is a repulsive potential, λ increases in
the interaction region causing a phase
  delay and, hence, a negative phase shift
2 d 2 2 l(l + 1) δ 0 . Including all angular momenta, or
− + + V(r) u(r)
2M dr 2 2M r 2 all ‘‘partial waves,’’ the differential cross
= Eu(r) (110) section can be expressed in terms of the
partial-wave phase shifts δ l as
The idea underlying the phase-shift anal-
∞ 2
ysis can be best illustrated for the phase

dσ 1
shift δ l for l = 0. In absence of the scat- = 2 (2l + 1)e sin δl Pl (cos θ )
iδl
d k
tering potential V(r), Eq. (110) represents l=0

a one-dimensional Schrödinger equation (112)


A particularly simple expression can be
of a free particle in the radial coordinate,
found for the total cross section:

2 d 2 4π

− u(r) = Eu(r) (111) σ = (l + 1) sin2 δl (113)


2M dr 2 k2
l=0

whose solution is u(r) = sin(kr) with k2 = For short-ranged potentials, δ l decreases


2ME/2 . Since the matter wave cannot with increasing l such that only a small
reach the unphysical region of negative number of terms must be included in
r, we have to impose the boundary condi- this sum. The scattering phase shifts δ l (E)
tion u(0) = 0, which excludes the alternative depend sensitively on the energy E of the
solution ∼cos kr. In the presence of the incident particle.
30 1 Quantum Mechanics

Fig. 1.11 Comparison between


δ0 the scattered wave in an
Free–particle wave attractive potential and a
1
Scattered wave free-particle wave (the scattering
phase δ 0 is shown modulo π ).

−1
Potential

0 10 20
r

1.8.3 the rapid change of the scattering phase


Resonances near the resonance can be modeled by

Often, one phase shift δ l (E) increases by



π within a narrow energy range. This δl (E)  tan−1 (115)
2(Eres − E)
corresponds to a phase shift of one-half
of a wave length, or to the addition of and that the energy dependence of the non-
one node to the wave function. A sudden resonant (smooth part) of the phase shift
increase of the scattering phase by π is can be neglected. The term  denotes the
a signature of a ‘‘resonance’’ in the cross width over which the phase jump occurs.
section. A resonance is characterized by a Inserting Eq. (115) into the lth term of Eq.
sharp maximum in the partial-wave cross (113) yields, for the shape of the resonant
section at the energy Eres = 2 k2res /2M: cross section, the Breit–Wigner formula
 

(σl )max = (2l + 1) (114)
k2res  2 /4
σl (E) = (σl )res (116)
(Eres − E)2 +  2 /4
where the phase shift passes through π/2
(or an odd multiple thereof). Resonances The shapes of resonant cross sections
appear when the scattering potential allows become more complicated when inter-
for the existence of ‘‘virtual levels’’ or ferences between the resonant and the
quasi-bound states inside the potential nonresonant parts of the scattering ampli-
well, whose probability density can leak tude come into play. The cross section
out by tunneling through the potential can then exhibit a variety of shapes
well. The additional node that the scattering (Beutler–Fano profile; Fano, 1961), includ-
wave function acquires as the energy passes ing asymmetric peaks or dips (‘‘win-
through E res reflects the presence of a dow resonance’’) depending on the rel-
quasi-bound state. An isolated resonance ative phases of the amplitudes. In the
can be parametrized by a formula due limit that a large number of reso-
to Breit and Wigner. One assumes that nances with energies E res,i (i = 1, . . . ,N)
1.9 Semiclassical Mechanics 31

are situated within the width  (‘‘over- can be written as


lapping resonances’’), the cross section
 
displays irregular fluctuations (‘‘Ericson iS(r)
ψ(r) = A exp (118)
fluctuations’’) (see Chapter 10). These 
play an important role in scattering
in complex systems, for example, in where the unknown function S(r) replaces
nuclear compound resonances (Ericson the exponent k · r of a plane wave (Eq.
and Meyer-Kuckuk, 1966) as well as in 4) for a free particle in the presence of
scattering in chaotic systems Burgdörfer et the potential. The idea underlying the
al., 1995). semiclassical approximation is that in the
limit  → 0, or likewise, in the limit of small
de Broglie wavelength λ → 0 (compared
1.9
with the distance over which the potential
Semiclassical Mechanics
changes), ψ(r) should behave like a plane
wave with, however, a position-dependent
While the correspondence principle pos-
de Broglie wavelength. Inserting Eq. (118)
tulates that in the limit of large quan-
into Eq. (117) leads to
tum numbers (or, equivalently,  → 0),
quantum mechanics converges to classi-
cal mechanics, the approach to this limit 1 i 2
E= [∇S(r)]2 + V(r) − ∇ S(r)
is highly nonuniform and complex. The 2M 2M
(119)
description of dynamical systems in the
with p = ∇S. Equation (119) is well
limit of small but nonzero  is called
suited for taking the classical limit  → 0
semiclassical mechanics. Its importance is
and yields the classical Hamilton–Jacobi
derived not only from conceptual but also
equation. In this limit, S is the action or
from practical aspects since a calculation
Hamilton’s principal function. Quantum
of highly excited states with large quan-
corrections enter through the term linear
tum numbers involved becomes difficult
in . Approximate solutions of Eq. (119)
because of the rapid oscillations of the
can be found by expanding S in ascending
wave function (see Figure 1.7). Semi-
powers of . The apparent simplicity of the
classical mechanics has been developed
classical limit is deceiving: the principal
only recently. The old Bohr–Sommerfeld function S appears inside the exponent Eq.
quantization rules that preceded the devel- (118) with the prefactor −1 and this causes
opment of quantum mechanics are now the wave function to oscillate infinitely
understood as an approximate form of rapidly as  → 0.
semiclassical mechanics applicable to sep-
arable (or integrable) systems. 1.9.1
The wave function of the time-indepen- The WKB Approximation
dent Schrödinger equation for a single
particle, For problems that depend only on one
coordinate, retention of the first two terms
  of the expansion of Eq. (119),
2
Eψ(r) = − ∇ 2 ψ(r) + V(r)ψ(r)
2M
(117) S(x) = S0 (x) + S1 (x)
32 1 Quantum Mechanics

with traveling wave:

 A
x ψ WKB (x) =  sin
S0 (x) = ± 
p(x )dx  |p(x)|
  x 
1 π
× p(x )dx + (122)
S1 (x) = 12 i ln|p(x)| + c (120)  a 4

Here a is the left turning point, the


leads to the Wentzel–Kramers–Brillouin boundary of the classically allowed region
(WKB) approximation where c is an (Figure 1.7). The additional phase of π/4
integration constant to be determined by is required to join ψ(x) smoothly with
normalization of the wave function, the exponentially decaying solution in the
classically forbidden region where p(x)
   becomes purely imaginary (‘‘tunneling’’).
A ±i x
With equal justification, we could have
ψ WKB (x) =  exp p(x )dx
|p(x)|  started from the classically forbidden
(121) region to the right of the turning point
Several features of Eq. (121) are worth not- b and would find
ing: the WKB wave function is an asymp-
A
totic solution of Schrödinger’s equation in ψ WKB (x) =  sin
the limit of small λ and satisfies all rules of |p(x)|
  
QM, including the superposition principle. 1 b π
× p(x )dx + (123)
All quantities in Eq. (121) can be calculated,  x 4
however, purely from classical mechan-
ics. In particular, the quantum probability The uniqueness of ψ WKB (x) mandates the
density |ψ WKB (x) | 2 can be directly associ- equality of Eqs (122) and (123). The two are
ated with the inverse classical momentum equal if the integral over the whole period
∼ | p(x) | −1 or, equivalently, the time the of the trajectory (from a to b and back to a)
particle spends near x. This result was equals
already anticipated for the highly excited   a 
b
state of the harmonic oscillator (Figure 1.7).
+ p(x)dx = p(x)dx
Furthermore, S0 (x) is a double-valued func- a b
tion of x corresponding to the particle = (n + 12 )2π  (124)
passing through a given point moving
either to the right or to the left with momen- Equation (124) can be recognized as the
tum ±p(x). It is the multiple-valuedness of Bohr–Sommerfeld quantization rule of the
p(x) or S(x) that is generic to semiclas- ‘‘old’’ quantum theory put forward prior
sical mechanics and that poses a funda- to the development of quantum mechan-
mental difficulty in generalizing the WKB ics. Quantization rules were successful in
approximation to more than one degree of exploring the discrete spectrum of hydro-
freedom. gen, one of the great puzzles that classical
For describing bound states, a linear dynamics could not resolve. The additional
combination of ψ WKB with both signs of term 12 × 2π in Eq. (124) was initially
the momentum in Eq. (121) is required introduced as an empirical correction. Only
in order to form a standing rather than a within the semiclassical limit of QM does
1.9 Semiclassical Mechanics 33

Fig. 1.12 Confinement of trajectories of an


integrable Hamiltonian for constant energy E
to the surface of a torus, shown here for a
system with two degrees of freedom (N = 2)
Path A
(after Noid and Marcus, 1977).

Path B

its meaning become transparent: it is the is well defined only under special circum-
phase loss in units of h = 2π of a matter stances, that is, when r p(r) is smooth and
wave upon reflection at the turning points the action S(r0 , r) = r0 p · dr is either sin-
(in multidimensional systems, at caustics). gle valued or (finite-order) multiple valued
These phase corrections are of the form (called a Lagrangian manifold; Gutzwiller,
α(π/4) with α = 2 in the present case. The 1990). Einstein pointed out, as early as
quantity α is called the Maslov index and it 1917, that the smoothness applies only to
is used to count the number of encounters integrable systems, i.e., systems for which
with the caustics (in general, the number the number of constants of motion equals
of conjugate points) during a full period of the number of degrees of freedom, N. For
the classical orbit. such a system, classical trajectories are con-
fined to tori in phase space (as shown in
1.9.2 Figure 1.12 for N = 2) and give rise to
The EBK Quantization smooth vector fields p(r) (Figure 1.13). The
corresponding quantization conditions are
The generalization of the WKB approxi-
mation to multidimensional systems faces   α
fundamental difficulties: the momentum p(r) · dr = 2π  n + (i = 1, . . ., N)
vector p(r) is, in general, not smooth but it is Ci 4
a highly irregular function of the coordinate (126)
vector r and takes on an infinite number where Ci (i = 1, . . . , N) denote N topolog-
of different values. Take, for example, a ical distinct circuits on the torus, that is,
billiard ball moving in two dimensions on circuits that cannot be smoothly deformed
a stadium-shaped billiard table. A ball of into each other without leaving the torus
a fixed kinetic energy T = p2 /2M can pass (for example, the paths A and B in
through any given interior point r with Figure 1.12). The conditions of Eq. (126)
a momentum vector p pointing in every are called the Einstein–Brillouin–Keller
possible direction. Consequently, the gen- (EBK) quantization rules and are the gen-
eralization of Eq. (121), eralization of the WKB approximation
  r  (Eq. 124). The amplitude of Eq. (125) is
 i
ψ(r) = |D(r)| × exp p(r ) · dr given by the projection of the point den-
 r0 sity of the torus in phase space onto the
(125) coordinate space. Since the point density is
34 1 Quantum Mechanics

Fig. 1.13 Regular trajectory in coordinate space for


an integrable system. The vector field p(r) is multiple
2 valued but smooth (after Noid and Marcus, 1977).

Y 0

−2

−2 0 2 4
X

uniform in the classical angle variables θ i semiclassical Van Vleck propagator (Van
conjugate to the action (Goldstein, 1959), Vleck, 1928),
the amplitude is given by
K SC (r1 , r2 , t2 − t1 )

∂θ1 ∂θ2 ∂θ3 = |D(r1 , r2 , t2 − t1 )|
|D(r)| ∝ det (127)   t 
∂x∂y∂z i 2 iαπ
×exp L(r(t ))dt − (129)
 t1 2
For a system with one degree of freedom
(N = 1), Eq. (125) reduces to Eq. (121) since where L is the classical Lagrange function
for the particle traveling along the classi-
cally allowed path from r1 at time t1 to
1/2 1/2
∂θ r2 at time t2 and α counts the number of
= ∂t ∂θ
∂x ∂x ∂t singular points of the amplitude factor D
(similar to the Maslov index). If more than
M 1/2 1

= ω ∝ 1/2 (128) one path connects r1 (t1 ) with r2 (t2 ), the
p |p|
coherent sum of terms of the form in Eq.
(129) must be taken. The Van Vleck prop-
1.9.3 agator was recognized only 20 years later
Gutzwiller Trace Formula (Feynman, 1948) as the semiclassical limit
of the full Feynman quantum propagator
As already anticipated by Einstein (1917),
tori do not exist in many cases of K(r1 , r2 , t2 − t1 )
   t2 
practical interest – for example, for the i
motion of an electron in a stadium-shaped = {Dr} exp L(r, t )dt (130)
 t1
semiconductor heterostructure (Marcus et
al., 1992). WBK or EBK approximations are where {Dr} stands for the path integral,
not applicable since the classical motion that is the continuous sum over all
of such an electron is chaotic, destroying paths connecting r1 and r2 including all
the smooth vector field p(r) as in the those that are classically forbidden. In
billiard discussed above. The semiclassical the semiclassical limit  → 0, the rapid
mechanics for this class of problems was oscillations in the phase integral cancel all
developed only recently (Gutzwiller, 1967, contributions except for those stemming
1970, 1971). The starting point is the from the classically allowed paths for which
1.10 Conceptual Aspects of Quantum Mechanics 35

the actions Ldt possess an extremum or, orbit when the initial condition is ever so
equivalently, for which the phase in Eq. slightly displaced (Gutzwiller, 1990). The
(130) is stationary. smooth average density of states,
Equation (129) provides a method for  
investigating the semiclassical time evo- d V(E)
nav (E) = (133)
lution of wave packets in time-dependent dE (2π )N
problems. The Fourier–Laplace transform
of Eq. (129) with respect to t yields the is given by the rate of the change of the
semiclassical Green’s function G(r1 , r2 , E). classical phase-space volume with energy
The spectral density of states of a quantum in units of the volume of Planck’s unit
system can be shown to be cell (2π)N associated with each quantum
state. The derivation of Eq. (132) assumes

n(E) = δ(E − Ei ) two degrees of freedom and makes explicit


i use of the fact that the physical system is
1 completely chaotic and integrable regions
= − Tr Im G(r1 = r2 , E) (131)
π are absent. Only (hyperbolic) unstable
periodic orbits are assumed to exist,
where the trace implies integration over the each with a positive Lyapunov exponent
variables r1 , r2 with constraint r1 = r2 . Note λi > 0 that measures the instability of
that each of these eigenenergies causes each periodic orbit. Equation (132) is
a sharp peak (resembling the singularity therefore complementary to the EBK
of the δ function) in n(E). Calculation of quantization formula (Eq. 126). This
n(E) therefore allows the determination of formula of ‘‘periodic orbit quantization’’ is
eigenenergies of quantum systems. It is an asymptotic series, that is, the sum over
possible to evaluate Eq. (131) in the semi-
all periodic orbits is, in general, divergent.
classical limit by using the semiclassical
For several systems, the sum over the first
Green’s function and evaluating the trace
few orbits has been shown to reproduce
in the limit  → 0 by the stationary-phase
the position of energy levels quite well.
method. The result is the trace formula
The mathematical and physical properties
1 of the Gutzwiller trace formula are still
n(E) = nav (E) + Im a topic of active research in quantum
iπ

Ti mechanics.
×
2 sinh(Ti λi /2)
periodic
orbits
  
Si π 1.10
× exp i − αi (132) Conceptual Aspects of Quantum Mechanics
 2

The sum extends over all unstable classical QM has proved to be an extraordinarily
periodic orbits, each characterized by the successful theory and, so far, has never
period T i , the instability (or ‘‘Lyapunov’’) been in contradiction to experimental
exponent λi , and the action Si , all of which observation. The Schrödinger equation
are functions of the energy E. An orbit has cleared every test carried out so far,
is unstable if its Lyapunov exponent is including a precision test of its linearity:
positive. The term λi measures the rate of the upper bound for corrections due to
exponential growth of the distance from the nonlinear terms in the wave functions,
36 1 Quantum Mechanics

ψ(x)n (n > 1), in the Schrödinger equation of the system. The point to be emphasized
stands currently at one part in 1027 is that it is not our incomplete knowledge
(Majumder et al., 1990). Even so, the of an independently existing objective real-
conceptual and philosophical aspects of ity, but the fate of the cat itself, that can
QM have remained a topic of lively only be resolved by the measurement. Such
discussion stimulated, to a large part, a measurement is therefore, quite literally,
by the counterintuitive consequences of a ‘‘demolition’’ measurement leading to
QM that apparently defy common-sense the ‘‘collapse’’ of the wave function to one
interpretation. The debate, which can be of the two possible states. In hypotheti-
traced to the origins of QM and Einstein’s cal thought experiments for macroscopic
objections summarized in his famous systems such as Schrödinger’s cat prob-
quote ‘‘Gott würfelt nicht’’ (‘‘God does not lem, the average over a huge number
play dice’’), is centered around two aspects: (∼1023 ) of uncontrolled degrees of free-
the superposition principle, Eq. (15), and dom will, for all practical purposes, destroy
the probability interpretation of the wave any well-definded phase relation or coher-
function, Eq. (39). ence between the ‘‘states’’ dead and alive.
In contrast, the axioms of QM can be put to
1.10.1 a rigorous test for microscopic systems, for
Quantum Mechanics and Physical Reality example, by performing correlation mea-
surements on two particles emitted from
The counterintuitive aspects of the super- the same source. Examples along these
position principle are highlighted by the lines were, for example, put forward by
thought experiment of ‘‘Schrödinger’s cat.’’ Einstein, Podolsky and Rosen (1935) to
The coherent superposition of two states, challenge the purely probabilistic interpre-
each of which is presumably an eigen- tation of QM.
state of an extremely complex many-body Consider the emission of two elec-
Schrödinger equation, describes the state trons from, for example, a helium atom
of the cat, being either dead, |d , or alive, (Figure 1.14). The spin of each electron
|a : i = 1,2 can be in a coherent superposition
state of spin up |+ and spin down |− ,
|ψcat = cd |d + ca |a (134)
|ψi = αi |+ + βi |− (136)
with |cd | 2 +|ca | 2 = 1. Only by way of
measurement, that is, by projection onto
relative to one particular Cartesian axis,
the states |a or |d ,
for example, the z axis. One possible
spin state of the electron pair (ignoring
Pa = | a|ψcat |2 , Pd = | d|ψcat |2 (135)
here the antisymmetrization requirement
for fermions) would be an uncorrelated
can we find out about the fate of the cat.
product state,
While the meaning of two sharply different
outcomes of Eq. (135), ‘‘dead’’ or ‘‘alive,’’ is
deeply engraved in our classical intuition, |ψ = |+ 1 |− 2 (137)
the quantum state of a coherent super-
position, Eq. (134), defies any classical Already Schrödinger pointed out that
interpretation. Yet it is an accessible state quantum theory allows to be alternatively
1.10 Conceptual Aspects of Quantum Mechanics 37

z z

y
x ++
e He e
L1 L2
x
Fig. 1.14 Two-electron emission from helium with the
electrons being spin-analyzed in two Stern–Gerlach-type
filters with different orientations (along the x and z axis,
respectively). The difference in length, L1 < L2 , makes the
spin selection in the left detector happen first (inducing a
‘‘wave function collapse’’) and in the right detector later.

in an entangled (or quantum-correlated) α = −β = 1/ 2,
superposition state,
1
|ψ = √ (|+ 1 |− 2 − |− 1 |+ 2 ) (139)
2
|ψ = α|+ 1 |− 2 + β|− 1 |+ 2 (138)
is rotationally symmetric and equally
applies when we inquire about the spin
In fact, the electron pair in the (singlet) orientation along the x rather than along
ground state of helium with total spin S = 0 the z axis. If we postselect the spin-up
forms such√ an entangled pair with α = direction (+) along the x axis for the first
−β = 1/ 2. The hallmark of entangled particle, the second particle will have spin
states of the form Eq. (138) is that they down (−) along the same axis. If we,
cannot be represented as a factorized instead, measured for the second parti-
product in the form of Eq. (137). cle the polarization relative to the z axis
The consequences of such quantum cor- as before (Figure 1.14), we would find
relations are profound: when measuring both spin up and down with 50% prob-
the spin of particle 1 to be up (+), we ability, and thus a vanishing average spin.
know with certainty that particle 2 has Different projection protocols for the first
its spin down (−). The ‘‘collapse’’ of the spin imply entirely different results for
entangled wave function when measur- the second. The emergence of a specific
ing particle 1 forces particle 2 to switch ‘‘physical reality’’ out of a specific mea-
its spin from a completely undetermined surement context is often referred to as
state to a well-defined value. We therefore quantum contextuality and is unknown in
obtain information about the orientation classical physics. It has therefore been
of spin 2 without actually performing a questioned whether the quantum mechan-
measurement on this particle. This corre- ical description of a state such as in Eq.
lation, often referred to as the nonlocality (138) can be considered a complete pic-
of QM, is a consequence of the probabilis- ture of physical reality (Einstein, Podolsky
tic description and superposition of wave and Rosen, 1935). In particular, it was
functions in QM. The emergence of a defi- conjectured that particles may have addi-
nite physical property or physical reality by tional properties that are underlying the
‘‘postselection’’ through the measurement quantum-mechanical picture but which are
of an entangled state goes even further. For hidden from our view. Such ‘‘hidden vari-
example, the entangled state Eq. (138) with ables’’ would ensure that physical systems
38 1 Quantum Mechanics

have only properties that are ‘‘real,’’ that of ‘‘nonlocality’’ are untenable (Hasegawa
is, independent of a measurement and et al., 2006; Gröblacher et al., 2007).
that the result of any measurement is
only dependent on ‘‘local’’ system prop- 1.10.2
erties rather than on a remote detection Quantum Information
of another observable. The knowledge of
‘‘hidden variables’’ would allow one to The most fundamental unit of informa-
remove the probabilistic aspects of QM. tion in classical information theory is the
The role of probability in QM could then so-called bit, which can take on only one
be the same as in statistical mechanics: of two mutually exclusive values (typically
a convenient way to quantify our incom- labeled as ‘‘0’’ and ‘‘1’’). In conventional
plete knowledge of the state of a system. computers, all numbers and operations
The hidden-variable hypothesis, or alterna- with them are encoded in this binary sys-
tem. The corresponding quantum analog
tively, the quantum correlation built into
of the bit – named qubit – is given by a
entangled states, Eq. (138), has been probed
two-state system with the two states |+
by testing the violation of Bell’s inequality
and |− from Eq. (136), now labeled as |0
(Bell, 1969; Clauser et al., 1969). Experi-
and |1 . What sets the qubit apart from the
ments with correlated pairs of particles, in
conventional classical bit is precisely the
particular photon pairs – for example, by
ability to be in a coherent superposition of
Aspect and coworkers (1982) – have fal-
both basis states,
sified the hidden-variable hypothesis by
showing that entangled states can violate
|Q = α|0 + β|1 (140)
Bell’s inequality while being in full agree-
ment with the predictions of QM. More
with |α | 2 +|β | 2 = 1. In quantum informa-
recent experiments have succeeded in mea-
tion theory, the intrinsic indeterminacy of
suring entangled photons at a distance of
the qubit’s value before a measurement is
several hundred meters apart and with
exploited as a resource rather than treated
light polarizers (used as photon analyz- as a deficiency.
ers), which switch their axes so rapidly One potential application of qubits is
that during the photons’ time of flight subsumed under the name ‘‘quantum
the polarizers have no chance to exchange computation’’. The basic building block
information on their respective orientation. of a so-called quantum computer would
(The ‘‘collapse’’ of the wave function by be a number of N entangled qubits
measuring one of the two photons occurs that can be in any superposition of
instantaneously for both photons, but can- 2N basis states. An example for N =
not be used to transmit any information 2 is just the maximally entangled pair
and therefore does not violate the relativis- in Eq. (139). While a collection of N
tic causality relations.) Even under such classical bits can only be in a single
stringent conditions, a violation of Bell’s one of the 2N possible states at a
inequality by more than 30 standard devia- given time, N qubits can be in any
tions was observed experimentally (Weihs superposition state and thus can clearly
et al., 1998). Also so-called noncontex- have a much higher information content.
tual hidden-variable theories and realis- Owing to the superposition principle, any
tic models that assume a certain type of the (unitary) logic operations of the
1.10 Conceptual Aspects of Quantum Mechanics 39

quantum computer performed on the A truly random sequence of numbers can


initial state could process all the involved be generated, for example, by repeatedly
2N basis states simultaneously. Several measuring the spin orientation (or photon
algorithms have been suggested (Shor, polarization) of an unpolarized spin, a
1994; Grover, 1996) for how this inherently situation that may be realized by first
quantum feature of ‘‘parallel processing’’ projecting the first spin of an entangled
could be used to solve fundamental pair along a perpendicular direction (see
computational problems much faster, i.e., the example of Fig. 1.14 above).
with much fewer operations than with the Interception of a message shared with
corresponding classical computations. the recipient through an entangled pair
The manipulation of quantum states by a third party would require a mea-
within the framework of quantum com- surement, which leads to a wave function
putation and quantum information pro- collapse and to a detectable reduction in
cessing involves a set of rules as, for quantum correlations. Using the princi-
example, the so-called no cloning theorem ples of QM for the generation and trans-
(Wootters and Zurek, 1982), which states mission of a secure encryption key – a
the impossibility to create identical copies technique that is referred to as quantum
of an arbitrary unknown quantum state. cryptography – has been successfully imple-
Although only imperfect ‘‘clones’’ of a mented for setups involving very large
quantum state are possible, a transfer of distances (144 km) between the sender and
a quantum state from one system to a recipient (Ursin et al., 2007).
second, possibly distant, system is allowed
according to QM. Such a ‘‘teleportation’’ of 1.10.3
a quantum state involves the destruction Decoherence and Measurement Process
of the original state during the teleporta-
tion procedure and is typically performed One fundamental obstacle in practically
through an intermediate entangled state implementing quantum computation and
of three particles (Bennett et al., 1993; quantum information protocols is the chal-
Bouwmeester et al., 1997). lenge to keep the coherent superposition
The laws of QM can also be used for of a quantum state, in particular of entan-
secure information transfer. In classical gled pairs, ‘‘alive’’ for extended periods
cryptography, secret messages are enci- of time. Any unintentional or intentional
phered by way of a code that the recipient of interaction of the system inevitably and
a message must know in order to decipher irreversibly converts the quantum super-
a message’s content. To prevent the break- position into a classical probability distri-
ing of the code by a third unauthorized bution and destroys the operating principle
party, the key for the encoding algorithm is of a quantum information device. The
preferably defined as a random sequence measurement process and the destruction
of numbers, which is generated anew for of coherence, frequently called decoher-
each message transmitted. ence, are closely intertwined. In the early
Challenges to classical cryptography are days of quantum theory, the measure-
the need of a truly random key and the risk ment device that caused the ‘‘collapse’’
of key corruption through the interception of the wave function was considered to
by a third party. Both of theses challenges be a ‘‘classical’’ apparatus decoupled from
can be met by intrinsic quantum features. the quantum dynamics to be observed
40 1 Quantum Mechanics

(Bohr, 1928). The notion of open quan- The expectation value of any physical
tum systems has superseded this view observable represented by an operator AS
and provides a widely accepted unified acting only on the states in the Hilbert
description of measurement and decoher- space of the system S gives in the presence
ence within the realm of quantum theory. of entangled environmental degrees of
The starting point here is the fundamental freedom,
insight that the quantum system (S) to be
observed is never completely isolated from AS = ψ|AS |ψ
the environment (E), which itself is quan- = |α|2S +|AS |+ S
tum rather than classical. The Hamiltonian
+|β|2S −|AS |− S
describing both system and environment
is given by H = HS +HE +HS−E , contain- = Tr(ρAS ) (143)
ing the Hamiltonian of the system (HS )
and its environment (HE ), as well as the with ρ = |α|2 | + S S + | + |β|2 |− S S −|.
interaction between the two (HS−E ). A It is the coherent entanglement with the
state describing both system and envi- environment, Eq. (142), that leads to (par-
ronment, |ψ , even when at the initial tial) decoherence within the system (S) and
time t = 0 prepared in a factorizable state, to a statistical mixture described by the
j
|ψ(t = 0) = |ψSi |ψE , will evolve in the density operator introduced above (Eq. 40)
course of time under the influence of the with probabilities P+ = |α|2 and P− = |β|2 .
interaction HS−E into an entangled, non- The ubiquitous coupling of a physical
factorizeable state system to the environment provides the

natural setting for the transition from
j
|ψ(t) = aij |ψSi |ψE (t) (141) the quantum indeterminism of superpo-
i,j sitions to the definiteness of the classical
measurement. The classical limit induced
j
Here, |ψSi (and |ψE ) are basis states of by decoherence is conceptually different
the system (of the environment), which from the classical limit of quantum sys-
often, but not always, are eigenstates of tems of large quantum numbers, which is
the Hamilton operator of the system HS the focus of semiclassical mechanics (see
(of the environment HE ). ‘‘Tracing out’’ Section 1.9). In spite of recent progress, it
the (unobserved) environment degrees of is generally still considered an open ques-
freedom of the entangled state by taking tion if decoherence concepts alone can fully
expectation values leads to decoherence explain the quantum-to-classical transition
in the (observed) system and mimics and provide definite answers to the widely
the effect of a classical measurement discussed questions regarding the ‘‘quan-
device. tum measurement problem.’’
Consider, as an illustrative example, The above considerations demonstrate
a situation in which the system and that it is very unlikely for a macroscopic
the environment are each represented object such as Schrödinger’s cat to ever be
by a two-dimensional Hilbert space and in a superposition state ‘‘dead and alive’’ as
described by the following entangled state: in Eq. (134) because it is extremely difficult
to isolate the cat from its environment.
For all practical purposes, an exchange of
|ψ(t) = α|+ S |− E + β|− S |+ E (142) particles and heat between the cat and
1.11 Relativistic Wave Equations 41

its environment will prevent the cat, for 1.11.1


itself, to be in a pure quantum state. The Klein–Gordon Equation
It will rather be described by a density
matrix with strongly reduced quantum We use the same recipe that led us in
correlations. As both the cat and the Section 1.3 to the Schrödinger equation,
i.e., we replace the quantities H and p by
environment are macroscopic objects with
differential operators (Eq. 11), motivated
many degrees of freedom, decoherence
by the de Broglie hypothesis. In this
sets in at an extremely short timescale.
case, however, we use the relativistic
Recent experiments have been able to
relation between energy and momentum
sufficiently decouple ‘‘mesoscopic’’ objects (see Chapter 2) instead of the nonrelativistic
as superconducting rings (Friedman et al., one, Eq. (9):
2000) from their environment to observe

these systems in superposition states.
E − V(r) = M2 c4 + p2 c2 (144)

Following this recipe, we are led to


1.11
Relativistic Wave Equations 
d
i ψ(r, t) = −2 ∇ 2 c2 + M2 c4
dt
When the characteristic speed v of particles 
in a physical system becomes comparable +V(r) ψ(r, t) (145)
to the speed of light, v ≤ c, the Schrödinger
equation, which incorporates nonrelativis- Equation (145) has a major drawback:
tic quantum dynamics, fails. Generaliza- it treats the space and time derivatives
tions to relativistic wave equations face, asymmetrically. The Lorentz invariance
however, fundamental difficulties due to of relativistic dynamics requires, however,
the possibility of particle creation and that space and time coordinates be treated
destruction in relativistic dynamics (see on equal footing. We thus start with the
Chapter 2). For example, if photons have square of Eq. (144) and arrive at
energies in excess of twice the relativis-  2

tic rest energy of electrons, E γ > 2me c2 , i − V(r) ψ(r, t)
∂t
a pair consisting of an electron and its
= (−2 ∇ 2 c2 + M2 c4 )ψ(r, t) (146)
antiparticle, a positron, can be created.
Therefore, the interpretation of |ψ(r)|2 as
This is the time-dependent Klein–
a single-particle probability density is no
Gordon equation, applicable to particles
longer possible. A consistent resolution of
with integer spin. The corresponding sta-
the difficulties can be accomplished within
tionary Klein–Gordon equation follows by
the framework of quantum field theory, replacing i∂/∂t by E. Equation (146) gives
which is outside the scope of this chapter. an eigenvalue equation in E 2 rather than
The relativistic analogs to the Schrödinger in E itself. Consequently, upon taking
equation discussed in the following section the square root of the eigenvalue, solu-

are only applicable to relativistic systems in tions with both signs, ± p2 c2 + M2 c4 ,
which particle production and destruction will appear. The physical meaning of
are not yet important. negative-energy solutions is discussed
42 1 Quantum Mechanics

below. For a charged particle with charge q spin 12 particle interacts with an electro-
in the presence of an electromagnetic field, magnetic field as described by minimal
both the scalar potential φ(V = qφ) and the coupling, the stationary Dirac equation
electromagnetic vector potential A must be becomes
included by ‘‘minimum coupling’’:
  [E − qφ(r)]ψ(r)
    
p −→ p − qA = ∇ − qA (147) 
i = cα ∇ − qA + βMc2 ψ(r) (151)
i

With this coupling, the Klein–Gordon Dirac spinors permit the probability-
equation is invariant under gauge trans- density interpretation, that is, |ψ(x)|2 =
4
formations of the electromagnetic field as
i=1 |ψi (x)| . The physical meaning of the
2
well as under Lorentz transformations. four components can be easily understood
by considering a free particle.
 The energy
1.11.2 eigenvalues are E = ± p2 c2 + M2 c4 . In
The Dirac Equation addition to the positive energies we find
negative eigenenergies, which are a feature
Instead of taking the square of Eq. (144), of relativistic quantum mechanics and
Dirac pursued a different approach in which result from the fact that the
developing a Lorentz-invariant relativistic square-root operator permits both the
wave equation. He ‘‘linearized’’ the square positive and negative sign of E. The
root four-spinor can be decomposed into two
   2-spinors ψ = (φ 1 , φ 2 ) one of which
c2 p2 + M2 c4 −→ c α · p + βMc2 represents the positive-energy solution and
(148) the other the negative-energy solution.
Each component of the two-spinor wave
with the subsidiary condition that the function can be understood as a product
square of the linearized operator satisfies of a configuration-space wave function
eikr and a spin 1/2 eigenstate |ms =
 2 ±1/2 . A major success of the Dirac
cα · p + βMc2 = c2 p2 + M2 c4 (149)
theory was that the intrinsic degree of
an electron spin, which was a mere
It turned out that solutions to this equation
add-on within the Schrödinger theory,
can be found only if all α i (i = x, y, z)
is incorporated into the Dirac equation
and β are taken as matrices of minimum
from the outset. Furthermore, the g
dimension 4 × 4. Equation (149) is satisfied
factor for the magnetic moment of the
if the matrices α i , β satisfy the relations
electron as well as the spectrum of the
hydrogen atom can be derived from the
αi αk + αk αi = 2δik Dirac equation to a very high degree of
αi β + βαi = 0 (150) approximation. The pieces still missing
are the quantum electrodynamic (QED)
Accordingly, ψ is now a four-component corrections that lead to small energy
entity (called a spinor), ψ = {ψ i , i = shifts of atomic levels (‘‘Lamb shift’’)
1, . . . , 4}. The resulting Dirac equation and the deviations of the g factor for
describes spin 12 particles. When a charged the anomalous magnetic moment from
1.11 Relativistic Wave Equations 43

Fig. 1.15 Particle–antiparticle production visualized as the excitation of a


E
negative-energy particle to positive energies leaving behind a hole.

E = 2mec 2

2, g = 2 × 1.001 16. These corrections can QM, whereas imperfections in the coher-
be determined only within quantum field ence are described by a density matrix with
theory, the theory of the interactions (partially) classical correlations. Decoher-
of charged particles with the quantized ence stands for the loss of coherence as
electromagnetic field. induced, for example, by coupling an iso-
The negative-energy states, according lated quantum system to its environment.
to Dirac, can be interpreted in terms of Correspondence Principle: The princi-
antiparticle states. The vacuum state of ple that quantum-mechanical and classical
empty space is represented by a fully description of physical phenomena should
occupied ‘‘sea’’ of negative-energy states. become equivalent in the limit of vanishing
While this ‘‘sea’’ itself is not observ- de Broglie wavelength λ → 0, or equiva-
able, ‘‘holes’’ created in the sea by excit- lently, of large quantum numbers.
ing negative-energy electrons to posi-
de Broglie Wave: According to de Broglie,
tive energies (Figure 1.15) are observ-
the motion of a particle of mass M and
able and appear as antiparticles. For
velocity v is associated with a matter wave
example, a missing electron with charge
with wavelength λ = h/(Mv). h is Planck’s
−e, energy −E, and momentum p repre-
quantum.
sents a positron with charge e, energy E,
and momentum −p. The negative-energy Delta (δ) Function: A particular case of a
states of electrons therefore offer a sim- distribution, frequently used in quantum
ple explanation of particle–antiparticle mechanics to formulate the orthogonality
pair production in terms of the excita- and normalization conditions of states with
tion of a negative-energy particle to a a continuous spectrum of eigenvalues.
positive-energy state (the particle) leaving Distributions: Also called generalized func-
a hole in the negative-energy ‘‘sea’’ (the tions. Linear-continuous functionals oper-
antiparticle) behind. ating on a metric vector space. They form
the dual space of the metric vector space.
Glossary In the context of quantum mechanics, the
space is a Hilbert space and the operation of
Coherence and Decoherence: In QM, the functional on an element of the Hilbert
the term coherence stands for the abil- space is called formation of a wave packet.
ity of a particle’s state to interfere. Eigenvalue Problem: The solution of
Perfectly coherent states (‘‘pure states’’) the eigenvalue equation for the matrix
are subject to the superposition principle of A, Aψ = λψ where λ is a number called
44 1 Quantum Mechanics

an eigenvalue and ψ is the eigenvector. Uncertainty Principle: Two canonically


In quantum mechanics, ψ is a vector in conjugate variables, such as position x and
Hilbert space. momentum p, can be simultaneously mea-
Entanglement: An intrinsic feature of QM, sured only with an intrinsic uncertainty of
occurring when particles or subsystems (in at least xp  /2.
either case more than one), are correlated Wave Function: The representation of the
beyond a degree allowed by classical state |ψ in coordinate space, ψ(r) = r | ψ .
physics. Entangled systems, even when far
apart, cannot be described independently
of each other. References
Hamilton–Jacobi Equation: A particu-
Andrä, H. (1974) Phys. Scr., 9, 257–80.
lar canonical transformation in classical Aspect, A., Dalibard, J., and Roger, G. (1982)
dynamics that leads to the vanishing of the Phys. Rev. Lett., 49, 91–94; 1804–807.
Hamiltonian function in the new coordi- (a) Bell, J.S. (1965) Physics (New York), 1,
nates. The transformation equation for the 195–207; (b) see also Clauser, J.F., Horner,
original Hamiltonian function is called the M.A., Shimony, A., and Holt, R. (1969) Phys.
Rev. Lett., 23, 880–83.
Hamilton–Jacobi equation and is equivalent
Bennett, C.H., Brassard, G., Crépeau, C., Jozsa,
to equations of motion. R., Peres, A., and Wootters, W.K. (1993) Phys.
Hilbert Space: Infinite-dimensional com- Rev. Lett., 70, 1895–99.
plete vector space over the complex num- Berry, M. (1984) Proc. R. Soc. London, Ser. A, 392,
45–57.
bers endowed with a metric induced by
Bohr, N. (1928) Nature, 121, 580–90.
a scalar product that satisfies Schwartz’s
Bouwmeester, D., Pan, J.-W., Mattle, K., Eibl, M.,
inequality. Weinfurter, H., and Zeilinger, A. (1997)
Laser: Light amplication by stimulated Nature, 390, 575–79.
emission of radiation. Modern source for Born, M. (1926) Z. Phys., 38, 803–27.
Brandt, S. and Dahmen, H.D. (1989) Quantum
high-intensity coherent electromagnetic
Mechanics on the Personal Computer,
radiation in the infrared, visible, and Springer-Verlag, Berlin.
ultraviolet spectral region. de Broglie, L. (1924) Thesis; reprint (1926), J.
Schrödinger Equation: Fundamental Phys. (Paris), 7, 321–337.
equation governing the dynamics of a Burgdörfer, J. (1981) Phys. Rev. A, 24, 1756–67.
Davisson, C.J. and Germer, L.J. (1927) Phys. Rev.,
nonrelativistic quantum system, playing a
30, 705–40.
similar role in quantum mechanics as do Dirac, P.A.M. (1930) The Principles of Quantum
Newton’s equations of motion in classical Mechanics, Oxford University Press, Oxford,
dynamics. Reprint (1986) Oxford Scientific Publications,
Oxford.
State: A mechanical system is at any given
Einstein, A. (1905) Ann. Phys. (Leipzig), 17,
time completely characterized in quantum 132–48.
mechanics by the state ψ. The projection Einstein, A. (1917) Verh. Dtsch. Phys. Ges., 19,
onto coordinate space, ψ(r), is called a wave 82–92.
function. States are vectors in Hilbert space. Einstein, A., Podolsky, B., and Rosen, N. (1935)
Phys. Rev., 47, 777–80.
Superposition Principle: Any superposi-
Ericson, T. and Mayer-Kuckuk, T. (1966) Annu.
tion in Hilbert space, ψ = aψ 1 + bψ 2 , of Rev. Nucl. Sci., 16, 183–206.
two states ψ 1 and ψ 2 forms another physi- Fano, U. (1961) Phys. Rev., 124, 1866–78.
cally realizable state of a physical system. Feynman, R. (1948) Rev. Mod. Phys., 20, 367–87.
Further Reading 45

Friedman, J.R., Patel, V., Chen, W., Tolpygo, Ursin, R. et al. (2007) Nature Physics, 3, 481–86.
W.K., and Lukens, J.E. (2000) Nature, 406, Van Vleck, J. (1928) Proc. Natl. Acad. Sci. U.S.A.,
43–46. 14, 178–88.
Goldstein, H. (1959) Classical Mechanics, Weihs, G., Jennewein, T., Simon, C., Weinfurter,
Addison-Wesley, Reading, PA. H., and Zeilinger, A. (1998) Phys. Rev. Lett.,
Goudsmit, S. and Uhlenbeck, G. (1925) 81, 5039–43.
Naturwissenschaften, 13, 953–54. Wootters, W.K. and Zurek, W.H. (1982) Nature,
Gröblacher, S., Paterek, T., Kaltenbaek, R., 299, 802–3.
Brukner, C., Zukowski, M., Aspelmeyer, M.,
and Zeilinger, A. (2007) Nature, 446, 871–75.
Grover, L.K. (1996) Proceedings, 28th Annual Further Reading
ACM Symposium on the Theory of
Computing (STOC) 212–19.
Bell, J.S. (1987) Speakable and Unspeakable in
Gutzwiller, M. (1967) J. Math. Phys., 8,
Quantum Mechanics, Cambridge University
1979–2000.
Press, Cambridge.
Gutzwiller, M. (1970) J. Math. Phys., 11,
Bjorken, J.D. and Drell, S.D. (1965) Relativistic
1791–1806.
Gutzwiller, M. (1971) J. Math. Phys., 12, 343–58. Quantum Mechanics and Relativistic Field
Gutzwiller, M. (1990) Chaos in Classical and Theory, McGraw-Hill, New York.
Quantum Mechanics, Springer-Verlag, New Bransden, B.H. and Joachain, C.J. (1983) Physics
York. of Atoms and Molecules, Longman, New York.
Hasegawa, Y., Loidl, R., Badurek, G., Baron, M., Burgdörfer, J., Yang, X., and Müller, J. (1995)
and Rauch, H. (2006) Phys. Rev. Lett., 97, Chaos Solitons Fractals, 5, 1235–73.
230401. Cohen-Tannoudji, C., Diu, B. and Laloë, F.
Heisenberg, W. (1927) Z. Phys., 43, 172–98. (1977) Quantum Mechanics, Vols. I and II,
Hellmuth, T., Walter, H., Zajonc, A. and John Wiley & Sons, New York.
Schleich, W. (1987) Phys. Rev. A, 35, Courant, R. and Hilbert, D. (1953) Methods of
2532–41. Mathematical Physics, Wiley Interscience,
Hylleraas, E. (1929) Z. Phys., 54, 347–66. New York.
Lighthill, M.J. (1958) Introduction to Fourier Eisberg, R. and Resnick, R. (1985) Quantum
Analysis and Generalized Functions, Physics of Atoms, Molecules, Solids, Nuclei and
Cambridge University Press, Cambridge. Particles, 2nd edn, John Wiley & Sons, New
Majumder, P.K., Venema, B.J., Lamoreaux, S.K., York.
Heckel, B.R., and Fortson, E.N. (1990) Phys. Feynman, R.P. and Hibbs, A.R. (1965) Quantum
Rev. Lett., 65, 2931–34. Mechanics and Path Integrals, McGraw-Hill,
Marcus, C.M., Rimberg, A.J., Westervelt, R.M., New York.
Hopkins, P.F., and Gossard, A.C. (1992) Phys. Feynman R.P., Leighton, R.B., Sands, M. (1965)
Rev. Lett., 69, 506–9. The Feynman Lectures on Physics, Vol. III,
Noid, D. and Marcus, R. (1977) J. Chem. Phys., Quantum Mechanics, Addison-Wesley,
67, 559–67. Reading, MA.
(a) Pekeris, C.L. (1958) Phys. Rev., 112, 1649–58; Krane, K. (1983) Modern Physics, John Wiley &
(b) Pekeris C.L. (1959) Phys. Rev., 115, Sons, New York.
1216–21. Landau, L. and Lifshitz, E. (1965) Quantum
Rutherford, E. (1911) Philos. Mag., 21, 669–83. Mechanics, Non-Relativistic Theory, Pergamon,
Schwartz, L. (1965) Méthodes Mathématiques pour Oxford.
Les Sciences Physiques, Hermann, Paris. Merzbacher, E. (1970) Quantum Mechanics, John
Shor, P.W. (1994) Proceedings of the 35th Wiley & Sons, New York.
annual symposium on foundations of Messiah, A. (1961) Quantum Mechanics,
computer science; revised version available North-Holland, Amsterdam.
at: (1997). SIAM J. Comput., 26, 1484–509. von Neumann, J. (1955) Mathematical
(a) Stern, O. and Gerlach, W. (1921) Z. Phys., 7, Foundations of Quantum Mechanics,
249–53; (b) Stern, O. and Gerlach, W. (1921) Princeton University Press, Princeton, NJ.
8, 110–111; (c) Stern, O. and Gerlach, W. Schiff, L. (1968) Quantum Mechanics,
(1922) 9, 349. McGraw-Hill, Singapore.
46 1 Quantum Mechanics

Sommerfeld, A. (1950) Atombau und Wichmann E.H. (1971) Berkeley Physics Course,
Spektrallinien, Vieweg, Braunschweig; Quantum Physics, Vol. 4, McGraw-Hill, New
Reprint (1978) Verlag Harri Deutsch, Thun, York.
Switzerland.

You might also like