mechanics
A Complete Introduction
Alexandre Zagoskin is a Reader in Quantum
Physics at Loughborough University and a
Fellow of The Institute of Physics.
He received his Masters degree in physics
from Kharkov University (Soviet Union),
and his PhD in physics from the Institute for
Low Temperature Physics and Engineering
(FTINT), also in Kharkov. He has worked at
FTINT, Chalmers University of Technology
(Sweden), the University of British Columbia
(Canada) and the RIKEN Institute (Japan).
In 1999, when in Canada, he co-founded the
company D-Wave Systems with the aim of
developing commercial quantum computers,
and was its Chief Scientist and VP of
Research until 2005.
He has published two books: Quantum
Theory of Many-Body Systems (Springer,
1998, 2014) and Quantum Engineering
(Cambridge University Press, 2011).
He and his colleagues are now working
on the development of quantum
engineering the theory, design, fabrication
and characterization of quantum coherent
artificial structures.
Quantum
mechanics
A Complete Introduction
Alexandre Zagoskin
First published in Great Britain in 2015 by John Murray Learning.
An Hachette UK company.
First published in US in 2015 by Quercus
Copyright Alexandre Zagoskin 2015
The right of Alexandre Zagoskin to be identified as the Author of the Work
has been asserted by him in accordance with the Copyright, Designs and
Patents Act 1988.
Database right Hodder & Stoughton (makers)
The Teach Yourself name is a registered trademark of Hachette UK.
All rights reserved. No part of this publication may be reproduced, stored
in a retrieval system or transmitted in any form or by any means, electronic,
mechanical, photocopying, recording or otherwise, without the prior written
permission of the publisher, or as expressly permitted by law, or under terms
agreed with the appropriate reprographic rights organization. Enquiries
concerning reproduction outside the scope of the above should be sent to the
Rights Department, John Murray Learning, at the address below.
You must not circulate this book in any other binding or cover and you must
impose this same condition on any acquirer.
British Library Cataloguing in Publication Data: a catalogue record for this
title is available from the British Library.
Library of Congress Catalog Card Number: on file.
1
The publisher has used its best endeavours to ensure that any website
addresses referred to in this book are correct and active at the time of going
to press. However, the publisher and the author have no responsibility for the
websites and can make no guarantee that a site will remain live or that the
content will remain relevant, decent or appropriate.
The publisher has made every effort to mark as such all words which it
believes to be trademarks. The publisher should also like to make it clear that
the presence of a word in the book, whether marked or unmarked, in no way
affects its legal status as a trademark.
Every reasonable effort has been made by the publisher to trace the copyright
holders of material in this book. Any errors or omissions should be notified
in writing to the publisher, who will endeavour to rectify the situation for any
reprints and future editions.
Cover image Shutterstock.com
Typeset by Cenveo Publisher Services.
Printed and bound in Great Britain by CPI Group (UK) Ltd., Croydon, CR0
4YY.
John Murray Learning policy is to use papers that are natural, renewable
and recyclable products and made from wood grown in sustainable forests.
The logging and manufacturing processes are expected to conform to the
environmental regulations of the country of origin.
Carmelite House
50 Victoria Embankment
London EC4Y 0DZ
www.hodder.co.uk
Contents
acknowledgements x
introduction xi
A fair warning
1 Familiar physics in strange spaces 1
Particles, coordinates, vectors and trajectories
Velocity, acceleration, derivatives and differential equations
Phase space
Harmonic oscillator, phase trajectories, energy
conservation and action
Hamilton, Hamiltonian, Hamiltons equations
and state vectors
Generalized coordinates, canonical momenta,
and configuration space
Statistical mechanics, statistical ensemble and
probability distribution function
Liouville equation*
2 Less familiar physics in stranger spaces 34
Quantum oscillator and adiabatic invariants
Action quantization, phase space and the uncertainty principle
Zero-point energy and complex numbers
State vector and Hilbert space
Operators, commutators, observables and expectation values
Eigenstates, eigenvalues, intrinsic randomness of
Nature and Borns rule
Quantization and Heisenberg uncertainty relations
3 equations of quantum mechanics 66
Poisson brackets, Heisenberg equations of motion
and the Hamiltonian
Matrices and matrix elements of operators
Schrdinger equation and wave function
Quantum propagation
A particle in a box and on a ring: energy and
momentum quantization
Quantum superposition principle
Quantum statistical mechanics, quantum ensembles
and density matrix*
The Von Neumann equation and the master equation*
4 Qubits and pieces 96
Qubits and other two-level quantum systems
Qubits state vector, Hilbert space, Bloch
vector and Bloch sphere
Qubit Hamiltonian and Schrdinger equation
Energy eigenstates and quantum beats
Bloch equation, quantum beats (revisited) and qubit control
Qubit observables, Pauli matrices and expectation values
Qubit density matrix, von Neumann and Bloch
equations and NMR*
Vintage charge qubits
5 Observing the observables 134
Measurement, projection postulate and Borns rule (revisited)
The collapse of the wave function, measurement
problem and quantum-classical transition
Observing a charge qubit
Bloch vector and density matrix; Bloch equations (revisited);
dephasing; and weak continuous measurement*
Measurements and Heisenberg uncertainty relations:
Heisenberg microscope
Standard quantum limit
Quantum non-demolition (QND) measurements
and energy-time uncertainty relation
Quantum Zeno paradox
6 strange and unusual 165
Quantum tunnelling: -decay
Quantum tunnelling: scanning tunnelling microscope
Waves like particles: photoelectric effect and the single-photon
double slit experiment
Particles like waves: diffraction and interference of massive
particles; de Broglie wavelength
7 the game of numbers 191
Electron in an atom
Electrons in an atom and Mendeleev periodic table
Pauli exclusion principle, fermions and bosons, and spin
Quantum many-body systems, Fock states, Fock space and
creation and annihilation operators
Quantum oscillator, Heisenberg equations and energy quanta*
Quantum fields and second quantization
Why classical light is a wave and a piece of metal is not
vi
8 the virtual reality 221
Electrons in a jellium, charge screening, electron
gas and FermiDirac distribution
Electrons in a crystal: quasiparticles*
Electrons in a crystal: Bloch functions; quasimomentum;
Brillouin zones; energy bands; and Fermi surface*
Quasiparticles and Greens functions
Perturbation theory
Feynman diagrams and virtual particles
Summing Feynman diagrams and charge screening
Dressing the quasiparticles
9 the path of extremal action 254
Variational approach to classical mechanics
Variational approach to quantum mechanics
Feynmans path integrals
Charged quantum particles in an electromagnetic field, path
integrals, the AharonovBohm effect and the Berry phase*
10 Order! Order! 281
Classical magnets, second order phase transitions
and the order parameter
Landaus theory of second order phase transitions and
spontaneous symmetry breaking
Superconductors, superconducting phase transition
and the superconducting order parameter
BCS theory, BoseEinstein condensate, Cooper pairs
and the superconducting energy gap
Josephson effect*
11 curiouser and curiouser 311
Einstein-Podolsky-Rosen paradox
Entanglement and faster-than-light communications
Spooks acting at a distance, no-cloning
and quantum tomography
Schrdingers cat and Wigners friend
Bells inequality*
Contents vii
12 schrdingers elephants and quantum
slide rules 339
The need and promise of quantum computers
Digital computers, circuits and gates
Quantum gates
Quantum parallelism and the Deutsch algorithm*
The Shor algorithm, code-breaking and the promise of an
exponential speed-up
Schrdingers elephants, quantum error correction and
DiVincenzo criteria
Quantum slide rules: adiabatic quantum computing and
quantum optimizers
13 history and philosophy 374
Sturm und Drang: the old quantum theory
When all roads led to Copenhagen: creation of quantum
mechanics
What the FAPP?
It was a warm summer evening in ancient Greece
The proof of the pudding
The final spell
Fact-check answers 393
taking it further 395
Figure credits 397
index 398
viii
How to use this book
?
The fact-check questions at the end of each chapter
are designed to help you ensure you have taken in the
most important concepts from the chapter. If you find
you are consistently getting several answers wrong,
it may be worth trying to read more slowly, or taking
notes as you go.
x
Introduction
A fair warning
The first thing a student of magic learns is that there are books
about magic and books of magic. And the second thing he
learns is that a perfectly respectable example of the former may
be had for two or three guineas at a good bookseller, and that
the value of the latter is above rubies.
Susanna Clarke, Jonathan Strange & Mr Norrell
Introduction xi
mechanics, and what this research tells us about the world
we live in. Besides, much of our technology uses quantum
mechanical effects routinely (like lasers and semiconductor
microchips, which were produced by the first quantum
revolution back in the 20th century), and if the current trends
hold, we will be soon enough playing with the technological
results of the second quantum revolution, using even more
subtle and bizarre quantum effects. It is therefore wise to be
prepared. As a bonus, you will learn about quite a lot of things
that are not taught in introductory quantum mechanics courses
to physics undergraduates.
This book contains quite a number of equations (though much
less than a book of quantum mechanics would), so brushing
up your school maths would be a good idea. I have tried to
introduce all the new mathematical things, and some things
you may already know. Still, this is just a book about quantum
mechanics. So why equations?
Quantum mechanics describes our world with an unbelievable
precision. But the scale of phenomena, where this description
primarily applies, is so far removed from our everyday experience
that our human intuition and language, however rich, are totally
inadequate for the task of expressing quantum mechanics directly.
In such straits it was common to use poetry with its allusions,
allegories and similes. This is indeed the only way one can write
about quantum mechanics using only words even if the poetry
lacks rhymes and rhythm. This is a nice and often inspiring way
of doing things, but poetry is too imprecise, too emotional and
too individual.
It is therefore desirable to use the language in which quantum
mechanics can be expressed: mathematics, a language more
remote from our common speech than any Elvish dialect.
Fortunately, this is also the language in which all physics is
most naturally expressed, and it applies to such areas like
mechanics where we do have an inborn and daily trained
intuition. I did therefore try to relate the one to the other you
will be the judge of how well this approach succeeded.1
xii
Familiar physics in
1
strange spaces
To an unaccustomed ear, the very language of quantum
mechanics may seem intimidatingly impenetrable. It is full of
commutators, operators, state vectors, Hilbert spaces and other
mathematical horrors. At any rate quantum physics seems
totally different from the clarity and simplicity of Newtonian
mechanics, where footballs, cars, satellites, stars, planets,
barstools, bees, birds and butterflies move along well-defined
trajectories, accelerate or decelerate when acted upon by forces,
and always have a definite position and velocity. All this can
be described by the simple mathematical laws discovered by
Sir Isaac Newton and taught at school or just intuitively
comprehended based on our everyday experience. Quantum
mechanics, on the contrary, is best left to somebody else.
1
Philosophy is written in that great book which ever lies before
our eyes I mean the universe but we cannot understand it
if we do not first learn the language and grasp the symbols,
in which it is written. This book is written in the mathematical
language, and the symbols are triangles, circles and other
geometrical figures, without whose help it is impossible to
comprehend a single word of it; without which one
wanders in vain through a dark labyrinth.
Galileo Galilei (15641642)
2
z
z
O y
O y
x
x
As you can see from the diagram, the coordinates (x,y,z) of the
same point A in the new system are given by
x' = x X, y' = y Y, z' = z Z.
Any coordinate transformation can be written as:
(x,y,z) = (x,y,z).
Here is shorthand for an explicitly known set of operations one
must perform with the set of coordinates in the original system, in
order to obtain the coordinates of the same point in the new one.
Mathematically, this means that is an operator the first scary
word is demystified.
Spotlight: at a tangent
According to classical mechanics, to fly at a tangent actually
means to keep moving in the same direction as before, and not to
start on something completely different!
4
Spotlight: Looking for treasure
If you wish, you can treat a radius vector as a set of instructions
on a treasure map. For example, from the old oak tree, go x feet
east, then y feet north, then dig a hole (-z) feet deep (or climb z
feet up a cliff).
N
W E
S
Dig here 7
150
50
O y
x
Figure 1.1 cartesian coordinates, radius vector and a trajectory.
6
Key idea: Derivative
The derivative
df is the mathematically rigorous expression for
dx
the rate of change of a function f (x) as the variable x changes.
For example, the incline of a slope is given by the derivative of
height with respect to the horizontal displacement:
dh(x )
tan =
dx
h
x
The Leibnizs notation for the derivative reflects the fact that it is a
limit of the ratio of a vanishingly small change f, to a vanishingly
small change x.
r
r + r
O y
x
Figure 1.2 Displacement
8
The velocity v(t) is always at a tangent to the trajectory r(t), as
you can see from Figure 1.3.
v
vav
r r + r
O
Figure 1.3 average and instantaneous velocity
dvy dvz d 2 x d 2 y d 2 z
( a , a , a ,) = dvdt
x y z
x
,
dt
, = , ,
dt dt 2 dt 2 dt 2
.
dt dt
r depends on time. It is this dependence (that is, where our
material point a football, a planet, a bullet will be at a given
moment) that we want to find from Newtons second law.
10
moment of time.3 Since we do not require any higher-order
derivatives of the radius vector in order to make this prediction,
we do not bother calculating them. Positions, velocities and
accelerations are everything we need.
Phase space
Now here surfaces a little problem with our picture of a
trajectory in real space, Figure 1.1. Unless we somehow put time
stamps along the trajectory, we cannot tell how fast or even
in what direction! a particle moved at any point, and will be,
of course, unable to tell how it will behave after flying off at a
tangent. Such time stamps are messy and, frankly, no big help
(Figure 1.4). Using them would be like a forensics unit trying to
figure out what happened during a road accident based on the
skid tracks on the road alone.
z z
32 27
30
20
5 25
10 30
25 29
0
O y O y
10
0 12
x x
Figure 1.4 time stamps along a trajectory
(a)
(b)
x
(c)
12
with a simple example of a phase space so simple, that it can
be drawn on a piece of paper.
Let us consider a particle, which can only move along one axis
like a bob of a spring pendulum, a fireman sliding down a pole or
a train carriage on a straight section of the tracks. Then the only
coordinate that can change is, say, x, and the rest do not matter
(it is customary to denote the vertical coordinate by z, but this
is more like a guideline). Now we can plot the corresponding
momentum, px, along an axis orthogonal to Ox, and obtain a
quite manageable, two-dimensional phase space (Figure 1.5). The
negative values of px mean that the particle moves in the negative
direction of the Ox axis. The trajectory of a particle in this space
is easy to read. For example, particle (a) started moving to the
right along Ox with a finite speed and slows down, (b) moves
right with constant velocity, and (c) was initially at rest and then
started accelerating to the left. On the other hand, you can easily
see that unlike the trajectories in real space inverting the
direction along these trajectories is impossible: a particle cannot
move left, if its velocity points right, etc. This indicates a special
structure of the phase space, which is indeed very important and
useful, but we do not have to go into this.
sin x cos x
1
2
x
x
0
Figure 1.6 a spring pendulum
14
Here the force (e.g. the reaction of the spring) is proportional to
the deviation from it (Hookes law): F(x) = kx.
The number k is the rigidity of the spring, x is the coordinate
of the bob, and we chose the origin at the point of equilibrium.
The minus sign simply indicates that the elastic force tends to
return the bob back to the equilibrium position.
Newtons second law for the bob is then written as:
d2x
m = kx,
dt 2
d2x
or = 02 x.
dt 2
and plot them for different values of the initial position and
momentum, x0 and px0 mv0 (Figure 1.7). It is enough to trace
the trajectory for time 0 t T0 2 . After that the motion
0
will repeat itself, as it should: this is an oscillator, so it oscillates
with the period T0. (The frequency 0 is the so-called cyclic
frequency and is measured in inverse seconds. More familiar to
non-physicists is the linear frequency 0 1/T0 = 0/2, which
tells us how many oscillations per second the system does, and
is measured in Hertz (Hz): 1Hz is one oscillation per second).
px
a a
x
16
One of the great achievements of Descartes was the development
of analytic geometry. In other words, he discovered how to
translate from the language of curves and shapes (geometry) to
the language of equations (algebra) and back again. In particular,
it turned out that ellipse, one of special curves known already to
ancient Greeks, is described by an equation
x 2 + y 2 = 1.
a2 b 2
First, we see that all the trajectories are closed curves; second,
that the motion along them is periodic with the same period
(this is why the oscillator is called harmonic: if it were a string,
it would produce the same tune no matter how or where it
was plucked); and, third, that these trajectories fill the phase
space that is, there is one and only one (as the mathematicians
like to say) trajectory passing through any given point. In other
words, for any combination of initial conditions (x0,px0 ) there is
a solution which is clear from its explicit expression and this
solution is completely determined by these initial conditions.
What is the shape of the curves in Figure 1.7? Recalling the basic
properties of trigonometric functions, we see that
x2 px2
+ = 1.
p 2 p 2
x0 x0
x02 1 + (m 0 x0 )2 1 +
m 0 x0 m 0 x0
0 0 m 0 x0
Spotlight: integration
Newton and Leibniz discovered integration independently, but we
are all using Leibnizs notations, as with the derivative.
x x
Let us consider a function f(x) and find the area enclosed by its
graph and the axis Ox, between the points x = a and x = b. To do so,
we can slice the interval [a,b] in a large number N of small pieces
b a
of length x = and approximate the area by the sum of the
N
areas of narrow rectangles of width x and height f(xi):
N
S f (x ) x
i =1
i
18
U(x) = kx2/2=m02x2/2, and the kinetic energy of the bob is
K=mv2/2=px2/2m, we can easily check that the equation for
the elliptic phase trajectory is nothing else but the mechanical
energy conservation law,
m 02 x2 px2
H ( x , px ) = U ( x ) + K ( px ) = + = constant.
2 2m
p 2
p x20
xo
S = x m pm = m 0 x02 1 + = m 0 x0 +
2 2
m 0 x0 0 m 0
2
= H ( x0 , p x 0 ) T0 H ( x0 , p x 0 ) .
0
S= p (x)dx.
x
x
xmin xmax
20
To put it simply, the Hamiltonian is the energy of a mechanical
system expressed through its coordinates and momenta.
How will the harmonic oscillators equation of motion look
after we replace the velocity with the momentum? Since
dx
px = mvx = m ,
dt
then
d2x d dx d px
= = = 02 x.
dt 2 dt dt dt m
These are the famous Hamiltons equations, which hold for any
mechanical system. This is a set of two first order differential
equations for two conjugate variables a coordinate and a
momentum.
y
x
22
We can take the set of coordinates and momenta and consider
them as the coordinates of a single point in a 6N-dimensional
phase space or, if you wish, as components of a radius vector
in this space (Figure 1.9):
R (x1,px1,y1,py1,z1,pz1, ..., xN,pxN,yN,pyN,zN,pzN).
One can call it the state vector, since it completely determines
both the current state of the system and its future evolution,
governed by the Hamiltons equations.
py1
pz1
pzN R px1
zN
x2
z1
y1
x1
Figure 1.9 6N-dimensional phase space of a system of N particles
24
The gist of it is that Lagrange developed a method of describing
an arbitrary mechanical system by a set of generalized
coordinates, qi and generalized velocities, qi dqi /dt. These
coordinates can be any convenient parameters that describe
our mechanical system: rotation angles; elongations; distances;
or even conventional positions in real space. The index
i = 1, 2, , M labels these coordinates; M is the number of
the degrees of freedom our system possesses. Please note that,
by tradition and for brevity, for the generalized velocities the
Newtonian notation for a time derivative is used.
dpi H
= ;
dt q i
i = 1,2, ... 2M
In many cases these equations are easier to solve than the
Lagrange equations. What is more important to us is that they
form a vital link between quantum and classical mechanics.
Hamiltons equations describe the evolution with time of the
state vector in a 2M-dimensional phase space of our system.
One cannot easily imagine this, but some geometrical intuition
based on our lower-dimensional experience can be actually
developed with practice. For example, you can draw the phase
trajectories in each of the M planes (qi,pi) and consider them in
the same way one does different projections of a complex piece
of machinery in a technical drawing.
We will try to show how it works on another example the last
in this chapter of those notions from classical physics, which
turn out to be not that much different from their quantum
counterparts.
26
Statistical mechanics, statistical
ensemble and probability
distribution function
One of the simplest and most useful model systems considered
by classical physics is the monoatomic ideal gas a collection
of a very large number N of identical material points, which do
not interact with each other. This model helped establish the
fundamentals of statistical mechanics the aspect of physics
that explains, in particular, why water freezes and how an
electric heater works. We already know that the state of this
system can be completely described by a state vector in a 6N-
dimensional phase space (or, which is the same, a single point in
this space the end of the state vector; see Figure 1.9).
As we know, the state vector determines all the positions
(in real space) and all the momenta of all N gas particles.
A single cubic centimetre of gas at normal pressure and zero
degrees Celsius contains about 2.7 1019 molecules. Therefore
the state vector contains a lot of unnecessary and unusable
information. It is unusable, because even if we had all this
information, we could never solve 6 2.7 1019 Hamiltons
equations per cubic centimetre of gas. Unnecessary, because
we do not care where precisely each gas particle is placed and
what momentum it has at a given moment of time. We are
interested in bulk properties, like temperature, pressure, local
flow velocity and so on at a given position in a given moment
of time.
The central point of statistical mechanics makes it so that a
single set of such bulk properties (P(r, t), T(r, t), v(r, t), ...) a
so-called macroscopic state of our system corresponds to
a huge number of slightly different microscopic states (each
described by a state vector in the 6N-dimensional phase space).
These are found in the vicinity of certain point R = (q, p) in this
space (Figure 1.11), which has volume (q, p).
q4
q3
q2
q1
Figure 1.11 a statistical ensemble in phase space
28
This function is called the N-particle probability distribution
(or simply distribution) function. It gives the probability to find
the system in a given microscopic state (because it still depends
on all the momenta and all the positions of all the particles).
Obviously, this is not much of an improvement, since this is as
detailed a description, as we had initially, and the equations
of motion for this function are as many, as complex and as
unsolvable as the Hamiltons equations for all N particles. But
once we have introduced the distribution function, some drastic
simplifications become possible. To begin with we can ask what
the probability is for one gas particle to have a given position
and momentum at a given time. Let us denote it as
f1(q, p, t),
where now (q, p) are the position and momentum of a single
particle in a phase space with a miserly six dimensions.
Since all gas particles are the same, the answer will do for any
and for all of them. Moreover, properties such as temperature
or pressure, which determine the macroscopic state of gas, can
be calculated using only f1(q, p, t). Finally, the approximate
equations, which determine the behaviour of f1(q, p, t), can be
written down and solved easily enough.
Sometimes we would need to know the probability that one
particle has the position and momentum (q1, p1), and other
(q2, p2), simultaneously. The approximate equations for the
corresponding two-particle distribution function, f2(q, p, t),
can be also readily written down and with somewhat
more trouble solved. Such equations can be written for all
distribution functions, up to the N-particle one from which
we started, but in practice physicists very rarely need any
distribution functions beyond f1(q, p, t) and f2(q, p, t).
The idea that sometimes we can and need to know something
only on average is a very powerful idea.4 It allowed the great
progress of classical statistical mechanics. We will see how it is
being used in quantum mechanics.
dfN (q , p , t) fN fN dq j fN dp j
dt
=
t
+ q
j j dt
+
p j dt
= 0,
fN (q , p , t) fN dq j fN dpj
t
= q
j j dt
+
pj dt
.
? Fact-check
1 Velocity is
a d2r/d2t
b dr/dt
c dt/dr
d d2r/dt2
2 Acceleration is
a d2r/d2t
b dr/dt
c dt/dr
d d2r/dt2
30
3 In order to predict the behaviour of a material point you need
to know its initial
a position
b position and velocity
c velocity and acceleration
d position and momentum
(b)
x
(c)
(a)
Dig deeper
Further reading I. Some serious reading on classical mechanics
and related mathematics.
R. Feynman, R. Leighton and m. sands, Feynman Lectures on
Physics, Vols i, ii and iii. Basic Books, 2010 (there are earlier
editions). see in particular Vol. i (read there about mechanics
and some mathematics, chapters 125).
P. hamill, A Student's Guide to Lagrangians and Hamiltonians.
cambridge university Press, 2013.
32
F. Reif, Statistical Physics, Berkeley Physics course, Vol. 5.
mcGraw-hill Book company, 1967.
L. susskind and G. hrabovsky, Theoretical Minimum: What You
Need to Know to Start Doing Physics. Basic Books, 2013. here
is a really welcome book: classical mechanics, including its
more arcane sides (like phase space and hamilton equations),
presented in an accessible form! susskind is a prominent string
theorist, which makes it even more interesting.
Ya. B. Zeldovich and i. m. Yaglom, Higher Math for Beginning
Physicists and Engineers. Prentice hall, 1988. introduces
mathematical tools together with the physical problems they
were developed to solve, and teaches learners how to use these
tools in practice.
Further reading II. About the ideas and scientists.
V.i. arnold, Huygens & Barrow, Newton & Hooke. Birkhauser,
1990. the same story, told by a great mathematician from his
unique point of view, and much more.
J. Bardi, The Calculus Wars. high stakes Publishing, 2007. how
newton and Leibniz discovered calculus.
L. Jardine, The Curious Life of Robert Hooke: The Man who
Measured London. harper Perennial, 2004. a very good biography.
t. Levenson, Newton and the Counterfeiter. Faber & Faber, 2010.
a fascinating story of how newton managed the Royal mint,
tracked and caught a gang of counterfeiters, and had their chief
executed.
34
Let us start from the beginning. In 1900, Professor Max Planck
of the University of Berlin introduced the now famous Planck
constant h = 6.6261034 J s. For very compelling reasons,
which we will not outline here, Max Planck made the following
drastic prediction concerning harmonic oscillators, which in
modern terms went as follows:
A harmonic oscillator with frequency 0 can change its energy
only by an amount h0 (the so-called energy quantum, from the
Latin quantus meaning how much).
In other words, E = ho.
Initially Planck referred to what we now call energy quantum
of an oscillator as one of finite equal parts of vibrational energy.
The term quantum was coined later. One of its first users, along
with Planck, was Philipp Lenard, a first-rate experimentalist,
whose work greatly contributed to the discovery of quantum
mechanics, and who was also regrettably a first-rate example
of how a brilliant physicist can go terribly wrong outside his
field of expertise. But, no matter what terms were used, the
notion of deriving discrete portions of energy from an oscillator
was revolutionary.
h
= = 1.054571726(47) 1034 J s.
2
36
where H(x0,px0) = E is the oscillator energy. What consequences
will follow from the energy quantization condition, E = hv0, or,
as it is usually written in modern books and articles, E = 0?
There seems to be little to worry about when dealing with
the events in our everyday life. Planck constant is exceedingly
small. Take as an example of an oscillator a childs swing,
with a typical frequency of about one swing per second,
which corresponds to the linear frequency of 1 Hz. Then the
corresponding energy quantum is h1 Hz = 6.6261034 J. This
is a very small amount of energy indeed. An average ant (which
weighs about 0.5 mg) falling from a height of 1 mm will acquire
energy about 1025 times greater than that. For comparison, the
Sun radiates approximately 1025 times more energy per second
than a standard 40 W light bulb (and remember here we count
all of the Suns energy production, not just the tiny amount that
reaches our insignificant planet). So, at least from the point of
view of everyday physical activities, quantization of energy does
not seem to lead to any perceptible effects.
You recall that this area is called action and can be written as
S = px ( x ) dx. Therefore, the Planck constant h is called the
quantum of action.
The relation
S=
p ( x) dx = (n + ) h; n = 0,1,2. ;0 < 1,
x
38
is known as the BohrSommerfeld quantization condition, and
it holds not just for a harmonic oscillator; it holds true in every
case, when such an integral can be calculated, e.g. for a non-
1 1
E = n +
o = n + h o .
2 2
40
Spotlight: Paul ehrenfest
Paul Ehrenfest (18801933) made deep and lasting contributions to
theoretical physics. One of his (quite unjustified) worries was that
he was not as good a physicist as his best friends Albert Einstein
and Niels Bohr.
Paul Ehrenfest noted that the ratio E/vo does not change if
the frequency of the oscillator is altered infinitesimally slowly
or, using a shorter term, adiabatically. Then, even though the
energy of the system changes, the number n of its quanta stays
the same!
1 E E
n+ = = = const.
2 o h o
2 Well, actually you will because not every system has nice closed phase
trajectories. Fortunately those were not important enough and did not have
to be considered when the very basics of quantum mechanics were being
discovered. Modern quantum mechanics has developed tools for dealing
with them.
x2
z1
y1
x1
Figure 2.4 elementary hypervolume of a 6N-dimensional phase space
3N
V = (x1 px1) (y1 py1) (z1 pz1) ... (xNn pxN) (yN pyN) (zN pzN) = h
42
Conjugate variables are those related by the Hamilton equations
and their dimensionalities are as such to make the dimensionality
of their product the same as that of the Planck constant, Js.
There are lots of pairs of conjugate variables in physics, and we
will meet some of them later but the position and momentum
are the most important ones for our current purpose.
On the surface of it, the uncertainty principle sounds pretty
simple; for some reason we cannot know exactly the position
and the momentum (i.e. velocity) of a particle. So what?
Suppose the phase space is neatly sliced in squares of area h (or
hypercubes of volume h3N). If that were the case, the Universe
would be something like a chessboard with very small squares,
and with game pieces elementary particles making very
small jumps between them (which would, probably, require
slicing time in tiny bits t as well). Of course then we would
only know the position and momentum of each piece up to the
size of the square it occupies and would not care, since a rook
can take, say, a bishop no matter where in its square it stands
(Figure 2.5).
pm H (xn +1 , pm ) H (xn , pm )
= .
t x
44
This is well and good for a classical oscillator. But a quantum
oscillator has the minimum energy (so-called zero-point energy)
E0 = 1 0 = 1 hv0 and the minimum action S0 = 1 h.
2 2 2
1
This is because of the value of = 2 in the BohrSommerfeld
quantization condition for the oscillator, found by Einstein and
Stern. The phase trajectory for a quantum oscillator at rest is
therefore not a point at origin, but an ellipse with area h/2 and
this is the smallest area that can be enclosed by a phase trajectory.
What is then the meaning of a point in the phase space?
z = x + iy
|z|
z* = x iy
z = x + iy
has the coordinates x along the real axis and y along the imaginary
axis.
z*= x iy.
The distance from the origin to the point z is denoted by |z| and, of
course, equals
| z | = {x 2 + y 2 }.
| z | 2 = (x + iy )(x iy ) = z z *.
46
mechanically, no matter how big their action is compared
to h.3 But special measures must be taken to observe such a
behaviour, and in the vast majority of cases, large systems (i.e.
those with S >> h) indeed do behave simply as if their phase
trajectories were drawn by a pen of thickness h per a degree of
freedom and this does not make any difference.
4 The terms real and imaginary come from earlier times when
mathematicians were less acquainted with strange and unconventional
mathematical objects and were not sure how to deal with them. To think
of it, they considered all numbers, which cannot be expressed as normal
fractions, irrational, and some of them even transcendent! By the time
these objects became well understood, it was pointless to rename them.
48
only takes some practice. For example, (1 + 2i)(3 4i) = 3 +
6i 4i 8i2 = 3 + 8 + (6 4)i = 11 + 2i. Instead, they allow
simplification of many mathematical operations.
One of the most useful simplifications one can get from complex
numbers is the Eulers formula for the exponent of a complex
number,
ez = ex+iy = ex (cos y + i sin y).
0 L
50
n
2 n 2 n
f (x) =
n=o
An cos
L
x+ B sin
n =1
n
L
x
L
2 n 2 m
0
dx sin
L
x cos
L
x = 0 for any m, n.
A= A e .
j =1
j j
Here the unit vectors ej are the basis vectors in our conventional
space they are simply vectors of unit length along the directions
Ox, Oy and Oz, and they are all at right angles to each other, that
is, their scalar (or dot) product is zero:
ej ek = 0 if j k.
Now, we know how to calculate a scalar product of two vectors
in conventional space. But the mathematicians found that if you
consider functions as vectors, you can consider their integral as
a scalar product. To be more specific, for any two functions, say
(x) and g(x) any two shapes of the string we can define the
scalar product (which is actually called a scalar product why
invent new terms, when the existing ones will do?) as
L
g|f = 0
g* (x)f (x)dx.
.
L
Obviously, C = 1 / f ( x) d x .
0
2
52
cos (2x/L)
cos (4x/L)
sin (8x/L)
sin (6x/L)
sin (4x/L)
sin (2x/L)
Figure 2.7 a hilbert space of functions, which can be expanded in Fourier
series
6 Yes, there exist infinities, which are larger than other infinities. This is a
fascinating subject, to which Hilbert contributed as well but unfortunately,
not for this book.
54
Yes, this is just giving a fancy name to a product, but this is the
simplest case. We can introduce more interesting operators, like
the differentiation operator, which produces the derivative of a
function:
(x) = d f (x)
Df
dx
x , p = x , D
= = i.
x i i
7 There will be yet another twist but we will deal with that later on.
8 This is true only for certain operators and this is how we figure out
whether this or that operator can correspond to a physical quantity.
56
Using the integral as the scalar product in our Hilbert space, we
see that their expectation values are given by:
L
x = x = 0
x( x)2 dx and
L
d (x)
p x = p x =
0
* (x)
i dx
dx.
|a = a |a
|a
58
|a
of
nt
me |
su
re
a
Me |3
|2
|1
Figure 2.8 the measurement of an observable A projects the state vector
on one of the eigenstates of this observable
We can make sense of this by considering the measurement
as the calculation of a component of some vector along a
certain direction. In conventional space, if we want to find the
component of a vector R along, say, the axis Ox, we project
the vector on this axis. This can be done, e.g. by calculating the
scalar product between R and ex, the unit vector along the axis.
In a similar vein, we can think of the result of measurement
of some observable in state as finding the projection of
on one of the eigenstates of the operator (Figure 2.8).
The square modulus of this projection gives the probability of
observing a particular eigenvalue, while the projection of state
= , which arose because of the action of the operator
on ), on the initial state gives the average result of many
repeated measurements, i.e. the expectation value .
So, if a system is in an eigenstate (e.g. a), then the measurement
will give the same result, , with probability P(a) = aa2 = 1,
that is, always. The average, of course, will be also , and it
coincides with the expectation value, = aa = aa = .
If this is not the case, then repeated measurements will produce
at random all eigenvalues of but with such probabilities that
on average we will obtain the expectation value, .
We have met with a similar situation in Chapter 1, when the
ideal gas was described by a probability distribution function
and the positions and velocities of particles could only be
known on average. There lies the reason for such vagueness:
there were too many particles. As we have stated there, given all
Spotlight: epicurus
Epicurus (341270 BCE) was a Greek atheist philosopher. He
further developed the atomistic philosophy of Democritus and
suggested that from time to time atoms in motion undergo
spontaneous declinations what can be considered as the
predecessors of random quantum leaps!
60
We have stated that Nature is intrinsically random. We have
related measurement of physical quantities with projections
of the state vector on some other vectors (the eigenstates of a
given observable) in Hilbert space. How will this help us with
quantization and uncertainty relation?
Earlier on in this chapter we have stated that for a particle in
one dimension, position and momentum are represented by
operators x and p x = d , and that the commutator of these
i dx
operators is
x , p x = i.
0 x L
Figure 2.9 the wave function of a particle in a one-dimensional box (in the
lowest-energy state)
state vector. Let us ask two questions: (i) What will be the
expectation value of position and momentum of such a particle,
i.e. the averaged measurements of these quantities? (ii) What is
the uncertainty of these measurements?
These are both easy to calculate:
L
L
x =
0
x | ( x) |2 dx =
2
;
2
L
L L 2 6 L
x = 0
x | ( x) | dx =
2
2
2 3
0.36 ;
2
L
d
p x =
0
* ( x )
i dx
( x) dx = 0;
2
L
d
2
px = 0
* ( x )
i dx2
( x) =
L
;
62
0.36
x px 0.57 .
2
? Fact-check
7 Hilbert space is
a an abstract space, where the state vector of a quantum
system lives
b an isolated space where a quantum system must be
placed in order to observe its properties
c the usual space as it is called by physicists
d the part of the Universe where quantum mechanics applies
64
10 Which of the below expressions is correct?
a x px
2
h
b x px
2
c x pz
2
d E
2
Dig deeper
Further reading I. Basics of quantum mechanics.
R. Feynman, R. Leighton and m. sands, Feynman Lectures on
Physics, Vols i, ii and iii. Basic Books, 2010 (there are earlier
editions). see Vol. i (chapters 37, 41 and 42) and Vol. iii (start
with chapters 1, 3, 5 and 8).
e. h. Wichmann, Quantum Physics (Berkeley Physics course,
Vol. 4). mcGraw-hill college, 1971. this is a university textbook,
but it is probably the most accessible of the lot.
Further reading II. About the ideas and scientists.
J. Baggott, The Quantum Story: A History in 40 Moments. Oxford
university Press, 2011. a modern history. Well worth reading.
G. Gamow, Thirty Years that Shook Physics: The Story of Quantum
Theory. Dover Publications, 1985. a history written by an
eyewitness and one of the leading actors in the field.
66
When introducing the Hilbert space, state vector and operators,
we claimed that this will allow us to keep some essential
features of classical mechanics such as the description of
a quantum systems evolution in terms of a vector in some
exotic, but comprehensible, space so that once we know the
state of the system, we can predict it at all future times.
The price we paid is pretty high. For a single particle in one
dimension instead of a two-dimensional phase space we now
have to deal with an infinite-dimensional Hilbert space! Therefore
let us check that we at least obtained what we have bargained for.
A part of the price was the split between the state of the
system, now represented by the vector , and the observables;
in this case the position and momentum operators, x and px.
This means that when considering the evolution of a quantum
system with time, we could either look at the change in ,
or in x and px. Both approaches are equally valid and both are
being used, depending on what is more convenient. The first is
due to Schrdinger, the second due to Heisenberg, but it was
Dirac who realized that they are actually the same and that they
are closely related to classical mechanics.
The similarity is easier to see if we start with the Heisenberg
representation.
68
derivative with respect to one independent variable, we keep all
others constant. Then indeed
dx x H H dpx H px H
= 0= ; =0 = .
dt x px px dt x px x
dA
i = [ A , H ].
dt
One of the great insights of Paul Dirac was that these Heisenberg
equations of motion are directly analogous to classical equations
of motion written using Poisson brackets. One only needs to
replace the classical functions and variables with quantum
70
operators, and instead of a classical Poisson bracket put in the
commutator divided by :2
1
{ f , g } [ f , g ],
i
And voil! You have a quantum equation of motion. This is a
much more consistent scheme of quantization than the one of
BohrSommerfeld.
Once we know the solutions to the Heisenberg equations, the
expectation value for an observable (t) at any moment of time
is immediately found:
A ( t ) = A (t) = | A (t) | .
Here is the state vector of our system. Note that it does not
change in time. This is because in the Heisenberg representation we
consider the time dependence of the operators themselves. The state
vector is static. We can do this due to the aforementioned split
between the state vector and the operators of physical quantities.
A B B A
Using the same rule, we can multiply a vector column by a square
matrix from the left, producing another vector column.
You can check that one can multiply a vector row by a square matrix
from the right and obtain another vector row.
In quantum mechanics vector columns are denoted by kets, , and
vector rows by bras, . Then an operator, represented by a square
matrix, can act on a ket from the left, and on a bra from the right:
) = ,
= .
| j |3
|1 |2
Figure 3.1 action of an operator in a hilbert space
An operator acts on a vector in a Hilbert space: = .
We can expand both state vectors, and :
| = j |j ;| ' = '
j | j .
Here j are basis states of the Hilbert space (like sine and cosine
functions in Chapter 2), which play the same role as unit vectors
72
ex, ey, ez in our conventional three-dimensional space. Depending
on what system we are dealing with, there can be finite or infinite
number of such basis states, but this will make no difference.
Recall that the scalar product of two different basis
vectors is zero: jk = 0, j k. Of course, the scalar product of
a basis vector with itself is the square of its length, that is, one:
j j = j2 = 1.
The notation we use for the scalar products of vectors in a
Hilbert space, ba, is another contribution due to Dirac. It
can be interpreted as a product of a vector a (ket-vector, or
simply ket) and a conjugate vector b (bra-vector, or bra).
Together they form a bracket, so the terms are self-explanatory.
A bra b is conjugate to the ket b : if b is described by
some complex coefficients, then b is described by complex
conjugates of these coefficients. Operators act on a ket vector
from the left: a, but on the bra vectors from the right: a.
A state vector is a ket: | = j | j ; the corresponding bra is
| = *j j | . They contain the same information about the
system.
The coefficients j , j (components of the state vector before
and after it was acted upon by the operator ) are just some
complex numbers. Therefore all we need to know about the
operator is how it changes these into the others.
In conventional space the action of operators is well known.
(They are not necessarily called operators, but this is a question
of terminology.) For example, a rotation of a vector in xy-plane
by the angle is given by an operator (). In a Hilbert space of
quantum mechanics we follow the same approach. An operator
can be written as a matrix, a square table of numbers:
A A12
11
A = A21
( )
A22 = Aij ,
or, keeping just the expression for the state vector components,
'j = Ajk k .
k
The rule for multiplying two matrices, e.g. and B is the same:
the (jk)th matrix element of the product = B is given by the
same row-by-column rule:
[ A , H ]ij = (A H
k
ik kj H ik Akj ).
74
This is not a single equation, but a set of equations for all pairs
(i, j). This may be even an infinite set, but such equations can
be solved either exactly, or more often approximately, or even
more often, numerically. The point is, such a solution (a set
of numbers, the time-dependent matrix elements A(t)) can be
obtained.
Now we can calculate the expectation value of the observable ,
A (t) = | A (t) | .
| ' = i | A (t)
i
*
i
j k
jk k | j =
A (t)i | k.
i j k
*
i k jk
(t ) = A (t).
i j
*
i j ji
d A d d d | .
= | A | + | A | + | dt
dt dt dt
On the other hand, if the operators are fixed and the state vector
changes in time, we get
76
d A d d
= | A | + | A | .
dt dt dt
d 1 d 1
The two equations agree, if | = H | and | = | H .
dt i dt i
These two relations are actually equivalent (they are conjugate).
Therefore we have derived the famous Schrdinger equation:
d
i | = H | .
dt
4 Dirac showed that the difference between the Schrdinger and Heisenberg
representations is like the one between different frames of references in
classical mechanics, and the choice of the one or the other is purely a question
of convenience. He also invented yet another representation (the interaction
representation), but unless you want to become a professional physicist you do
not need to bother with it.
=
( k)2 .
2m
What does this solution look like? Using Eulers formula, we see
that (assuming C is a real constant)
(x, t) = C {cos(kx t) + i sin(kx t)}.
78
Both the real and the imaginary parts of (x,t) are waves, which
have frequency and wave vector k and propagate with the
speed s = (Figure 3.2). Naturally, is called a wave function.
k
The wavelength (the distance between the consecutive maxima
or minima) is found from the condition cos (k(x + )) = cos kx
2 6
and equals = .
k
In an oscillator, the combination is equal to its energy.
This holds true here as well (after all, a free particle is just an
oscillator with a broken spring): the energy of a particle is
E = .
We can therefore write
iE
t
(x, t) = (x)e
.
2 2 2 2
i (x, y, z , t) = + + (x, y, z , t),
t 2m x2 y 2 z 2
Quantum propagation
One conceptual problem with the use of a wave function is how
a wave can describe the propagation of a particle in a definite
direction. Here again much can be inferred from classical
physics this time from optics.
80
The wave theory of light was well established in the 19th
century, though it hails back to Christiaan Huygens, a Dutch
contemporary and competitor of Newton. In the process of
its development, the question of how a wave can propagate as
a ray of light was successfully answered (with a contribution
from Hamilton). Roughly speaking, if the wavelength is small
enough, one can shape the waves into a beam in a definite
direction.
iEt ipx
d iEt ipx
ipx ipx
p = | p | = {C*e e
Ce e }dx = | C |2 p e
e
dx
i dx
= | C |2 p
1 dx.
82
2 2
This relation holds if, and only if, the energy E = n .
2m L
This is important; a particle in a box cannot have an arbitrary
energy. Its energy is quantized, taking one of the allowed values:
2
2 n
En = , n = 1,2, ...
2m L
(x, y)
y
y = 0, M, 2M, x = 0, L, 2L,
f(x, y, z) = f(x + nL, y + qM, z + pQ)
n, q, p = 1, 2,
y
z
84
Mathematically, the trick means that we should look for periodic
solutions to the Schrdinger equation, so that (x + L) = (x).
You can check directly that two such solutions are
1 i ( pn x Ent ) / 1 i ( pn x Ent ) / 2 n nh
+n = e and n = e , if only pn = L = L ,
L 2
L
pn
and, of course, En = .
2m
These states are eigenstates of the momentum operator with the
eigenvalues pn :
d 1 i (pn x En t ) 1 i (pn x En t )
p +n = e = pn e = pn +n ;
i dx L L
d 1 i ( pn x Ent ) 1 i ( pn x Ent )
p n = e = pn e = pn n
i dx L L
First, since the length of the segment L = dx , the momentum
quantization condition in a ring is nothing else but a version of
the BohrSommerfeld quantization condition p dq = nh .
Second, you may ask what happened to the uncertainty
principle? After all, the momentum of the particle in the states
n is determined, p = 0, while the uncertainty of the position
cannot be more than the length of the segment L. The answer
is rather subtle. Imposition of periodic boundary conditions
on a segment is not quite the same as constraining a particle
to a ring of the same circumference. In the former case we
essentially assume that the particle can travel indefinitely
far, but only in a periodic environment (e.g. a crystal), which
removes the problem.7
(
A 3 = C 2 1 A 1 + 2 A 2 + 1 A 2 + 2 A 1 )
=C 2
(A
1 )
+ A 2 + C 2
(
1 A 2 + 2 A 1 )
The first bracket does not cause any objections: this is the sum
of averages in the states 1 and 2 . For example, if is the
position operator, then our particle in state 3 on average will
be somewhere between its positions in the states 1 and 2.
Unfortunately, there is another bracket, which contains cross
terms those with the matrix elements of between the states 1
and 2. These terms may completely change the pattern. Again,
if is the position operator, the presence of these terms can lead to
a strange situation when the particle will certainly not appear on
average anywhere near its positions in the states 1 and 2.
86
Key idea: Wave superposition
Wave amplitude = f(x) + g(x)
+ =
Destructive interference
+ =
Constructive interference
+ =
88
We should therefore generalize our approach to include this
extra averaging over the details we cannot (or do not find
expedient to) control.
|n , Pn
| j
|3
|1 |2
Figure 3.3 a quantum ensemble: the system can be in any of a number of
quantum states, with certain probability
A = n Pn n A n .
n = C nj j ,
j
90
A = C
k j n
nk n P Cnj* j A k.
A =
k j
kj j A k.
A = tr( A ),
fN (q, p, t) f dq j fN dp j
t
= q N
j dt
+
p j dt
.
j
fN (q, p, t) fN H fN H
t
= q
j j pj
= { f N , H },
p j q j
i = [ , H ].
t
is called the Liouville-von Neumann, or simply von Neumann
equation. It looks almost like a Heisenberg equation, but the
opposite sign indicates that we are working in the Schrdinger
representation. Here the state vectors (and the density matrix)
evolve in time, but the operators stay put!
In the Heisenberg representation it is the other way around. The
density matrix is constant, the operators depend on time, but
the expression for the expectation value,
A = tr( A ),
will be the same.
92
The von Neumann equation would be exact if it included the
entire Universe. This is, of course, impossible and unnecessary.
After all, we are interested in a particular quantum system and
do not care much what happens to everything else. Therefore,
instead of the Hamiltonian and the density matrix of the
Universe, we modestly use in this equation only the density
matrix and the Hamiltonian of the system we are actually
interested in. Unfortunately since our system does interact with,
if not everything else, then at least with much of the systems
environment, this will not be enough. One has to include in the
equation some terms that will describe this interaction. As a
result, one obtains the master equation
i = [ , H ] + [ ].
t
? Fact-check
1 Heisenberg equations of quantum mechanics describe the
evolution of
a state vectors
b operators
c Poisson brackets
d density matrix
94
9 A mixed quantum state is a state of a quantum system, which
can be described by
a an ensemble of quantum state vectors
b a density matrix
c a statistical operator
d a single state vector
Dig deeper
Further reading. Quantum evolution, Heisenberg and
Schrdinger equations.
R. Feynman, R. Leighton and m. sands, Feynman Lectures on
Physics, Vols i, ii and iii. Basic Books, 2010 (there are earlier
editions). Vol. iii (chapters 7, 8, 16, 20).
e. h. Wichmann, Quantum Physics (Berkeley Physics course,
Volume 4), mcGraw-hill college, 1971. chapters 68.
96
We have done a lot of heavy lifting in order to explain that the
basic structures of quantum mechanics share a lot with those
of classical mechanics. It is not our fault that these similarities
lie pretty deep and can be seen only using some mathematical
tools. But now we can take a breath and apply these tools to
simpler quantum objects, which are simpler precisely because
they do not have classical analogues.
We have seen examples of quantum systems, which have Hilbert
spaces with infinitely many dimensions (that is, with infinitely
{ }
many basis vectors: j : 0, 1, 2, ). But a quantum
system can have as few as two dimensions and such a system is
much easier to investigate. (A one-dimensional system would be
even simpler, but as such a system can be only in one and the
same state all the time, there is nothing to investigate there.)
We will use the example of quantum bits, or qubits. In principle,
any quantum system with a two-dimensional Hilbert space can
be a qubit. In practice there are certain additional requirements,
which such a system must satisfy. They stem from the intended
use of qubits as the basic units of quantum computers, and their
physical realization as mostly quite macroscopic devices. But
here we will consider the behaviour of ideal qubits.
One important remark before starting: a qubit is actually the
simplest physical system one can consistently investigate from
the first principles, that is, based on the basic physical equations
without any simplifications. This seems to contradict our
intuition: the behaviour of a particle in space or a harmonic
oscillator is much easier to understand. But this ease is the result
of a long habit: we are macroscopic beings and are conditioned
by evolution to feel at home within classical physics.
What about classical bits the units of standard computers,
which can only be in two states, zero and one? Are not they
even simpler? Yes, they are, but they cannot be investigated
from the first principles. Classical objects are characterized
by a position and momentum, and these can take any value
whatsoever within some continuous range and change
continuously with time. In order to describe a classical bit
ez
r
ey
y
ex
98
Greenwich meridian is called the longitude and the radius vector
(the height) is measured from sea level (that is, from some
idealized surface of our planet).
1 Despite significant progress since the turn of the century, the very
possibility of quantum computing on a practically useful scale remains a
matter of serious controversy, which can only be resolved by experiment. We
will discuss this topic in Chapter 12.
100
N = 5.0507 1027 J/T, that is, even smaller. Nevertheless
this did not stop researchers from using spins as qubits.
One approach was based on using a large number of identical
molecules, each containing several atoms with nuclear spins
(Figure 4.1) e.g. 1H, 13C or 19F. These spins interact with
each other and could be controlled and measured using nuclear
magnetic resonance (NMR) techniques.2 The enormous number
of spins involved compensated the weakness of each spins
signal. Unfortunately, this approach to quantum computing
does not scale. As one increases the number of qubits (i.e. the
spins in a molecule), the signal from them grows relatively
weaker, and in order to measure it, the overall number of
molecules involved must grow exponentially faster. NMR
quantum computers are not scalable. For anything practically
useful one would have to use an NMR computer of the size of
the Sun not quite an option.
19
19F F
19
F 13
C
13
C 19F
19
F Fe
C2H2 CO
CO
Figure 4.1 nmR quantum computing uses as qubits quantum spins
of certain atoms in a molecule. even though a macroscopic number
of molecules are used, the effective number of qubits is the number
of working spins in a single molecule. the substance in this picture
(perfluorobutadienyl iron complex) contains seven qubits and was used in
2001 to perform a real quantum algorithm: factor the number 15 = 3 5.
This downside of the NMR approach to using all-natural qubits
is due to its bulk character, when all spins are controlled and
measured at once. In order to overcome it, it is necessary to
address each spin individually and in many cases it is simpler
V0 V1
|0 |1
Figure 4.2 a quantum dot-based qubit
102
The advantage of using superconductors is that, due to their
special properties, they are better protected from external noise.
For example, a loop made of a superconducting wire can carry
the electric current forever. A small enough superconducting
loop, interrupted by Josephson junctions4, is a superconducting
flux qubit (Figure 4.3b). In the state 0 the current flows
clockwise and in the state 1 counterclockwise. (The flux refers
to the magnetic flux, which these currents produce.)
While with charge qubits, the states 0 and 1 differ by
a microscopic quantity: the electric charge of one or two
electrons; the currents in flux qubits involve a macroscopic
number of electrons. Nevertheless these qubits behave as
quantum two-level systems, which is a very remarkable fact.
There are many more different types of qubits, which were
successfully fabricated and tested, but we now have enough
examples of quantum two-level systems and can move on to the
description of their behaviour.
Gate voltage
Bulk superconductor
|0 ( |1 )
(a)
Superconductor
|1
(b)
= 0 + 1 .
( ) (
P0 = P0 = * 0 + * 1 P0 0 + 1 = ;
2
)
( ) (
P1 = P1 = * 0 + * 1 P1 0 + 1 = .
2
)
5 That is, with zero scalar product: 0 1 = 1 0 = 0.
104
2 2
Geometrically, the requirement + = 1 simply means that
state vector must be normalized, that is
2
( )(
= 1 = * 0 + * 1 0 + 1 = ) 2
00 +
2
11 =
2 2
+ .
= cos ; = e i sin .
2 2
ez
|0
ey
y
ex
x
|1
0 0
= cos 0 + ei sin 1 = 0 ,
2 2
while for =
= cos 0 + e i sin 1 = e i 1 = 1 .
2 2
(Since we are at the pole, the azimuthal angle does not matter
any direction from the South Pole is due north!)
What happens when = /2? Then our vector lies in the xy-
plane:
= sin cos e x + sin sin e y + cos e z = cos e x + sin e y .
2 2 2
0 + e i 1 0 + ( cos + i sin ) 1
= cos 0 + e i sin 1 = = .
4 4 2 2
y
x
1
106
Now we see that the azimuthal angle of the vector is the
same, as the angle, which determines the phase of the complex
coefficient in the complex plane. As increases from zero
to 2, the complex number makes a counter-clockwise turn
about the origin in the complex plane and so does the vector
, turning counter-clockwise about the vertical axis. This is, of
course, true not only when lies in the xy-plane (Figure 4.6).
|0
x
|1
+
a * * +
a*
= (a b );(ab) = * .
b b
For example, a bra vector in the Hilbert space is the Hermitian
conjugate of the corresponding ket vector, and vice versa:
+ +
= ; = .
108
Qubit Hamiltonian and Schrdinger
equation
Since the Hilbert space of a qubit is two-dimensional, any
operator of a physical observable anything that can happen
with a qubit and can be measured is represented by a
two-by-two matrix. This is indeed the simplest one can go in
physics.
Such small matrices we can draw directly this is simpler and
clearer than a more general form (like (a)), which we had to
use when dealing with infinitely dimensional Hilbert spaces. It is
then convenient to represent the state vector as columns (for ket
vectors) or rows (for bra vectors) of two numbers:
= 0 + 1 = * * * *
; = 0 + 1 = .
( )
We see incidentally that using the standard rule for multiplying
matrices
2
2 2
= = ( * * ) = + = + ,
* *
as it should be.
Not any two-by-two matrix of complex numbers can represent
an operator of a physical quantity (an observable). The key
requirement is that this matrix must be Hermitian, that is,
possess a special kind of symmetry. Its diagonal elements must
be real, and off-diagonal ones complex conjugate, that is
a c id
A = ,
c + id b
E1
E0
110
namely, for the development of the theory of nuclear magnetic
resonance (NMR) an important physical effect, which is, among
other things, the basis of life-saving technology. This technology is
not usually called by its proper name NMR imaging but MRI,
magnetic resonance imaging, because of some peoples irrational
fear of the word nuclear although every single atom in their
body contains a nucleus, and some of these nuclei are naturally
radioactive.
d 1 d 1
i = ; i = ,
dt 2 dt 2
or
d i d i
= ; = .
dt 2 dt 2
1
= E .
2
112
The lower sign corresponds to the energy Eg of the ground state
g and the upper sign to the energy Ee of the excited state
e; see Figure 4.8. The explicit expressions for and in these
states (with the normalization condition 2 + 2 = 1) can be also
written easily enough, but for us it is sufficient to know that
1 0 0
g () = and e () = 1 , when ; g ( ) = 1 and
0
1 1 1 0
e ( ) = , when ; and g (0) = + and
0 2 0 1
1 1 0
e (0) = , when = 0.
2 0 1
This makes more sense if you look at Figure 4.8. You will see
that at a large bias (when the wells are far apart) the state of the
system is almost as if it was in either one, or the other of these
wells. In other words, the state vector of the system is either
1 0
L = or R = (calling them left and right is, of course,
0 1
E
|L |R
Ee
|L |R
Eg
Figure 4.8 two-well potential: energies and wave functions of the ground
and excited qubit states
t=0
t = T/2
t=T
e (0, t) = 1
2
L R exp i2t . (We have restored the time-
dependent exponents.)
At t = 0 the system is in state L, which can be written as
1 1
L =
2
( ) (
2 )
L + R + L R = 2 g (0, t = 0) + 2 e (0, t = 0)
.
114
1
(t) = 2 g (0, t) + 2 e (0, t)
2
1 i t
i t
i t
i t
= L e 2 + e 2 + R e 2 e 2
2
t t
= L cos + i R sin .
2 2
This is interesting. Initially (t = 0) our qubit was in the left
state. We do absolutely nothing with it but at the time t = /
it will be in the right state. And it will keep oscillating between
these two states. The probability to find the qubit in state L or
R is
2
PL (t) = L (t) = cos 2 ( )
t 2
=
1
2 (1 + cos ) or
t
2
PR (t) = R (t) = sin 2 ( )
t 2
=
1
2 (1 cos ) respectively.
t
d
= .
dt
= e x + ez .
z
W
x
Figure 4.10 time evolution of the Bloch vector
dA/dt
116
(Approximately, because the Earth also rotates around the Sun, and
its axis also wobbles slightly because of the imperfectly spherical
shape of the globe and the influence of other celestial bodies.)
If a vector A of constant length rotates about its origin with the
angular velocity , its time derivative is given by a simple formula,
which involves the cross product of two vectors:
d
A = A,
dt
that is,
AB
A B = AB sin ,
118
obtaining an arbitrary state of a qubit starting from some other
state remains the same.
z
|0
|
y
x
|1
Figure 4.11 controlling the Quantum state of a qubit
a+b ab
0 0
2 2 + 0 c + 0 id
A = +
a+b a b c 0 id 0
0 0
2 2
a+b 1 0 ab 1 0 0 1 0 i
= + + c +d .
2 0 1 2 0 1 1 0 i 0
0 1
x = ( * *)
* *
= + ,
1 0
0 i
y = ( * *) * *
= i + i ,
i 0
1 0 2 2
z = ( * * ) = .
0 1
120
Spotlight: Wolfgang Pauli
Wolfgang Pauli (19001958) was an Austrian physicist, Nobel Prize
winner (1945) and one of the founders of quantum physics. He
discovered the Pauli principle (that two electrons cannot share the
same quantum state), proved the spin-statistics theorem (which
we will discuss later) and predicted the existence of neutrino.
Pauli did not tolerate vague and nonsensical arguments, labelling
them not even wrong.
tr z = tr x = tr y = 0.
x x = 1; y y = 1; z z = 1,
and that
x y = i z; y z = i x; z x = i y .
d
i = [ H , ].
dt
1 1
= ( x x + y y + z z + 0 ) = ( + 0 ).
2 2
The role of the unit matrix is here simple: since all Pauli
matrices have trace zero, the unit matrix must ensure that
the density matrix has unit trace. All the information about
the qubit state is contained in the three coefficients x,y,z,
which we have gathered into the single vector (which in the
expression for the density matrix is dot-multiplied by the
vector all this happens in the space of all two-by-two
Hermitian matrices).
A simple example is the density matrix of a qubit, which is in
a pure state described by a state vector . It can be written
simply as
= ,
that is,
2
*
= ( 0 + 1 ) ( * 0 + * 1 ) = 2
*
=
1
2{( ) ( ) 2
( 2
) ( 2 2
* + * x + i * * y + z + + 0 . ) }
You can check that all the coefficients in the round brackets are
always real numbers.
An important point is that the vector in the expression for the
density matrix is the same Bloch vector, which we introduced
earlier in this chapter when describing pure states of a qubit.
While the Bloch vectors of pure states had unit length, with
their ends on the surface of the Bloch sphere, the ends of the
Bloch vectors describing mixed states are inside the Bloch
sphere (Figure 4.12).
122
z z
q q
y y
x x
|1 |1
As before, the Bloch vector of a mixed state will rotate with the
angular velocity determined by the Hamiltonian of the qubit,
d
dt
= But now we can take into account the influence
of the environment. The resulting equations are also called the
Bloch equations and have the following form (Figure 4.12):
d
x = y z z y x ;
dt T2
d y
y = z x x z ;
dt T2
d z z
z = x y y x .
dt T1
d M
Mx = y Mz z My x ;
dt T2
d My
My = z Mx x Mz ;
dt T2
d Mz Mz
Mz = x My y Mx .
dt T1
124
H m
Larmor precession
Dephasing
Relaxation
Figure 4.13 Dephasing and relaxation in nmR
NM = = Ntr( ),
where = is the operator of the magnetic moment, is some
constant coefficient, and has Pauli matrices as its components.
126
important, but not relevant here, reasons in the experimental
device the island is attached to the reservoir in two spots).
Another insulating layer separates the island from the probe
electrode, which is eventually connected to a sensitive ammeter.
The electrodes (bright) are made of aluminium deposited on
the insulating substrate (dark). In Figure 4.14 the island, the
reservoir and the probe seem to be of one piece, but in reality
they are placed on top of each other, with very thin insulating
barriers in between.
Reservoir
DC gate
Probe
Box
1 m Pulse gate
8 This means, in particular, that all the connections between the qubit
and the electrical equipment outside the fridge must be very thoroughly
engineered to avoid heat and noise entering the fridge.
1
H =
( )
Vg
.
2
( )
Vg
128
15
h/Tcoh
80
Microwave spectroscopy
EJ (eV)
Pulse-induced current (pA) EJ(=0) cos (/0)
40
10
0
0.0 0.2 0.4 0.6
/0
0/e = 0.51
5 /0 = 0.31
Tcoh
0
0 200 400 600
t (ps)
Gates
1.5
tp (ns)
0.5
np 0.5
Figure 4.16 Quantum beats in a double-quantum dot charge qubit (from
hayashi et al. 2003: 226804)
130
can again control whether the state 0 (the extra electron is on
the left dot) or the state 1 (the extra electron is on the right
dot) has lower energy. (The whole device, of course, had to be
placed inside a dilution fridge and cooled down.)
The experiment was essentially the same as with the
superconducting charge qubit: starting from the state 0 (extra
electron on the left dot) we bring the qubit to the position when
the quantum beats begin, wait for a period of time , and then
measure the current from the right dot. Do this many times.
Then change and repeat. The quantum beats, obtained after
such a procedure, are seen quite clearly though not as well as
in the superconducting device. (We have already stated that
superconductors have intrinsic advantages as qubit material).
The devices in the pictures are quite small about a micron
across. Nevertheless they already contain a very large number
of atoms, while showing a typically quantum behaviour. This
was one very important result, supporting the opinion that in
quantum mechanics size is not everything: it seems you can be
both big and quantum.
? Fact-check
1 How many quantum spin states does a system with spin 3/2
have?
a 3
b 4
c 8
d 2
132
9 NMR uses the following physical process:
a compression
b evaporation
c precession
d expansion
Dig deeper
Further reading I. Quantum mechanics of two-level systems.
R. Feynman, R. Leighton and m. sands, Feynman Lectures on
Physics, Vols i, ii and iii. Basic Books, 2010 (there are earlier
editions). Vol. iii (chapters 6, 7, 10, 11).
Further reading II. Quantum bits.
m. Le Bellac, A Short Introduction to Quantum Information and
Quantum Computation. cambridge university Press, 2007.
chapter 6.
a. m. Zagoskin, Quantum Engineering: Theory and Design of
Quantum Coherent Structures, cambridge university Press, 2011.
chapters 13.
134
Measurement, projection postulate
and Borns rule (revisited)
The result of a measurement is that the state vector of the
quantum system is projected on one of eigenvectors ai of one
of the observables , which characterize the system. (Which
observable it will be depends on the apparatus.) At the same
time, the apparatus will evolve from its initial state to some final
classical state, which is determined by which eigenvector the
quantum state was projected to (Figure 5.1). This is called the
measurement of the observable .
QUANTUM CLASSICAL
a1
a2
| a1
| a2
|
| a3 Apparatus a3
| a4
| a5
a4
a5
a1
| | a2
a2
Apparatus a3
a4
a5
136
(t2 ), without any randomness involved. What we have
here instead is that the quantum state vector is turned into a
collection (i.e. ensemble in the sense of statistical mechanics)
of possible state vectors, out of which a random choice is
made every time this measurement of the same quantum state
vector is repeated. The results of this choice are described by
a probability distribution, the set of probabilities Pi. Unlike
classical statistical mechanics, though, this randomness is
intrinsic. It is not due to the complexity of the system, and it
cannot be removed by any Laplacian demon, however powerful.
This is worth repeating:
1 If the system was in a mixed state to begin with, the outcome will be
the same: after a measurement of an observable , with certain probability
Pi, the system will be found in an eigenstate ai of this observable. The
formula for calculating Pi is just slightly more complex than for the initially
pure state.
138
practical engineering. This is why the so-called measurement
problem, along with the intimately related problem of
quantum-classical transition, was forcefully pushed to the
forefront of research.
Bulk
superconductor Probe
electrode
0 0 l
2 Why is the Bloch vector in the yz and not in the xz-plane (or any other
plane containing the z-axis)? It is because of the imaginary unit i multiplying
the coefficient at 1. Recall that the phase shift between the complex-
number coefficients in the expression for the state vector is reflected in
the azimuthal angle of the Bloch vector.
140
z
|0
|1
1 0 0
n = ( 0 z ) = .
2 0 1
0 0 1 0
n 0 = = =0 0 ,
0 1 0 0
0 0 0 0
n 1 = . = =11 .
0 1 1 1
142
After the measurement of the observable n the quantum
state vector will be projected on one of its eigenstates,
either 0 or 1. The corresponding probabilities are
0 0
= 2e n = 2e tr( n ) = 2e tr P0 ( )
Q
0
0 P1 ( ) 0 1
0 0
= 2e tr = 2eP1 ( ).
0 P1 ( )
1 1 0
H0 =
2 2 0
1 0
H1 = and 1 ex .
2 0
144
Now the angular velocity once again points along the z-axis.
Therefore the Bloch vector will keep turning around it, tracing
one of the parallels on the Bloch sphere, but this rotation is not
important. Indeed, it does not change the z-component of the
Bloch vector, and this is loosely speaking the component we
are measuring.
The measurement, that is, the escape (or the non-escape) of the
two electrons from the island to the probe electrode, reduces
the pure state (with its Bloch vector lying on the surface of the
Bloch sphere) to a mixed state (with the Bloch vector inside the
Bloch sphere). This density matrix is
P0 ( ) 0
P0 ( ) + P1 ( ) P ( ) P1 ( )
( ) = = 0 + 0 z.
0 P1 ( )
2 2
( ) = 1
2 ( 0 )
+ x x + y y + z z . Therefore the Bloch vector
|0
Projective
measurement
q
q proj
y
|1
|0
q qproj
|1
Figure 5.5 superconducting charge qubit experiment: weak continuous
measurement
146
During this time the Bloch vector will keep rotating about the
z-axis, approaching it along a kind of a spiral. This spiral is
conveniently and accurately enough described by the Bloch
equations in Chapter 4 (Figure 5.5):
d x y d y y d z
= y z z y ; = z x x z ;
dt T2 dt T2 dt
= x y y x .
3 Note that the relaxation time T1 does not appear in the above equation:
the measurement by itself does not bring the system to an equilibrium
state. In the language of the Bloch vector and Bloch sphere, during the
measurement process the Bloch vector will spiral in the same plane, as in
Figure 5.6.
148
Figure 5.6 Weak continuous measurement (Zagoskin 2011: 266)
p F p F
e p + p e
p + p
150
S
The light pressure is P = , where S is the energy flux in the
c
electromagnetic wave (that is, the amount of light energy
passing through unit area per unit time), and c is the speed of
light. If A is the area of the particle, which scatters light (called
cross-section), then the light will push the particle with force
SA W
F = PA = c = c . Here W is the light power (energy per unit
time) scattered by the particle.
Suppose we illuminate the particle for time t. Then light will
transfer to its so-called impulse F t = Wct = cE , which is
proportional to the energy E of light, which was scattered
by the particle during the observation. It is known from
Newtonian mechanics that an impulse changes the momentum
of the body it acted upon: p = Ft. Therefore the momentum
of the particle will inevitably change after and because of the
observation of its position, and this change is p = F t = cE .
Can this be avoided? In classical physics certainly: it would be
enough to make E arbitrarily small. (In a Gedankenexperiment
you do not care about the implementation, only about what is
possible in principle.) In quantum physics, not so: we have seen
that the energy of a quantum oscillator can only change in portions
of E = h, and this holds true for light as well. Therefore
the minimal change in the particle momentum due to the
observation of its position through the Heisenberg microscope is
h h h
p = = .
c x
152
Standard quantum limit
Suppose now that we want to repeatedly measure the
momentum of the same particle. The straightforward (and
rather impractical, but we are still in the Gedankenexperiment
mode) way of doing it is to measure its position at times t = 0
and t = t calculate
p = mv = m(x(t) x(0))/t.
The first and second position measurements were made with
uncertainties x0 and xt . Moreover, because of the first
measurement the momentum was changed by at least p =
/2x0, and the velocity by v = /2mx0. Therefore this will
produce an additional uncertainty of the particle position at
time t:
xt = v t =
2 m x0
t
m
pSQL = .
t
1 t
xSQL = .
2 m
154
search for the gravitational waves requires very precise
measurements of macroscopic systems (e.g. the position of a
massive slab of metal, which would be very slightly displaced
by a passing gravitational wave), and since the 1960s it was
understood that quantum effects would be important. Though
no direct observation of gravitational waves was achieved so
far,5 the effect of this research on quantum theory was profound
and very fruitful. Yet another confirmation of the old adage: in
science, a negative result is still a valuable result.
In precise measurements it is important to determine all possible
sources of errors. Assuming that classical noise (such as thermal
fluctuations in the system itself and ambient noise reaching it
from the outside) is suppressed (by lowering the temperature of
the system and properly shielding it), the accuracy seems to be
restricted by the standard quantum limit. And it is not.
Here we should be cautious. If we do not know what the
quantum state of the system we are going to measure is, then on
average the expectation value of an observable can be found
with an accuracy no better than the standard quantum limit. This
was the standard experimental situation for quite a long time
after the discovery of quantum mechanics. The situation changed
when it became possible to control quantum systems better
in particular, to perform repeated measurements of the same
observable in the same quantum system.
Why was it a game changer? Recall the projection postulate:
after the measurement the system ends up in an eigenstate
of this observable. Therefore if we immediately measure the
same observable again, we will with certainty obtain the same
eigenvalue of this observable, with (at least in theory) 100%
precision, and the system in question will be found in the same
quantum state (e.g. aj) as before the observation. Therefore
this kind of measurement is called a quantum non-demolition
(QND) measurement: the state of the system is not demolished
and can be observed again and again.
( )
of ), but also H s + H a a j = hj a j . If substitute the latter
expression in the Schrdinger equation, we see that
d
i a j = H a j = hj a j .
dt
156
energy of a closed system) are called the integrals of motion,
and so they are called in quantum mechanics.
This so-called strong QND condition is pretty restrictive. (There
is also a weak QND condition, which allows non-demolition
measurement of other observables as well, but as a price it
requires a subtle manipulation with the quantum state, which is
at the moment rather impractical.) Nevertheless it is satisfied by
some important observables, such as the Hamiltonian itself.
Recall that the time-independent Schrdinger equation,
H = E , clearly says that the eigenvalues of the
Hamiltonian are the allowed energies of the system. Obviously,
the Hamiltonian commutes with itself, and the strong QND
condition is satisfied. Therefore the simplest example of a QND
measurement is the measurement of energy.
Let us take a quantum particle, freely flying in space. One
(rather Gedankenexperiment-esque) way of measuring its
energy is to measure its momentum and calculate E = p2/2m
(we do not bother here with relativistic particles). When
considering a similar measurement earlier on in the part on
Standard quantum limit, we assumed xt = 0 , since we did
not care that such a precise measurement of the position will
change the momentum afterwards. Now we can make xt =
x0 instead, repeat the calculations and see that the uncertainty
of momentum measurement still depends on the duration of
measurement as m / t (up to a slightly different numerical
factor). Therefore it can be made arbitrarily small at the expense
of a longer measurement time (and a totally undetermined
position of the particle). The energy uncertainty is then
(p)2
E = , and we come to the energytime uncertainty
2m 2t
relation:
E t .
2
This is a very useful relation. For example, it tells how long
it will take you (in the very least) to measure energy of some
quantum object to the desired precision. And if you want to
change the quantum state of a qubit, you cannot go too fast,
if you do not want to change the energy of the qubit too much.
158
Aristotle in his Physics recounted this paradox as follows: If
everything when it occupies an equal space is at rest, and if that
which is in locomotion is always occupying such a space at any
moment, the flying arrow is therefore motionless.
In other words, if a moving object at any moment of time can be
found at some particular place, then it is not different from an
object, which would be resting at the same place at that time.
Therefore a flying arrow must be at rest.
Here ( E02 ) is the dispersion of energy in the state a0, that is,
( ).
2
E02 = a0 H 2 a0 a0 H a0
160
? Fact-check
1 The measurement requires
a a measuring device, which obeys the laws of classical
physics
b a measuring device, which obeys the laws of quantum
physics
c an observer, who can manipulate the measuring device
d a sentient being capable of making an observation
3 Quantum randomness is
a the result of interactions of quantum system with its
environment
b a fundamental property of Nature
c the expression of our limited knowledge
d the result of measurement
162
9 The problem of quantum-classical transition is about
a how the laws of quantum mechanics emerge from the
laws of classical mechanics
b how the laws of classical mechanics emerge from the
laws of quantum mechanics
c what is the source of dephasing
d how to resolve the internal contradictions of quantum
mechanics
164
Strange and
6
unusual
We have spent much effort on elucidating both similarities and
dissimilarities between quantum and classical mechanics. Some
of them are pretty subtle. But the ability of a quantum system to
go through solid walls the quantum tunnelling was certain
to attract much attention as soon as it was discovered that it
must follow from quantum theory.
165
We have taken pains to underline the deep similarities between
quantum and classical mechanics, in order to better understand
the fundamental differences between them. In short, these
differences are that:
In classical mechanics the state of the system is essentially
identical to the set of physical quantities which characterize the
system in this state; they can be considered the coordinates of
the state vector.
In quantum mechanics the state and these quantities are
separate: the observables are operators, which act on the state
vector.
In classical mechanics all aspects of a systems evolution are in
principle fully deterministic.
Quantum mechanics is fundamentally stochastic: the state
vector satisfies a fully deterministic Schrdinger equation.
Nevertheless the state vector generally does not determine
the result of a single measurement (i.e. the outcome of a
specific instance of quantum systems interaction with a
classical apparatus), but only its probability. Moreover
and importantly, this probability is a nonlinear (quadratic)
function of the state vector.
166
In 1919 Rutherford artificially transmuted nitrogen to oxygen by
bombarding it with particles.
His most celebrated discovery was his model of an atom, deduced
from the scattering of particles. Scattering experiments remain
the backbone of high-energy physics, including the Large Hadron
Collider (LHC).
During his Nobel ceremony Rutherford remarked that he had
dealt with many different transformations with various time
periods, but the quickest he had met was his own transformation
from a physicist to a chemist.
168
escape the mug at 300 K is, as you can easily see, approximately
exp [ 1.7 106 ] 1. This stands to reason: if such light atoms
as helium-4 could not fly at room temperature, nor would
much heavier nitrogen or oxygen molecules, and most of the air
would just drop on the floor and stay there (forming a snow
this happens on planets far enough from the Sun).
The main point of the comparison is however that either an
ant or a helium atom, which escapes the cup, must have energy
enough to clear the top of the potential barrier U, so when it
will drop back to the ground level it will have at least this much
kinetic energy: energy is conserved.
This created a major problem when physicists began seriously
thinking about the mechanism of -radiation. It was known
due to Rutherford and Soddy that -particles are emitted by the
nuclei of certain radioactive elements, and that -particles are
the nuclei of helium-4. Positively charged -particle is strongly
repulsed by the positively charged rest of the nucleus and
therefore must be held inside by some non-electromagnetic force
(so-called strong force). Imagine the potential of this force as a
mug (properly called a potential well). Its outer side represents
the repulsive Coulomb potential, which will accelerate the
-particle once it is out of the nucleus.
The nature of strong force was understood only much later.
But whatever it is, if an -particle escapes from the nucleus, it
should have at least enough energy to jump the potential barrier.
Then it will accelerate away along the external potential slope.
Far enough from the nucleus the repulsive Coulomb potential
will become negligible, and the kinetic energy of the -particle
will be equal to its potential energy at the moment it cleared
the top (assuming it did so with zero kinetic energy but this
assumption is not actually important).
170
energies (4-5 MeV). Finally, it was established early on that
heating or cooling does not affect radioactivity at all.
George Gamow solved the puzzle by applying to it the new
quantum mechanics. He considered the Schrdinger equation
for a particle in a potential well (Figure 6.3). The external slopes
go as 1/r representing the Coulomb potential, and the internal
shape of the well could be for the moment approximated by a
vertical wall producing a shape similar to the nautical mug,
which is hard to knock over.
n
p p
n
|(r)|2 n
p p
U(r) n
Let us first neglect the outside of the mug and solve the
stationary Schrdinger equation inside the well:
2
2m + U(x, y, z) (x, y, z) = E (x, y, z).
It will have only a discrete set of solutions with energies E1, E2,
E3, etc., as we have discussed in Chapter 3. But what is crucially
important is that these solutions do not immediately go to zero
inside the wall, though they drop quite fast exponentially fast.
Recalling that the square modulus of the wave function gives
the probability of finding the particle at a given point, we see
that a quantum particle can penetrate inside the potential
barrier. (Of course, a classical particle would be just reflected.)
Now you already guessed what happens to -particles. If the
thickness of the mug wall is not too great, the tail of the
172
Gamov was able, using this model that is, a more elaborate
and accurate one than the sketch presented here to explain the
GeigerNuttall rule and reproduce the experimental data.
U(x)
(x)
x
2m dx 2
at x > 0. You can check that the first equation has a solution
(
one is satisfied by (x ) = C exp k (E )x ) with k(E ) = ( E ) / 2 .
2mV
Here A, B and C are some constants we do not care about at the
moment they are determined from the condition that the wave
function behaves well at x = 0. As you can see, the wave function
indeed does not immediately drop to zero inside the potential
barrier, but of course the taller the barrier compared to the
energy of the particle, the faster it drops.
174
V
3 We could also heat the system, but this would drastically increase the
amplitude of the thermal motion of atoms both in the substrate and in the
tip, defeating the purpose of the whole exercise.
4 One electron Volt (1 eV) is the energy that an electron gains after
passing through the voltage difference of 1 Volt. This energy is equivalent to
the temperature of approximately 11,600 K. The temperature of the surface
of the Sun is only about 6000 K.
176
their contributions to the development of wireless telegraphy and
the 1912 one to Nils Daln for his invention of automatic regulators
for use in conjunction with gas accumulators for illuminating
lighthouses and buoys and to the ones given for the invention
of cloud (C.T.R. Wilson 1927) and bubble (Donald Glazer 1960)
chambers (at the time, indispensable tools for particle physics);
of the cyclotron (E. Lawrence 1939); of the laser (Charles Townes,
Nicolay Basov and Aleksandr Prokhorov 1964); and of the transistor
(William Shockley, John Bardeen and Walter Brattain 1956).
Figure 6.6 the stm tip can not only feel, but also move single atoms and
place them at the desired spot. here you see the stm images of corrals
built of iron atoms on the (111)-surface of a copper crystal. the ripples are
those of the electron density (roughly, the square modulus of the electron
wave function). in the case of a round corral the solution is given by a
Bessel function (one of mathematical functions akin to sine and cosine).
using an stm is probably the most expensive way of tabulating this
function. (a. http://researcher.watson.ibm.com/researcher/files/us-flinte/
stm15.jpg b. http://researcher.watson.ibm.com/researcher/files/us-flinte/
stm7.jpg
178
Waves like particles: photoelectric
effect and the single-photon double
slit experiment
It has been known since the 19th century that light is a wave
and that its properties were not quite like the properties of a
sound wave. Sound in gases and liquids is a longitudinal wave
(that is, the particles oscillate along the same direction in which
the wave propagates); sound in solids has both longitudinal
and transverse components (in a transverse wave the particles
oscillate in the direction, which is orthogonal to the direction of
its propagation). But light and it was established beyond any
doubt had only a transverse component, so it could not be a
wave in an ordinary liquid or solid, and the attempts to explain
it as a wave in a special substance called aether run into
insurmountable difficulties. These difficulties were got rid of by
getting rid of aether altogether and recognizing that light (or,
more generally, the electromagnetic field) is just a special kind
of matter governed by Maxwells equations. In the process the
special relativity was built primarily by the efforts of Lorentz,
Poincar and, of course, Einstein (and a very important later
contribution from Minkowski), but this is not our concern here.
We have mentioned earlier in Chapter 2 the photoelectric effect.
If a metal is illuminated, it may under certain condition emit
electrons. At first, nothing strange was seen about this. Electrons
are electrically charged, light is an electromagnetic wave,
and under the influence of oscillating electric and magnetic
fields an electron can absorb enough energy to get loose and
escape the metal (where it is held by the collective force of
electrostatic attraction of all the negatively charged electrons
to all the positively charged ions see Chapter 8). The flux of
emitted electrons the photoelectric current was found to be
proportional to the light intensity (so-called Stoletovs law, after
Aleksandr Stoletov (18391896)), as one would expect.
Nevertheless there were features of the phenomenon that
refused to fit this picture. To begin with, there was no electron
emission if the light frequency was below a certain threshold
value 0 (different for different materials), no matter how bright
h
W = h 0
(work function)
Electrons
Conductor Vacuum
Figure 6.7 Photoelectric effect and Lenards formula
al 4.1
au 5.1
Be 5.0 1200 ultraviolet
ca 2.9 700 violet
cu 4.7
Fe 4.5
180
hg 4.5
K 2.3 550 green
Pb 4.1
182
form two blurred lines instead they produced an interference
pattern, as if each single photon was a wave in and of itself and
passed through both slits at once (Figure 6.8).
2 2 2 2
+ + ( x, y, z ) = E( x, y, z ).
2m x2 y 2 z 2
184
We must find its solution with appropriate boundary conditions.
It is reasonable to demand that the wave function is zero on and
inside the impenetrable screen (the particle has zero probability
to get there, so ||2 should be zero as well). The wave function
before the screen should look like a combination of incoming
and reflected waves. This equation is exactly like a wave
equation from the classical theory of light and the mathematical
methods for solving such equations were worked out long
ago we have already mentioned this in Chapter 3.6 Therefore
it is not surprising that the probability |(x, y, z0)|2 to find the
particle at some point (x,y) in the picture plane z = z0, which
follows from its solution, demonstrates the tell-tale interference
pattern, and not just two blurry strips expected from a flow of
classical particles (Figure 6.9).
This is a very remarkable result. Here too all attempts to
explain away the wave-like behaviour of particles by assuming
that it happens only when there are many of them, somehow
forming a wave, were futile. Experiments with single electrons
and other particles sent one by one towards a double slit (or
a more elaborate structure, like a number of parallel slits
so-called diffraction grating) demonstrated the same result:
an interference pattern. In the two-slit case this can only
mean one thing: a particle passes through both slits. In case
of a diffraction grating it passes through all the slits at once.
Nevertheless it is registered as a particle by a detector (e.g. a
sensitive film) in the picture plane. Recently such behaviour was
observed for a molecule containing almost a thousand atoms,
weighing about 10,000 atomic mass units (that is, about 10,000
times heavier than a proton or a neutron, and about 20 million
times heavier than an electron).
In its time this experimental result produced a kind of shock,
despite the fact that Louis de Broglie predicted the outcome
three years earlier. The particle-wave duality became a stock
phrase, with some unfortunate consequences. Not for physics
or physicists physicists know what that means (that quantum
objects are neither particles nor waves, though in certain
Spotlight: Louis-Victor-Pierre-Raymond,
Duke de Broglie
Louis-Victor-Pierre-Raymond, 7th Duke de Broglie (18921987)
was not a duke (he inherited the title from his brother later) when
he formulated his famous hypothesis, which was rewarded with
the Nobel Prize for Physics in 1929 (after it was confirmed by the
electron diffraction experiments).
186
nh
L = n = .
p
R
R R
R R
R R R R
NH N
R R
N NH
R R R R
R R
R=F
C6F13
or = S
R R
C6F13
R
2000
1600
1200
800
400
0
0 100 200 300 400 500 600 700 800 900 1000 1100
z/nm
Figure 6.9 here the quantum interference is demonstrated with heavy
molecules: their masses exceed 10,000 amu (approximately 10,000
masses of a hydrogen atom) (eiberberger et al., 2013, 1469614700, Figs
1 and 3)
= h/p.
? Fact-check
1 Tunnelling was proposed as the mechanism of -decay, because
a the energies of particles do not depend on temperature
b the energies of particles coming from the same decay
are identical
c the energy of particles strongly depends on the half-life
of the -radioactive element
d radium-222 and polonium-210 have almost identical
half-lives
188
4 The subatomic STM resolution is due to
a the subatomic scanning tip
b the exponential dependence of tunnelling probability on
distance
c the shortwave probing radiation
d electrons being smaller than the nuclei
Dig deeper
R. Feynman, R. Leighton and m. sands, Feynman Lectures on
Physics. Basic Books, 2010 (there are earlier editions). Vol. i,
chapter 37 and Vol. iii, chapters 13.
e. h. Wichmann, Quantum Physics (Berkeley Physics course, Vol.
4), mcGraw-hill college, 1971. chapter 7.
190
The game of
7
numbers
The general formalism of quantum mechanics, which we have
described so far, does not care about the number of particles in
the quantum system. The quantum state satisfies the same
Schrdinger equation, i t = H , whether it describes
a single electron, an electron beam, an atom, a molecule or
a qubit (Figure 7.1). Similarly, the expectation value of an
observable A is given by A , and the Heisenberg equation
for it is always i dA = [ A , H ]. The devil is, as usual, in the
dt
details. We know at least, nominally how to find the wave
function (x, y, z , t) , which represents the state vector of
a single quantum particle (e.g. an electron). Using it for an
atom or a molecule (like we did describing certain quantum
interference and diffraction experiments) is already a stretch
an approximation, which only holds while we can neglect both
the size of the object so described and its interaction with other
such objects (and, however strange that sounds, the latter can
never be neglected we shall see why later in this chapter).
Therefore single-particle quantum mechanics only takes us so
far. It needs further elaboration.
191
me
R
+
mp
Electron in an atom
Before qubits, the simplest quantum system used as a textbook
example was the hydrogen atom. The atom consists of the
nucleus (a single proton with the electric charge +1 in electron
charge units) and an electron. The proton is almost 2000 times
more massive than the electron1, and therefore can be, to a
good accuracy, considered as nailed down being at rest at
the origin. Then its only role (neglecting some subtler effects we
are not ready to treat yet) is to be the source of the electrostatic
Coulomb potential,
e
(r) = + r ,
acting on the electron.2 The stationary Schrdinger equation for
the electron with energy E is thus
2 2 e2
( R ) ( R ) = E( R ).
2me r
192
Here R is the radius vector pointing at the electron; its length is
r. The symbol 2 is pronounced nabla square and is called the
Laplace operator, or simply the Laplacian. This is shorthand
for the expression we used before: for any function of Cartesian
coordinates (x, y, z):
2 2 2
2 f (x, y, z) = 2 + 2 + 2 f (x, y, z).
x y z
r
y
2 1 2 2 1
r (r , , ) sin (r , , )
2me r r r
2
2me r sin
2
2 1 2 e2
(r , , ) (r , , ) = E (r , , ).
2me r 2 sin2 2 r
But in our case, when the potential depends only on the distance
r between the electron and the proton, it is not convenient to
use the Cartesian coordinates. The solution is greatly simplified
if spherical coordinates are used instead (recall Chapter 4), and
in these coordinates (r, ,) the Laplacian looks quite differently.
There are other coordinate systems as well, fitted to specific
problem symmetries. Using everywhere the same symbol 2
shortens equations and allows us not to specify a particular
coordinate system.
We chose to write the Coulomb (electrostatic) energy of the
proton as er . (As we know from classical mechanics, any
2
194
( r ) = C exp r E me / 2 , r .
2 1
En =
2me a02 n2
2me a0
is called a Rydberg (and denoted by Ry). This is the ionization
energy the minimal energy one must supply to the hydrogen
atom in this state to break the electron loose and produce the
positively charged hydrogen ion that is, a proton.
The quantum formula for the electron energy levels explains
the emission (and absorption) spectra of hydrogen (Figure 7.2).
Suppose the hydrogen atom emits a photon. This can only
happen if an electron drops from a level Em to a level E n (m > n).
The energy difference, carried away by the photon, is
1 1Ry 1 1 1 1
= = RH 2 2 .
hc n2 m2 n m
0 l n 1; l m l .
E=0
n=4
n=3
n=2
1 Ry
n=1
me e
millionth part of a centimetre).
196
Spotlight: Johann Balmer and Johannes Rydberg
Johann Jakob Balmer (18251898) was a Swiss mathematician,
who in 1885 found an empirical formula that quantitatively
described the frequencies of visible light emitted by hydrogen.
(Some say he did it on a bet a friend dared him to find a simple
formula fitting the hydrogen spectroscopic data.)
In 1888 Johannes Robert Rydberg (18541919), a Swedish
physicist, found the general phenomenological formula, which in
modern notation looks like:
1 1 1
= RH .
n 2
m2
Balmers formula corresponds to the case n = 2.
That is, for each n there are n solutions with different l, and
for each l there are (2l+1) solutions with different m. The total
number of independent solutions with the same energy En is
thus N n = l =0 (2l + 1) = n2.
n 1
Figure 7.3 shapes of s, p, d, f wave functions (the shade marks the sign of
the wave function) (sarxos 2007)
198
quantum numbers can be shared by two electrons, no more,
and they are being taken starting from the lowest energy state
like a rooming house being occupied starting from the ground
floor (Figure 7.4).
1s 2s 2p 3s 3p 3d
200
Remarkably the scheme generally holds even though the
hydrogen-like formulas only work properly either for one-
electron ions (when a single electron remains bound to the
nucleus), or for the Rydberg atoms, when one electron occupies
such a high energy level n>>1, that on average it is very far
away both from the nucleus and from the rest of electrons,
which keep close to the nucleus. Generally, hydrogen-like
electron wave functions are distorted by electronelectron
interactions out of all recognition. For example, in a carbon
atom, its four electrons with n = 2, instead of occupying a
spherically symmetric s-state and dumbbell-like p-states along
x, y and z-axes, occupy four identical states directed at the
vertices of a tetrahedron (Figure 7.5). (This is very lucky for
us it is this electronic structure of carbon that makes it such a
useful building block for all organic compounds, including those
our brain is made of.) Nevertheless, the systematics of electron
states based on the hydrogen-like solutions of a single-electron
Schrdinger equation for some reason remains valid.
202
Case study: Principle of indistinguishability
of elementary particles
The principle of indistinguishability of elementary particles states
that there is absolutely no way to tell apart any two particles of
the same kind. You cannot do so even in a Gedankenexperiment
(e.g. by painting one electron red). This principle is supported by
experimental evidence. That is, if it were not satisfied, the observable
properties of matter and radiation would be massively different.
204
This leaves us with two possibilities. Either =1 (that is, the
wave function does not change at all, if two identical particles
are exchanged), or = 1 (the wave function changes sign). In
the former case we have bosons named after Satyendra Nath
Bose, and in the latter fermions.
A rule of thumb allowing us to distinguish between bosons
and fermions is, that fermions are matter (or substance), while
bosons are fields. For example, substances are made of atoms,
atoms are made of electrons, protons and neutrons, and protons
and neutrons are made of quarks. All of these are fermions with
s = 1/2. On the other hand, light quanta photons are bosons
(with s = 1), and all electric and magnetic forces binding electric
charges also can be described in terms of photons. The celebrated
Higgs boson has spin s = 0. It is expected that the gravity
quanta are bosons with s = 2. (They were not observed yet and a
satisfactory quantum theory of gravity remains to be built.)
The distinction agrees with our intuition that substances
occupy space, and you cannot put one substance into the space
already occupied by another.
ci+ n1 , n2 , ni , n1 , n2 , ni + 1, ;
ci n1 , n2 , ni , n1 , n2 , ni 1, .
The factors in these expressions are different for bosons and for
fermions.
206
First, note that creation and annihilation operators c+i and ci do
not correspond to any observables. They are not Hermitian
as a matter of fact, they are Hermitian conjugates of each
other (we have seen examples of this in Chapter 4). But their
combinations may be Hermitian.
( )
i c + c = i c +
( ) (c ) = i (c )
+ + +
+
j cj .
j j
j j
ck+ c j n j , nk , n j 1, nk + 1, .
{c , c } = c c
j
+
k
+
j k { } { }
+ ck+ c j = jk, and c j , ck = c +j , ck+ = 0,
Fourth, the quantum state, which does not contain any particles,
is called the vacuum state and is denoted by 0 :
0 = 0,0,0,0, .
This is, of course, the lowest energy state the ground state of
the system. Since there are no particles to annihilate, if you act on
the vacuum state by any annihilation operator, you obtain zero:
c j 0 = 0.
Note that the vacuum state is not zero a vacuum is not nothing!
On the contrary, any physical state of the system can be obtained
by acting on the vacuum state by the appropriate creation
operators as many times as necessary (the nth power of an
operator means that the operator was acting n times in a row):
( ) (c )
nj + nk
n j , nk , c +j k n j , nk , .
The vacuum state and all states that can be obtained from it
in this way are called the Fock states. They form the basis of a
special kind of Hilbert space. It is called the Fock space (after
Vladimir Aleksandrovich Fock).
Let us see how this works. We will start with fermions.
The anticommutation relation for two Fermi creation
operators means that we cannot create more than one particle
in a given state (say, state j). Indeed, in order to create a
208
quantum state with two fermions in a state j, one must twice
act on the vacuum state by the creation operator c +j . Since
{c , c } = 2c c ( )
+ + + + 2
j j j j = 2 c +j = 0, we will obtain zero as a result of
this operation. Therefore the only possible states for a many-
fermion system are of the form
0,1,1,0,1, .
The operator
N j = c +j c j
c +j c j n j , nk , = n j n j , nk , ,
that is, it does not change the vector |, but simply multiplies
it by the number of particles nj in the single-particle state j (for
example, by the number of particles with a given momentum
and spin). As you recall, this is another way of saying that
N j = N j = n j = n j = n j .
c j = n1 , n2 , c j n1 , n2 , = 0; c +j = n1 , n2 , c +j n1 , n2 , = 0
210
This is reasonable: either of these operators changes the number
by one in the state on the right, but not on the left; and the
scalar product of two different basis vectors is zero.
1 1 px2
H (x, px ) = m 02 x2 + .
2 2 m
2 d 2 1
2
( x ) + m 02 x2 (x) = E(x).
2m dx 2
1 2
H = (a2 + (a+ )2 ) m 02 K 2 + the rest.
4 mK 2
1 1 aa+ + a+ a
H = (aa+ + a+ a) m 02 K 2 + 2
= (aa+ + a+ a) 2 0 = 0 .
4 mK 4 2
212
This Hamiltonian is expressed in terms of creation and
annihilation operators. They act in the Fock space, which has
the basis 0 , 1 , 2 (We have only one kind of operator,
without any indices, so we need only one number to label a
quantum state.)
Now we can easily find the quantized energy levels of harmonic
oscillator. Remember, that the operator N = a+ a is the particle
number operator, and its expectation value in state |n is simply
n. Therefore the expectation value of the Hamiltonian in state
|n is
1 1 1
En = n 0 N + n = 0 n + = h 0 n + .
2 2 2
m 02 1
n H n = n x 2 n + n p x2 n = En > 0.
2 2m
that is, the integral d 3r *j (r ) l (r ) is zero unless j = l).
An example of such set of functions is a set of plane waves,
k (r ) = C exp (ik r ) (C is a number a normalization factor.)
A particle in a state k ( r ) has an exactly known momentum of
k and an indefinite position (because the probability of finding
2
a particle k ( r ) = C 2 is the same at any point of space).
Now let us put together field operators:
(r ) = (r) c ;
j
j j
+
(r ) = (r) c .
j
*
j
+
j
214
The operator + (r ) creates a particle at the point r, and the
operator (r ) annihilates it. The density of particles at this point
is given by the operator + (r ) (r ). These operators describe
the particle field, i.e. the situation when a particle can be found
anywhere in space.
()
n r =
() *j r c +j
( r )c
l l
j l
= ( r ) ( r ) c c
*
j l
+
j l
j l
(r ) ( r ) (r ) c c
2
= j
c +j c j + *
j l
+
j l
.
j =l j l
by U = d 3r (r ) e n (r ) , where n (r ) = * (r ) (r ) is the
probability density of finding this electron at the point r. Then the
potential energy operator for a many-electron system is
U = e d r (r ) ( r ) ( r ) .
3 +
2m
216
if you want to consider a single quantum particle (e.g. a single
electron) accurately enough. And, of course, it is extremely handy
when dealing with condensed matter systems, such as metals
or semiconductors, which contain huge numbers of quantum
particles. These questions we will consider in the next chapter.
{ , } = 0; { , } = 0; { , } = 0.
Therefore no matter what, the fermion wave functions cannot
become usual waves in the classical limit: such a wave would
have to have zero amplitude. So, e.g. a peace of metal, which
consists of fermions, must be solid to exist.6
? Fact-check
1 The principal quantum number determines
a the bound state energy of an electron
b the ionization energy
c the number of electrons in an atom
d the Rydberg constant
6 You may ask: What about atoms, which are bosons like helium-4?
The answer is, compound bosons are what is called hard core bosons in
the sense that they behave like bosons only until you try pushing them too
close together.
218
4 Spin characterizes
a an orbital motion of a quantum particle
b an orbital angular momentum of a quantum particle
c an intrinsic angular moment of a quantum particle
d the number of electrons in an atom
Dig deeper
Further reading I
R. Feynman, R. Leighton and m. sands, Feynman Lectures on
Physics. Basic Books, 2010 (there are earlier editions). Vol. iii,
chapter 4 (identical particles and the Pauli exclusion principle);
chapter 19 (the periodic table).
Further reading II. About the ideas and scientists.
J. Baggott, The Quantum Story: A History in 40 Moments, Oxford
university Press, 2011. a more modern history. Well worth
reading.
G. Gamow, Thirty Years that Shook Physics: The Story of Quantum
Theory, Dover Publications, 1985. a history written by an
eyewitness and one of the leading actors.
220
The virtual reality
8
Some of the most striking successes of quantum mechanics
come from its application to the systems that contain a huge
number of quantum particles. I mean condensed matter systems,
like metals, dielectrics, semiconductors, superconductors,
magnets and so on. These successes are also among those with
the greatest impact on technology, economy and society.
One of the reasons for this success is the same as that of
classical statistical mechanics. The behaviour of a large system
is often determined by averages average speeds, average
densities, etc and these are much easier to measure and calculate
than the properties of individual atoms or molecules. Finding
averages in large quantum systems adds to this an extra layer:
finding quantum expectation values. But we have already seen
in Chapter 3 that switching from the state vector to the density
matrix formalism can do this quite naturally.
Moreover, in this chapter we shall see that the description of
a system of many quantum particles, which strongly interact
with each other, can be often reduced to that of a system of
almost non-interacting quasiparticles with excellent results. It
is the behaviour of a modest collective of quantum particles
that are usually the hardest to describe and predict. There are
exceptions, however. If quantum correlations involving many
particles simultaneously are important, as is the case, for
example, in quantum computing, these simplifications do not
apply. For discussion of this situation please see Chapter 12.
221
Electrons in a jellium, charge
screening, electron gas and
FermiDirac distribution
We begin with the simplest quantum theory of metals. The
characteristic properties of metals their high electrical and
thermal conductivity, their metallic lustre, their strength, their
ability to be melted or hammered or pressed into a desired
shape without breaking apart in short, everything that makes
them useful is due to the fact that metals contain a large number
of free electrons.
If look at Mendeleevs Periodic Table, we see that the most
typical metals, the alkali metals (lithium, sodium, potassium,
etc.) are in the same column as hydrogen that is, they have
one unpaired electron in their outermost s-shell. This electron is
the closest to the top of the potential well and is therefore easily
lost by the atom. This, in particular, explains the extraordinary
chemical activity of the alkali metals.
Most other elements in the Table are metals that also lose
some of their outermost electrons pretty easily. As a result,
when atoms of these elements group together, these electrons
collectivize. That is, they are no longer bound to any specific
atom. Instead they can move anywhere in the piece of metal
and, e.g. can carry electric current. This is why they are called
conduction electrons. Other electrons, which occupy lower-
energy states, remain with their nuclei and form the positively
charged ions (Figure 8.1).
Though there is no individual bonding between ions and
conduction electrons, they are bound to each other collectively.
The two subsystems, positively charged ions and negatively
charged conduction electrons, are strongly attracted by the
electrostatic forces. This forms a specific metallic bonding.
The electrons, while free to move about and repel each other,
cannot leave the metal in any significant number, because then
they would be attracted back by the strong electric field of the
uncompensated ions. (A relatively tiny number of electrons can
be taken from the metal this is what happens in photoelectric
222
+Z
from the origin will be attracted not to the bare charge +Q,
but by the smaller screened charge QR = Q QR. The same
+
+Q
+ Q
+Q
224
happens if the probe charge is negative. Now the conduction
electrons will be repulsed and the positively charged jellium
will screen the probe charge, with the same result.
Taking one of the electrons as the probe charge, we see that
these very electrons also screen its interaction with other
electrons. It acts over a short distance only, and it is therefore
plausible that the conduction electrons in a metal behave as a
gas the electron gas. (Recall that classical statistical physics
treats a gas as a system of particles, which only interact when
colliding with each other.)
This is not quite a usual gas however. Electrons are fermions,
and their distribution function nF (E) (the number of electrons
with energy E) is strongly affected by the fact that they must
obey the Pauli principle. Therefore instead of the familiar
Boltzmann distribution we have the FermiDirac distribution:
1
nF (E) = E F
exp K T +1
B
1
2 (r ) = (r ).
TF
2
1 F
TF = .
e 6 n0
The Fermi energy F is the key characteristic of the electron gas and
is usually of the order of a few electron volts.
226
nF
0 E
EF
nF
~kBT
0 E
EF
0
x
A function (x), which is zero for all negative and one for all positive
values of x is called the Heaviside step function (because he was
p2
E(p) = .
2m
At absolute zero the maximal momentum an electron can have
is the Fermi momentum,
pF = 2mF .
pz
pF
py
px
228
state, we can represent all electron states on the same picture:
electrons fill the Fermi sphere.
N 1 4 4 3
= 2 3 pF3 = 2 3 (2mF ) 2 .
h 3 3h
The factor of two is due to the electron spin: two electrons with
opposite spins can occupy each state.
In a more general case (for example, at a finite temperature) the
electron density is given by the expression
N 2
=
h3 d p n (E( p)).
3
F
d (E )
D (E ) = .
dE
3 3 1
230
material, e.g. nickel) produces an interference pattern. The
same happens when a conduction electron moves through the
lattice of positive ions, which are arranged in a regular way: it is
scattered. As the result, the electron wave function will produce
an interference pattern, and the expectation value of momentum
and energy will be affected compared to what one would expect
for an electron wave in the free space. This is not surprising,
after all, since the electrons are acted upon by the ions. The ions
themselves are much heavier than electrons and the reciprocal
action of electrons on the ions can be usually neglected.
Since conduction electrons in a metal (or in a semiconductor)
usually have properties very different from those of a free
electron in space, they are called quasiparticles.
Conduction electrons are not the only quasiparticles in solids
(Figure 8.5). There are such quasiparticles as phonons, polaritons,
magnons, spinons, polarons, etc. Unlike conduction electrons, some
of them cannot be brought to the outside. They only exist within,
like the units in a MMPORG (a massively multiplayer online role-
playing game). For example, a phonon is a quantum of sound. But
in the solid it has all the properties of a quantum particle, with its
own energy, momentum, dispersion relation and quantum statistics
(phonons, by the way, are bosons, not fermions).
2 2
(r ) + U (r )(r ) = E (r )
2m
p (r ) = up (r )e i p.r / .
232
ipx
the form of a Bloch function, p (x) = up (x)e . Note that this
wave function is not itself periodic:
Band gap
Band
p
3G/2 G/2 G/2 3G/2
Figure 8.6 electron in an empty 1D lattice: bands and band gaps
1 They are called the inverse lattice vectors, but you do not have to
remember this.
234
derivative dE
dp
= 0. Therefore a gap (the band gap) opens in the
spectrum: some values of energy for a conduction electron are
not allowed.
In a real 3D lattice the resulting band structure can be very
intricate, but the principle is the same.
2 1
p
G/2 G/2
236
Note that in the presence of electric field electrons and holes
will move in the opposite directions, but carry the electric
current in the same direction.
In a metal the surface in the quasimomentum space, determined
by the condition E(p) = F, is called the Fermi surface. In the
jellium model it was simply a sphere of radius pF. In real
metals it can be much more interesting (Figure 8.8). The shape
of the Fermi surface and the band structure to a large degree
determine what kind of quasiparticles exist in the system.
238
Fortunately, there are ways around this obstacle. Many
properties of a physical system can be obtained from its
response to an external perturbation like knocking on a wall
to discover a hidden cavity. One way of perturbing a system of
quantum particles is to add a particle at some point, wait and
see what happens, and then take it away at some other point.
Alternatively, we can instead remove a particle, wait, and then
replace it (Figure 8.10).
In the first case, the quantum state vector of the system will be
= (r , t ) + (r , t)
= iG+ (r , t ; r , t).
Perturbation theory
Very few problems in physics have an exact solution. This is not
to say that all such problems have already been discovered and
solved, but it is far more probable that any new problem does
not have an exact solution. Numerical methods are helpful and
powerful, but despite the ever growing computing power there
are many problems they cannot solve now and will never solve
because of the sheer scale of required computations (see the
3 For example, you can check that the average particle density in the system
is given by iG(r , t ; r , t). .
240
discussion in Chapter 12). Moreover, numerical computation
is akin to experimental physics: it does not necessarily give
insight into the key features of the effects that are being
modelled (wherefore the term computer experiment). Numerical
computations are also vulnerable to flaws in the underlying
computational models, and without some guidelines obtained in
other ways it is hard to distinguish model-dependent artefacts
from real effects.4 It is always good to have some guidelines
against which to measure the numeric, and which can provide
insight into the mechanism of an effect and its likely dependence
on the model parameters. In short, physicists still prefer to have
analytical solutions, even in the situation when exact solutions
are not known or do not exist.
The way of dealing with this situation was first found in
Newtons time, and it has served as the mainstay of theoretical
physics ever since. I am talking about the perturbation theory.
The idea of the approach is very simple. If our problem does
not have an exact solution, then we will look for a similar
problem that has. We then will modify that solution to fit
our problem. There are systematic ways of doing it, and
they are very useful. Laplace very successfully applied it to
the calculation of planetary motions in his famous work on
celestial mechanics, published long before the appearance of
computers.
In quantum mechanics, the perturbation theory works if the
Hamiltonian of the system can be presented as a sum of the
main term and the perturbation:
H = H 0 + H 1.
F (x ) = kx k x 3 .
x (t ) = x 0 (t ) + x 1(t ) + 2 x 2 (t ) +
Since 1 the corrections will be small, and only the main terms
need to be kept. Substituting this expansion in our equation, you
will find:
d 2x1
m = kx 1(t ) k(x 0 (t ))3 +
dt 2
d 2x1
m = kx 1(t ) k(A sin t )3 .
dt 2
This equation can be solved explicitly and yield an approximate
solution for the cubic oscillator. If better accuracy is required,
we look at the terms with higher powers of . There will be some
subtle problems with this expansion, but they also can be taken
care of in a similarly systematic way.
242
Here I explicitly include a small parameter 1. The state vector
of the system will be sought in the form of a power series:
= 0 + 1 + 2 2 + + n n +
d
0 i 0 = H0 0
dt
d
1 i 1 = H 0 1 + H 1 0
dt
d
2 i 2 = H 0 2 + H 1 1
dt
244
interactions between the particles. We will again consider this
against the background of the jellium model of a metal.
First, there is an electrostatic interaction between electrons. The
corresponding perturbation term in the Hamiltonian is
H c = d r d r ( r , t ) ( r , t )V ( r r ) + ( r , t ) ( r , t ).
3 3 +
H ph = g d 3r + (r , t) (r ) (r , t).
246
B C A
(a)
B C
D
(b) C A
(c)
(d)
Figure 8.12 second order Feynman diagrams for the electrons Greens
function
+ + +
248
(the star here stands for a certain set of mathematical
operations, like integrations over positions and momenta of the
virtual particles). The solution is
1
Veff V.
1 V 0
= + + + +
+ +
= + +=
+ + + =
250
? Fact-check
1 A metal is held together by
a strong forces
b electromagnetic forces
c electrostatic attraction between conduction electrons and
lattice ions
d gravity
7 A Greens function is
a an expectation value of a product of two field operators
b a scalar product of two quantum state vectors
c a measure of a many-particle quantum systems response
to an external perturbation
d the quantum state of a system of quasiparticles
8 A Feynman diagram is
a a graphic representation of a quantum system
b a graphic representation of a state of a quantum system
c a graphic representation of one term in the perturbation
expansion of a Greens function
d a graphic representation of a series of terms in the
perturbation expansion of a Greens function
9 A virtual particle
a cannot have momentum, energy or electric charge
b is represented by an internal line in a Feynman diagram
c is a quantum particle, which exists for a limited time
d does not satisfy the relation between the momentum and
energy
252
Dig deeper
t. Lancaster and s. J. Blundell, Quantum Field Theory for the
Gifted Amateur. Oxford university Press, 2014.
254
Spotlight: Richard Feynmann
Richard Feynman (19181988) was an American theoretical
physicist. He was awarded the Nobel Prize in Physics (in 1965,
together with Shin-Itiro Tomonaga and Julian Schwinger) for the
development of quantum electrodynamics. Feynman participated
in the Manhattan Project, developed the path integral approach to
quantum mechanics, and obtained a number of other important
results. In particular, he finished the theory of superfluid helium
built by Lev Landau, and his diagram technique helped elucidate
the concept of quasiparticles introduced by Landau.
Feynman and Landau were much alike in many respects. Both had
flamboyant personalities, both were brilliant, both achieved great
results across the spectrum of theoretical physics, and both left
behind great textbooks.
The Feynman Lectures on Physics remains one of the best
undergraduate courses on the subject. Characteristically, for each
three copies of English editions of this book ever sold, its Russian
translation sold two.
Besides that, Feynman published higher-level physics textbooks,
popular books on physics and other subjects. He played a key role
in finding the reason behind the Challenger disaster. Feynman
predicted and encouraged the development of nanotechnology and
what is more important for our purpose quantum computing.
And yes, Feynman played bongo drums. Landau did not.
Variational approach to
classical mechanics
The usual approach to classical mechanics is via the three
Newtons laws. For example, for a system of N material
points massive particles of zero size we need to solve the
N vector (or 3N scalar one for each Cartesian component)
equations of motion
d 2 ri d r
mi 2
= Fi r ,
dt dt
l = x2 (t) + y2 (t),
256
x
mg
y
Figure 9.2 a pendulum
is fixed. There are two forces acting on the bob, gravity, mg,
and the reaction of the rod, f. The Newtons equations are (it is
convenient to direct the y axis downwards):
d2x x(t)
m = f x = f (t) ;
dt 2 l
d2y y(t)
m 2
= f y + mg = f (t) + mg.
dt l
258
and define the action as its integral along a given path q j (t), j =
1 N , which the system follows between the initial and final
times:
tf
S[q(t)] =
dt L({q, q }).
ti
S = 0.
Then, if we slightly shift the path, that is, replace q (t) with q (t) + q(t)
(an infinitesimally small function q(t), which is zero at t = ti and
t = tf , is called a variation) (Figure 9.3), the action will not change:
S[q (t)] = S[q (t) + q(t)] S[q (t)] = 0.
q(t)
ti
Figure 9.3 Variation of a path
260
Key idea: newtonian and Lagrangian differences
2
md r =F
dt 2
S = 0
d L L
= 0.
dt q j q j
m(l )2
L( , ) = mgl (1 cos ).
2
2 g
The Lagrange equation, dtd (ml 2 ) + mgl sin = 0, or d 2 = l sin ,
dt
is the usual equation for a pendulum, which could be obtained
without using the Lagrange approach. Still, in this way it is
obtained effortlessly, and the advantages of the method become
evident for more complex systems.
Practically all of theoretical physics (including electrodynamics,
special and general relativity, hydrodynamics, etc.) can be
presented starting from variational principles. For example,
Landau and Lifshitz take this approach in their multivolume
Course of Theoretical Physics, which helped educate generations
of physicists worldwide.
262
Variational approach to quantum
mechanics
Let us now return to the literal interpretation of the variational
principle. Suppose we could calculate action for all possible
paths, S[q(t)], compare them and choose the extremal one. How
could we simplify this task?
To begin with, we can calculate for each path qj(t) a fast
oscillating function, for example, cos ( S[ q j (t )]), where is
a large number. A small change in q(t) will produce a small
change in S[q(t)], but if the factor is large enough, the cosine
can change all the way between minus one and one. Now we
simply add together all the contributions from different paths.
In the resulting sum, cos ( S[q j (t)]), the contributions from
j
different paths will mostly cancel out (see Figure 9.4): for each
path qj(t) there will be a nearby path q j (t) + (t) (the small
number is of the order of 1 ) such that
cos( S[q j (t) + (t)]) cos( S[q j (t)] + ) = cos( S[q j (t)]).
Figure 9.4 contributions from different paths. the solid line is the
surviving extremal path, which corresponds to the classical motion.
As you can see, it oscillates very fast everywhere, except near the
point x = 3, where the argument of the cosine has a minimum.
2 Similar reasoning explains why light, being a wave, propagates along straight
lines. There the role of the small parameter plays the light wavelength : the
effects of light interference and diffraction appear on the scale comparable to .
264
But now we can turn around and say: if we start our reasoning
from the sum like j cos(S[q j (t)]) and consider which term
contributes the most, we will inevitably arrive in the end at
the principle of stationary action. The principle, which
looked like some mysterious revelation from the top and
caused so many philosophical speculations, becomes simply a
mathematical necessity.
The question of what is the large number is easily answered:
it is 1 . Indeed, then the argument of the oscillating function is
a dimensionless number, S = 2 S , that is, 2 times the action
of physical system in the units of action quantum. An extra
quantum makes the oscillating function run the full circle.
This was the inspiration for the approach suggested by Dirac in
1932 and realized by Feynman a quarter of a century later.
The oscillating function used is exp( iS[q j (t)] ) = cos( S[q j (t)] )
+ i sin( S[q j (t)] ). The question is: how do we calculate the
contributions from different paths, and contributions to what?
ti tf
t
Figure 9.5 slicing the interval and trajectories
266
The action integral is approximated by a sum:
tf N 1
S[q(t)] = L(q, q ) dt L t.j
ti j =1
N 1
i
i
exp S[q(t)] exp Lj t .
j =1
of these paths, S[q(t)] = L(q, q) As a more accurate analysis
demonstrates, one should include certain normalization factors
C t in each of the integrals over qj before taking the limit
t0, so this is not such an easy operation as a transition from
a sum to an ordinary integral. Nevertheless it is quite doable.
The result is
Cdq1
Cdq2
CdqN 1 i N 1
path intergal = lim
t
N
t t
t
exp
L t
j =1
j
qf (t f )
i
=
qi (ti )
Dq exp S[q(t)]
q f (t f )
tf
i
i f =
qi (t i )
Dq exp
ti
L(q, q ) dt .
One can derive from the path integral formulation all the
equations of quantum mechanics we have encountered before,
even in the case of many-body systems. In the latter case we can
define Feynman integrals over fields. Instead of the Lagrangian
we will have the Lagrangian density (which will be integrated
over all of space to obtain the Lagrangian) and, instead of Dq,
we may encounter D D (pointing at the integration over all
possible configurations of fields).
We know that quantum fields can be bosonic or fermionic. With
bosons there is no problem the corresponding fields in path
integrals are usual functions, with real or, at worst, complex
values. With fermions we would seem to get into trouble; there
are no usual numbers, which would anticommute. Fortunately,
Felix Berezin discovered the method of calculating it using
special numbers (so-called Grassmann variables). We will not
discuss how it works (though this is a fascinating and rather
simple topic), but will just assure you that it does work.
Besides the philosophical attractiveness (the explanation
of where the variational principles of physics are coming
from), there are other advantages of using the path integral
formulation. The main one is that the path integrals include the
Lagrangian, and not the Hamiltonian.
This is not an important point as long as we are happy with
the non-relativistic world, where space and time do not mix.
268
The Schrdinger equation, the Heisenberg equation, the von
Neumann equation everywhere time is a variable apart. For
example, the Schrdinger equation contains the first order time
derivative, and the second order position derivative.
The problems appear when one can no longer neglect the
relativistic effects. As we have known since 1907, courtesy of
Hermann Minkowski, what seems to us three-dimensional
space + one-dimensional time is actually the four-dimensional
spacetime. This means, in particular, that coordinates and
time get mixed, if we deal with objects moving with relativistic
speeds (that is, with speeds comparable to the speed of light in
vacuum). Therefore a quantum theory of such objects can no
longer treat them unequally, like in the Schrdinger equation.
It must be what is called Lorentz invariant.
S [ , ] = dt d rL ( (r , t ) , (r , t )) = d X L ( (X ) , ( X )).
ti
3 4
does not make any explicit difference between space and time,
and it can be used as the basis for a relativistic quantum field
theory. Of course, you still do not know how to calculate such
an integral, but then, we are only learning here about quantum
theory. If you decide to study quantum theory at a rather
advanced level, you will know how.
270
Case study: Vector potential
B
B d s = A dl.
S
272
electromagnetic field has its own momentum, and this is the
simplest (and correct) way to account for it.
But how do we include the magnetic field in the Lagrangian
function? Recall that while the Hamiltonian is the sum of
kinetic and potential energy terms, the Lagrangian function is
their difference: H = T + U, L = T U. Therefore all we need is
to rewrite the kinetic energy term through velocities rather than
through momenta.
Take a single charged particle of mass M. Then its kinetic energy
in the magnetic field is
(p A)
2
Q
c p2 Q Q2
T= = Ap+ A2 .
2M 2M Mc 2Mc2
The last term relates to the energy of the magnetic field itself
and is of no concern to us. The first term is just the kinetic
energy of the particle without any magnetic field. All the
particle-field interaction is contained in the second term, which
we can rewrite as
Q
A v,
c
p
where v = x = M is the velocity of the particle. Therefore in the
presence of the magnetic field the Lagrangian of a charged
particle becomes
Q Q dx
L ( x , x , A ) = L ( x , x ; A = 0 ) A x = L ( x , x ; A = 0 ) A .
c c dt
i ie
C1 exp Spath1 exp
c
path1
A dx
i ie
+ C2 exp Spath2 exp
c
path2
A d x .
Here Spath 1 and Spath 2 are actions in the absence of the magnetic
field, and we choose for simplicity a pair with Spath1 = Spath2 = S.
274
B
e
A
ie ie
= C1C2 exp
c
path 1
A dx
c
A dx
path 2
ie ie
A d x .
+ exp
c A dx +
c
path 1 path 2
e
2
AB
path1, path2
cos
c
path 1
A dx
path 2
A d x .
This does not seem a very useful expression, with all its
integrals. But let us reverse the path 2: instead of integrating
the vector potential from A to B we integrate it from B to A
(Figure 9.6). The integral will simply change sign (since we go
along the same path, but in the opposite direction). But then the
sum of two integrals becomes one integral along the closed path
from A to B and back to A, which goes around the solenoid:
e e e
c
path 1
A dx
path 2
A dx =
c
A dx = c .
On the last step we have used the defining property of the
vector potential. Now we see that each pair of paths with
2
Spath1 = Spath2 = S will add to the probability A B a term that
e
contains the same oscillating factor cos c . There will be also
contributions from other terms, but they will not completely
erase these oscillations. Therefore the interference pattern on
the screen will change, if the magnetic field in the solenoid
changes even though the electrons never go into the regions
of space, where the magnetic field B is nonzero! The vector
potential of the field is nonzero there and this is enough, despite
the fact that there is no force acting on the electron. A classical
charged particle would not notice anything.
This beautiful and counterintuitive result is purely quantum
mechanical, and it was observed in a number of experiments.
For example, if you measure the electrical conductance of a
small (micron-sized) metal ring, it will oscillate as you change
the magnetic field penetrating the ring (conductance is, after all,
determined by the probability that an electron can get from the
one electrode to another).
276
The oscillating factor cos e
c
,
. can be rewritten as cos 2 0
to stress that the period of oscillations is
= hc .
0
e
ei .
Of course, the phase factor does not affect the state vector
itself after all, the probability is given by its modulus squared.
But if there is interference between two vectors with different
Berry phases, as in the case of the AharonovBohm effect, there
can be observable consequences.
Now we can ask whether such effects could happen to large
systems. Suppose a big particle can go from A to B along two
mirror-symmetric ways around the solenoid, with exactly
the same actions S, and the only difference due to the vector
potential. Should not we see the AharonovBohm effect? The
particle can be big even though the action S is much, much
greater, than , it cancels out. The same reasoning should apply
? Fact-check
1 In classical mechanics, a physical motion realizes
a a maximum of the Lagrangian
b a minimum of the Lagrangian
c an extremum of the action
d a minimum of energy
278
5 The AharonovBohm effect demonstrates that
a the phase of a charged quantum particle depends on the
magnetic field
b a charged quantum particle interacts with the magnetic
field
c a charged quantum particle interacts with the vector
potential
d a quantum particle can travel through a solenoid
280
Order! Order!
10
Quantum systems are in principle more complex than the
classical ones, in particular those quantum systems that contain
many quantum particles. They possess specific, strange and
outright weird properties, such as entanglement, and their
behaviour is, in principle, impossible to model in all detail using
classical computers (the fact pointed out by Feynman, which
initiated the search for quantum computers). Nevertheless,
as we have seen in Chapter 8, there are situations when an
approximate treatment of quantum many-body systems
are not only possible but also pretty straightforward and
accurate. Fortunately, these situations include a large number
of scientifically interesting and practically very important
cases. We will start with the workhorse of class demonstration
experiments: magnetism.
281
Classical magnets, second order
phase transitions and the order
parameter
Magnetism is one of the oldest subjects of scientific investigation.
In the classic treatise On the Magnet, Magnetic Bodies, and the
Great Magnet of the Earth published in 1600, William Hilbert
(15441603) had already stated that magnetic attraction is a
separate phenomenon, not to be mixed up with the force exerted
by an electrified piece of amber. Since then, the investigation of
magnetism became one of the most important areas of physics.
Many results, methods and approaches migrated from there to
far removed areas of physics, to cosmology, biology and even
economics.
A ferromagnet what is colloquially called a magnet in
equilibrium has a magnetic moment M, which disappears above
certain temperature TC (the Curie point). This is because certain
atoms in this system have microscopic magnetic moments of
their own, and their interaction with each other tends to align
them (Figure 10.1). At high enough temperatures the thermal
fluctuations rattle the system so much that this alignment is
destroyed, and the magnet loses its magnetization. If we cool the
magnet down, the magnetization is restored. (However we will
not go any deeper into some important details here, e.g. that a
ferromagnet is usually split into regions with different directions
of magnetization (domains) and will consider a single domain.)
This process is called phase transition.
282
Classical statistical physics has a number of ways of describing
magnets, which successfully reproduce their key features. In
a simple theoretical model of a classical magnet we assume
that each atom in a crystal lattice has a magnetic moment sj
(j labels the lattice site, and we neglect all the atoms that do
not possess a magnetic moment of their own). Here is the
absolute value of the atomic magnetic moment (assuming for
simplicity that they are all the same), and sj is a vector of unit
length, which shows its direction. The Hamiltonian function (i.e.
the energy) of such a magnet is
H= J jk s j sk h s .
j
jk j
284
Here Jjk is the coupling between two magnetic moments
(which has the dimension of energy), and h is the external
magnetic field. In equilibrium, physical systems have the
lowest possible energy.1 The signs are chosen to ensure
that, first, in the presence of the magnetic field each atomic
moment tends to align with it, as a compass needle: the
second term in the Hamiltonian will correspond to the lowest
energy, when for each site h.sj>0. Second, if all Jjk are non-
negative, the first term in the Hamiltonian function will give
the lowest energy, when all the dot products are positive,
sj sk >0. This is the case of ferromagnetic coupling, as all the
magnetic moments will want to point in the same direction.
On the contrary, if the Jjk were negative, the pairs of coupled
moments would want to point in the opposite directions
(antiferromagnetic coupling).2 If the couplings can have
either sign, especially at random, the situation becomes really
interesting and often intractable. But here we are concerned
with the ferromagnetic coupling only.
Let us switch off the external magnetic field (put h = 0) and look
at the average magnetic moment of the system of N magnetic
atoms. Here we can use essentially the same approach as in
Chapter 1: consider the ensemble of systems in a configuration
space. There will be 2N coordinates, since the direction of a unit
vector sj is determined by two numbers (for example, the polar
angle j changing from zero to radians, and the azimuthal
angle j changing from zero to 2, Figure 10.2).
The ensemble is represented by a swarm of M 1 points, each
corresponding to a possible state of the system. The average
magnetization on the site number j of the Ising magnet is
s(k) . (The upper index labels the point in
M
s j = lim M 1
M K =1 j
I II
286
freezing point until it melts. This is because different phases have
different energies per molecule, which must be supplied before,
e.g. an H2O molecule in liquid water can evaporate.
The appearance and disappearance of magnetization at the Curie
point is different. If the external magnetic field is zero, this is a
second order phase transition. It means, in particular, that the
transition happens all at once. The reason is that the energies of
both phases at this point are the same. What is different is their
symmetry.
3 The fact that a sphere (in three dimensions) and a circle (in two
dimensions) are the most symmetric objects was recognized early on. This is
why ancient philosophers envisaged a perfectly spherical Moon and perfectly
spherical planets revolving around the Earth along circular trajectories.
288
in your pocket.) Nevertheless, if we assume that such quantum
objects as atomic magnetic moments exist, we can for a while
treat them as classical variables.
This is what Landau did in his theory of second order phase
transitions. He developed it when working on the theory of liquid
helium, and it proved to have a much wider applicability. Since
then more accurate theories of phase transitions were developed,
but it remains a very powerful, insightful and useful tool.
His idea was startlingly simple. It is known from classical
thermodynamics and statistical mechanics that in equilibrium
a physical system is in a state with the lowest free energy. Free
energy is a certain thermodynamic function, which at absolute
zero coincides with the total energy of the system.4 Landau
guessed that near the phase transition point (critical temperature,
Tc ) it would have the following form (Figure10.4):
F() = a(T Tc ) 2 + b 4 .
F F
T > Tc T < Tc
Figure 10.4 Free energy above and below the second order phase transition
temperature
dF
( Tc ) + 4b 3 .
= 2aT
d
Check that
if T > Tc, the only minimum F() is at = 0.
If T < Tc, then F() has a maximum at = 0 and a minimum at
a
= (T T ).
2b c
290
Spontaneous symmetry breaking occurs when the symmetry of a
solution is lower than the symmetry of the problem.
What does this actually mean? In our case the problem the
function F() is symmetric: it does not depend on the sign
of , only on its absolute value. The solution, on the other
a a
hand, is either + 2b
(T Tc ) or 2b
(T Tc ), so that has a
definite sign. The choice between them is made at random any
infinitesimally small perturbation at the transition point will
push the system towards the one or the other.
A standard textbook illustration of spontaneous symmetry
breaking is a sharp pencil balancing on its tip on a smooth
horizontal surface: it can fall in any direction with equal
probability, so the problem is symmetric, but the slightest whiff
of air will bring it down, and it will end up pointing in one
definite direction. A less standard one involves a mosquito and
two heraldic animals carrying a heavy object.
Figure 10.5 spontaneous symmetry breaking (Zagoskin 2014: 167, Fig. 4.5)
A similar situation arises in the Heisenberg magnet. There
we can take for the order parameter the absolute value of the
magnetization, M = |M|, and write
F ( M ) = a (T Tc ) M 2 + bM 4 .
H = J jk s j sk h s ,
j
jk j
Superconductors, superconducting
phase transition and the
superconducting order parameter
The time span between the discovery of superconductivity
in 1911 and its explanation in 1957 may seem excessive,
but only to a public accustomed to a steady and very fast
progress of physics over the last hundred years. When it
was discovered, quantum theory was still in its infancy, and
superconductivity is a macroscopic quantum phenomenon.
The theoretical description of superconductivity did not
appear all of a sudden in a finished form: there were a
number of intermediate steps, with stop-gap assumptions and
plausible guesses, and remarkably most of this guesswork
was eventually shown to follow from a consistent microscopic
theory.
292
Spotlight: heike Kamerlingh Onnes
Superconductivity was discovered by Heike Kamerlingh Onnes
(18531926) in the course of his experiments on the properties of
matter near the absolute zero of temperature. These experiments
were only made possible by his development and refinement of
cryogenic technology. He received his Nobel Prize in Physics
(1913) for this research, and in particular, for the liquefaction of
helium. This was a well-deserved award: all cryogenics is based
on the use of liquid helium (the common helium-4 and the rare
isotope, helium-3), and costs of helium loom big in most physics
departments research budgets.
f f f
f (r ) = e x + ey + ez .
x y z
5 This applies if it takes place with zero external magnetic fields. If the
system is initially placed in a magnetic field, as in Figure 10.6, this will be a
first-order phase transition. The role of latent heat is played by the energy
required to expel the magnetic field from the volume of space occupied by
the superconductor.
294
The essence of this theory is contained in the expression for the
free energy density6 of a superconductor:
B2 ( r )
2
2 4 1 q
F(( r )) F0 = a(T Tc ) ( r ) + b ( r ) +
A ( r ) + .
2m i c 8
Here F0 is the free energy density of the system in its normal state.
First, the right-hand side of this equation contains familiar
terms, like in a magnet, only with the magnetic moment
replaced by some complex function (r). This should be the
order parameter in our system below the superconducting
transition temperature Tc. The last term is (as you probably
know from the theory of electromagnetism) the energy density
of the magnetic field B(r).7 But what is the third term?
( )
2
2
d 3r (r ) 1 (r ) = 1 d 3r i (r ) . Then, in Chapter 9
2m i 2m
we saw that in the presence of magnetic field, described by a
vector potential A, the quantum momentum operator p is replaced
q
by p c A for a particle with the electric charge q. Therefore
the density of the kinetic energy in the system should indeed be
( )
2
1 q
c A (r ) .
2m i
6 Free energy per unit volume. In order to find the free energy of the
system, one must integrate this expression over all of space.
7 Again, we are using Gaussian units your textbook is probably using
the SI, so do not worry about the coefficients like 1/8 or 1/c.
( r ) (r )e i (r ) ,
separating its absolute value and its phase, and . This makes
sense. The absolute value determines the very existence of
superconductivity: it is zero above Tc and nonzero below it. In
the absence of external magnetic field, when both B and A are
zero, we can assume that is the same across the system, and
therefore the kinetic term in the free energy (which contains
the derivative of ) will be zero as well. Then what remains is
the same as in a magnet, with playing the role of the absolute
value of the magnetic moment, M, and nothing depends on the
superconducting phase (as nothing depended on the direction
of the magnetic moment).
Here we will again run into the same problem as with a magnet:
since nothing depends on , it can take any value, and the
ensemble average of the complex order parameter would be
zero. Again, this is resolved by using quasiaverages.
The phase of the order parameter of an isolated superconductor
has no physical meaning much as the overall phase of a wave
function. But its gradient the magnitude and direction of its
change in space plays an important role. It determines the
magnitude and direction of the supercurrent the very current
that flows without resistance and that attracted attention of
Kamerlingh Onnes in the first place.
296
Key idea: Velocity operators
p
The velocity operator is = 1 . The expectation value of
m m i
velocity at the point r is then (r ) 1 (r ) . Assuming
m i
that (r ) ei and is a constant, we see that the velocity
q
is , and the electric current density j . For
m m
example, j q d .
x
m dx
298
1
nB ( E; , T ) =
exp ( )1
E
KBT
n (E; , T ) = N.
E
B
can live with. Therefore e KBT 1 > 0,that is, e KBT > 1 for any
energy E 0. Therefore is always negative (compare this to
the always positive Fermi energy).
In a large system, where the energy levels are very dense, the
sum E nB (E; , T ) can be usually rewritten as an integral,
0
(E)nB (E; , T ) dE. Here (E) is the density of states in a
small energy interval E there are (E)E energy levels. Here
usually is not a figure of speech: the condition
300
in the physics of strong magnetic fields before returning to the
Soviet Union to discover the superfluidity of helium. He was
Rutherfords student and friend, giving Rutherford the pet name
The Crocodile (which stuck, and Rutherford approved of).
Kapitsa commissioned the relief of a crocodile by Eric Gill on the
wall of the Mond Laboratory in Cambridge, which was built by the
Royal Society for the research of strong magnetic fields.
cp c p = 0 cp c p 0 0.
0
302
nonzero anomalous averages in the ground state of a system is
the signature of the condensate.
( r ) (r ) (r ) .
0
by adding to the Hamiltonian the term d 3r g + +
(g is a positive constant), and further approximated by writing
d r g d r g
3 +
+
3 +
+
0
+ d 3 r g + +
0
=
d r ( r ) ( r ) .
3
+
+
304
An important note: since electromagnetic fields cannot penetrate
inside a superconductor, and electric currents produce magnetic
fields, the supercurrent must flow in a thin layer near the
surface of a superconductor. (The thickness of this layer is
called the superconducting penetration depth). In the bulk of a
superconductor there are no currents and no fields.
Josephson effect*
Now let us consider a beautiful and instructive phenomenon,
which has many applications in science and technology. This is
the Josephson effect.
We have seen that the GinzburgLandau theory was put on
a solid microscopic foundation. Its superconducting order
parameter (r) can indeed be considered as a macroscopic
wave function of the superconducting condensate of Cooper
pairs. The effective mass m* is twice the mass of an electron (a
conduction electron, not the one in vacuum), and the charge
q =2e is twice the electron charge. What is more to the point,
it is the condensate density, that is, the anomalous average,
0, and not the energy gap, g 0, that determines the
superconductivity. The coupling strength g, and with it the energy
gap, can be zero in some part of the system, but there will still be
superconductivity there.
ns1 1
ns2 2
ns1 e i1
= .
ns 2 e i2
0 K
H = ,
K 0
306
The Schrdinger equation will be then
i = i
{ n s1
2 ns1 }
+ i1 e i1
= K
ns 2 e i2
.
t
{ n s2
2 ns2 }
+ i2 e i2
ns1 e i1
ns1 2K n 2K
= ns1ns2 e i(2 1) ; s2 = ns1ns2 e i (1 2 ) .
t i t i
Let us subtract the one from the other and recall that
sin x = (e ix e ix ) 2i. We will find, that
4K
(ns2 ns1) = ns1ns2 sin (1 2 ) .
t
? Fact-check
1 The adjacent atomic magnetic moments tend to point in the
same direction, if the coupling is
a antiferromagnetic
b ferromagnetic
c diamagnetic
d paramagnetic
308
4 Spontaneous symmetry breaking means that
a the symmetry of the system is randomly destroyed
b the system undergoes a phase transition
c the symmetry of the systems state is lower than the
symmetry of the equations which govern its behaviour
d the system becomes more symmetric
6 Free energy
a is the energy that can be freely extracted from a physical
system
b is a quantum mechanical operator of a physical system
c coincides with the energy of a system at absolute zero
d is a thermodynamic function, which is at a minimum when
the system is in equilibrium
Dig deeper
F. Reif, Statistical Physics (Berkeley Physics course, Vol. 5).
mcGraw-hill Book company, 1967.
For an introduction to superconductivity (its science and
scientists) see
J. matricon and G. Waysand, The Cold Wars: A History of
Superconductivity. Rutgers university Press, 2003.
V. V. schmidt, The Physics of Superconductors: Introduction to
Fundamentals and Applications. springer, 1997, 2010.
310
Curiouser and
11
curiouser
One of the main objections against Newtons theory of gravity
was that it introduced action at a distance: in order for the law
of inverse squares to work, any two gravitating masses had to
know the positions of each other at any instant, no matter how
large a distance separated them. The adherents of Descartes
physics, which dominated European natural philosophy before
Newton and did not go down without a fight, pointed that out
as the cardinal flaw of the new fangled theory. To them action
at a distance was inconceivable, and gravity was supposed to
be the result of interaction of bodies with the aether an all-
pervading medium. The vortices in aether were supposed to be
responsible for the formation of planetary orbits and for the fall
of heavy bodies.
Unfortunately, Cartesian physics itself had a cardinal flaw:
paraphrasing Laplace, it could explain anything and predict
nothing. Newtonian mechanics, despite (or rather thanks
to) its unphysical action at a distance could and did make
quantitative predictions, and that sealed the matter for the time.
The Coulomb law was another version of action at a distance.
Only when the attempts to present electrodynamics in the same
way ran into serious problems in the mid-19th century while
Maxwells theory was successful, the action at a distance theories
started going out of fashion. When the general relativity took
over from the Newtonian gravity, it seemed that local theories
would rule forever. And then came quantum mechanics
311
Einstein-Podolsky-Rosen paradox
The intrinsic randomness of quantum theory was, of course,
an irritant for the physicists accustomed to the (theoretically
strict) Laplacian determinism of classical mechanics. But
this irritation was mitigated by the experience of classical
statistical mechanics and the practical impossibility to follow
on this determinism for a macroscopic system (for example,
a roomful, or even a fraction of the cubic millimetre, of
gas). It seemed that quantum mechanics could well be
an approximation, which can be derived from a more
complete theory. Such a theory would be deterministic
and operate with some quantities (hidden variables),
which for some reason only manifest themselves through
quantum mechanical observables. There were precedents. For
example, equations of the theory of elasticity were known
and accurately described the properties of solids long before
the atomic structure of matter was established, not to speak
about the observation of atoms themselves.
312
depends on the particle being able to go through both slits. It
is clear in Feynmans path integrals formulation of quantum
mechanics. It is obvious when you look at the Aharonov
Bohm effect, when the behaviour of a particle depends on the
magnetic field, though the particle cannot even in theory travel
through the region where this field is present. But long before
Feynmans and AharonovBohms work, in 1935, Einstein,
Podolsky and Rosen invented a Gedankenexperiment (the EPR
experiment), which, as they thought, established that quantum
mechanics is incomplete that is, must eventually reduce
to some theory of hidden variables. Anything else would be
inconceivable.
O
A B
We can choose the coordinates in such a way that the up state (along
the axis Oz) is the state ( ) , and the down state is ( 10 ) . Then the
1
0
expectation value of S x in the state EPR = is (figure out
2
(
EPR Sx EPR = 41 xA 0B + 0A xB . )
Take the xA 0B -term first.
1 1 1
xA 0B = xA 0B = xA 0B xA 0B
4 4 4
1 1 1
xA 0B + xA 0B = xA 0B
4 4 4
1 1 1
xA 0B xA 0B + xA 0B .
4 4 4
We have here explicitly written that xA only acts on the first arrow
(particle A), and 0B only on the second (particle B). The matrix 0B is a
unit matrix. It does not change the state it acts upon: for example,
0 = 1 0 1 = 1 = .
0 1 0 0
314
The remaining terms are 1
4
xA 0B + xA 0B ,
You can check that the 0A xB -term will also produce a zero.
Repeating this with Sy and Sz will improve your quantum
mechanical intuition and build character.
Let us align both detectors with the axis Oz. Suppose the
apparatus B is a little farther from O, or that the particle B flies
a little slower, so that the detector A always fires first.
According to the rules of quantum mechanics, during
measurement the quantum state of the system is randomly
projected on one of the eigenstates of the operator being
measured, and the measured value will be the corresponding
eigenstate. The operator measured by the detector A is (we
measure spins in the convenient units of ) SzA = 12 zA 0B .
The 0B -term simply means that the detector A only measures
the spin of particle A and does not do anything to particle B. (It
cannot: the interaction of the particle with a detector is local,
and the detectors are so far apart that even light cannot reach
B from A before the particle A is measured.) The eigenstates
1
of z are = with the eigenvalue +1 and = 0
0 1
with the eigenvalue -1. The initial state of the system was
EPR = . Therefore after measurement it will be either
2
(if the measured spin projection SZA is +1/2) or (if SZA is
1/2). Either outcome has the probability .
Now an important point: s A + s B = 0, the total spin of the system
is zero. Therefore at the same moment, when the apparatus A
measures the spin sZA of particle A, the spin of particle B becomes
fully determined: szB = szA . If szA = 12 , then szB = 12 , and vice versa.
+
= ; = .
2 2
316
A simple exercise
Check that all this is actually so.
= ,
2
+
= ,
2
1 Do not worry about the minus sign. You can multiply an eigenstate by
any number, and it remains an eigenstate with the same eigenvalue. In other
words: it is the modulus squared of the state vector that matters.
1 +
= +
2 2 2
1 +
= + .
2 2 2
If now we measure the x-component of spin of the particle
+
A, the state of the particle B will be either =
( ) or ( )
2
if s xA = 1
= if s xA = 1 . In other words, if the
2 2 2
system is in the state , both particles spins are pointing in
the same direction. (Please do not think that this means that the
expectation value S x = 1. It is zero: there will be equal
contribution from the cases s xA = s xB = 1
and sxA = s xB = 1 . )
2 2
As you may recall from Chapter 7, if the total angular moment of a
system is S, then its projection on any given axis can take exactly
(2S + 1) values in our case, 2 1 + 1 = 3. The state has SZ = 0.
The other two states are with SZ = 1 and with SZ = 1.
These three states form a triplet, while the EPR state with its zero
spin and therefore zero projection of spin on any axis is a singlet.
318
with equal probabilities (if the detectors are at 90). What is
worse, the detector A can be reset at any moment, e.g. right
before the particle A arrives, which completely eliminates any
possibility of a signal from A reaching B to tell about the setup.
It looks like the events at A and B are, after all, connected by
what Einstein called spukhafte Fernwirkung or spooky action
at a distance. Einstein was very unhappy about it. According to
Einstein, the particles A and B should have real (that is, really
existing) values of z- and x-components of their spins all the
time.2 The fact that the operators z and x do not commute and
therefore cannot have common eigenstates meant to Einstein
that quantum mechanics is not complete: the state vector does
not tell us everything about the system.
Niels Bohr responded by saying that the correct way
of describing a quantum system depends on how this
system interacts with its macroscopic environment (e.g.
with the measuring apparatus). For example, a system
of two particles can be described by the operators
p1 , x 1 , p 2 , x 2 of their momenta and positions such that
[ x 1 , p1 ] = i, [ x 2 , p 2 ] = i, but [ x 1 , p 2 ] = 0, [ x 1 , x 2 ] = 0, etc.
On this point of view, since either one or the other, but not both
simultaneously, of the quantities P and Q can be predicted, they
are not simultaneously real. This makes the reality of P and Q
depend upon the process of measurement carried out on the
first system, which does not disturb the second system in any
way. No reasonable definition of reality could be expected to
permit this.
(Einstein et al., 1935: 777)
2 The EPR paper used a different model, but the gist of it is the same.
320
measurement.3 We can even draw a table illustrating the results
of measurements performed on two particles of spin 1/2 in the
EPR state, EPR = :
2
1/2 50%
+ + + + +
= = = ,
2 2 2
= AB.
Such states are called factorized. If a system is in a factorized
quantum state, one can measure one particle without disturbing
the other. Any non-factorized state is entangled.
322
The answer is negative. Suppose we have set up a superluminal
communication channel by placing somewhere in space a
source of entangled pairs of particles of spin (or photons:
instead of spins, we could use the polarization of light, and
instead of detectors, polarizers, but that would not change the
outcome). One particle of each pair arrives at the detector A,
which we will use as a transmitter, and the other (a little later)
at the detector B (serving as a receiver). Let every pair be in an
EPR state, and assume that none is lost in space, detectors work
perfectly, and so on and so forth. If both A and B detectors are
aligned in z-direction, then every time there is a +1/2 readout
there, there will be a 1/2-readout at B:
a readout:
+ + +
sz
B readout:
+ +
sz
324
Classical 1-1234 1-1234
2-4678 2-4678
3-9067 3-9067
4-AUG6 4-AUG6
5-77UJ 5-77UJ
6-6531 6-6531
7-HG09 7-HG09
8-7751 8-7751
9-1132 9-1132
10-34FZ 10-34FZ
11-Pq09 11-Pq09
EPR
0 1
1 1 1 1
0 1 0 0
0 1
0 0
1 0 0 1
1 0 0 1 0
1
M N O P
L VW X Y Q
K U Z
T
R
F G Q I J
S
A
S T D V W
R
B
H
C
O P
U
E
N
F
E
G
M H X
D J K L I Y
Z A B C
326
The strength of this kind of encryption is that there are absolutely
no statistical correlations in the coded message, which could be
used by the cryptanalysts to break the code. Another letter Y in
the message (say, on the 227th place) will be encrypted using
another, totally unrelated random number, e.g. 616: 25 + 616
(mod 26) = 17; therefore this time Y Q. So as long as Eve does
not have access to the one-time pads, and each pad is only used
once and then destroyed, the system is unbreakable.
The only problem is how to securely distribute the one-time
pads and ensure that they are really unique and random. And
here the EPR pairs can be of assistance (Figure 11.5). Let the
sender and the recipient (traditionally called Alice and Bob)
align their detectors and receive a stream of EPR pairs. Then
each will register a random sequence of zeros and ones (or
ups and downs, corresponding to the detected spins). These
sequences can be transformed into lists of random binary
numbers, that is, ready one-time pads. They are as random as
one can get, because their randomness follows directly from the
fundamental randomness of Nature. They are identical (Alices
zeros correspond to Bobs ones and vice versa they can agree
beforehand, that Bob will flip his sequence). And they cannot be
stolen or listened to on the way.
A B
A B
Figure 11.5 a very simple (and not quite practical) scheme for quantum
cryptography
328
H , that an evolution described by the Schrdinger
equation i t = H ( t ) would take the state of the
system from (t = 0) = to (t = T ) = . Therefore
Eve cannot make a copy of the state , keep it for further
investigation, and send the original to Alice and/or Bob.
The no-cloning theorem does not say that you cannot reproduce
a known quantum state. But to know a quantum state means
knowing its projections on all the axes of the Hilbert space,
that is, measuring it in an appropriate basis as many times as
the dimensionality of the Hilbert space. This is what is being
done in the process called quantum tomography of an unknown
quantum state. Since a measurement generally changes the
quantum state, you would need at least as many copies of your
system in this state, while by the conditions of the theorem you
have only one.
Now suppose you do not know the quantum state of your
system, but you can repeatedly prepare it in this state. There is
no contradiction here. You can always initialize your system
in some eigenstate n of an observable by measuring this
observable and using only the outcomes when the measured
eigenvalue is an. Then you can evolve the system using the same
Hamiltonian H (t) (e.g. applying a given sequence of pulses of
microwave or laser radiation), and the result should be each
time the same unknown quantum state vector ) (allowing for
the inevitable experimental errors, of course). Therefore we can
measure this state repeatedly and find it. If the Hilbert state of
the system is infinite-dimensional, or if it is in a mixed state, the
procedure is more involved, but will yield the quantum state of
the system with arbitrarily high accuracy (at least, in theory).
We will discuss the cost of this accuracy in the next chapter.
330
Some of this unhappiness could be attributed to the attitude
of some physicists and philosophers, who saw in quantum
mechanics the downfall of traditional scientific objectivism,
according to which whatever we observe and investigate exists
independently of our conscience, and trumpeted the triumph
of the subjectivist worldview. The news about the demise of
scientific objectivism was premature, though our understanding
of what is objective reality had to be seriously amended in
view of what quantum mechanics tells us about the world.
The famous Schrdingers cat Gedankenexperiment is a good
example (Figure 11.6). A cat is placed in a closed container
together with a radioactive atom (e.g. polonium-210), and a
Geiger counter connected to the valve of a can of poisonous
gas. Polonium is an -radioactive element with half-life T=138
days, which decays into the stable isotope, lead-206. The
quantum state of the polonium atom can be therefore written as
a superposition
2tT 210 Tt
(t ) = 2 84 Po + 1 2 206
82 Pb + .
210
84 Po closed can live cat or 206
82 Pb + open can dead cat . . )
An extreme subjectivist point of view states that the collapse
is caused by the act of observation by a sentient being (e.g. a
human). Logically it cannot be refuted, but not all statements,
which cannot be logically refuted, are true.5 Eugene Wigner
proposed an extension of the cat Gedankenexperiment, in which
Wigner himself does not open the container. Instead he sends
in his friend and waits for his report. Then, the vector should
describe the state of the system up to the moment when the
friend returns (assuming that Wigners friend likes cats):
2tT 210
(t ) = 2 84 Po closed can live cat happy friend
Tt 206
+ 1 2 82 Pb + open can dead cat sad friend .
This state collapses only when Wigner sees his friend and this
is not quite a reasonable picture despite its logical irrefutability.
Then, there is a question of how sentient the observer must be.
Will a cat be able to observe itself? Will there be any difference
between a clever and a dumb cat (or a clever and dumb friend)?
How about mice? Where do we stop? Will getting a PhD
improve your ability to collapse a quantum state?
A more reasonable approach would be to ascribe the collapse to
the interactions inside the container, which will either kill or not
kill the cat even before anyone bothers to look. The umbrella
term for these interactions is decoherence. The key requirement
5 For example, one can claim that he or she is the only person in
the universe and everything else is just an illusion. This is a logically
unassailable position but trying to live it would be a disaster, and writing
books promoting such a point of view (solipsism) to allegedly illusory
readers is not quite consistent.
332
of decoherence is the presence of a macroscopic system, but the
question remains how macroscopic it should be.
The problems concerning the existence of a macroscopic
quantum system in a quantum superposition of states could
be considered of no practical interest, until it turned out that
quite large systems could be put in a quantum superposition of
different states, if only one manages to properly insulate them
from the environment. We will therefore consider these questions
in the next chapter, when discussing quantum computing.
Bells inequality*
The nonlocality of quantum mechanics found its exact
mathematical expression in Bells inequality (derived by John
Stewart Bell (19281990) in 1964).
Bell was interested in what kind of a theory with hidden
variables could reproduce the predictions of quantum
mechanics, particularly for the EPR experiment.
Let us once more consider two spin-1/2 particles with zero total
spin, which are measured by detectors at points A and B, and
assume that there is a set of hidden variables , which completely
describe these two particles. If we reject action-at-a-distance
and insist on locality, then the result of measurement at A must
be independent on that at B (Figure 11.7). For example, the
probability of measuring spin up at A and spin down at B will be
PAB (; ) = PA (; )PB (; ).
A B
z z
x x
A B
A B
PAB () = 0 = d P
AB (; ) ( ) = d P (; ) P (; )().
A B
334
variables. Let us call a the variable, which determines the
outcome of the sx measurement on particle A: a = 1 if the
measurement gives sAx = 12 . In the same way we introduce
a = 1 for the measurement of sAz , and a = 1 for the
measurement of sA45 , and similar variables for the measurements
performed on the particle B.
Now, the probability for a particle from the EPR pair to have
values = 1, = 1 equals
P( = 1, = 1) = P( = 1, = 1, = 1) + P( = 1, = 1, = 1)
(since either = 1, or = 1).
Probabilities are non-negative, therefore
P( = 1, = 1, = 1) P( = 1, = 1) and
P( = 1, = 1, = 1) P( = 1, = 1), and we have
P( = 1, = 1) P( = 1, = 1) + P( = 1, = 1).
P( A = 1, B = 1) P( A = 1, B = 1) + P( A = 1, B = 1).
336
c an unknown quantum state cannot be copied
d a known quantum state cannot be copied
338
Schrdingers
12
elephants and
quantum slide
rules
The key feature of science is its ability to not just explain, but
to predict the behaviour of the world around us. From the
predictions of eclipses to the calculations of planetary orbits
and trajectories of spacecraft and ballistic missiles, from atomic
transitions to chemical reactions, from pure science to applied
science to engineering the power of science is built on its
predictive force, and the latter means a lot of computations. The
reason is that while the mathematical expression of the laws
of Nature may be quite simple, the corresponding equations
of motion practically never have exact solutions. One must
develop either some approximate schemes, or numerical methods
of solving these equations, or both. And in any case solving a
problem from the first principle remains a very challenging task.
In dealing with it we come to rely more and more on the support
of computers. But the use of computers creates problems of its
own, from developing a general theory of their operation to the
practical questions of their design, manufacture and maintenance,
to the point when engineering and fundamental physics are no
longer distinguishable. This makes the field of quantum computing
and more broadly quantum engineering especially interesting.
339
Celestial mechanics the application of the Newtonian
dynamics to the motion of the celestial bodies was the first
science where large-scale computations were required.
The ideal picture of planets following the elliptic orbits
around the Sun in accordance with Keplers laws is easily
derived from the Newtons laws and universal gravity, but only
if there are only two bodies in the system the Sun and a single
planet. This is just an approximation to reality: there are more
than two objects in this Solar system, and, unfortunately it is
impossible to solve exactly the equations of motion already
for three gravitating objects (except for some very special
cases). This is the famous three-body problem. It does have
some equally famous exact solutions (some of them found
by Lagrange), but it has rigorously proven that this problem
has no good solutions for the absolute majority of initial
conditions.
The situation with more than three objects is even worse.
Therefore in order to solve the Newtonian equations of
motion for the Solar system (and thereby confirm or refute
the validity of Newtonian dynamics and universal gravity,
by comparing these theoretical predictions with the actual
astronomical observations), it was necessary to build the
perturbation theory. For example, when making calculations
for Venus, first one had to take into account the pull from the
Sun (which is the greatest), then from Jupiter, then from other
planets, and so on and so forth.
In perturbation theory the solutions are found as a sum of many
(in principle infinitely many) contributions, finding which is a
time-consuming and labour-intensive process, requiring many
computations. First this was done manually (with the help of
the logarithmic tables, invented in the 17th century by John
Napier). The accuracy of these calculations by the mid-19th
century was sufficient to predict and then discover a new planet,
Neptune, based on the deviations of Uranus from its calculated
orbit. Nevertheless many effects in celestial mechanics such
as the relation between the satellites of giant planets and the
shape of their rings could only be investigated with the arrival
340
of powerful computers in the latter half of the 20th century,
because they involved such a large number of bodies that any
manual computations, even assisted by mechanical calculators
or slide rules, were hopeless.
The growing power of computers prompted the development
of a separate branch of physics, computational physics, along
with theoretical and experimental physics, devoted to a direct
modelling of the behaviour of various physical systems by
solving the corresponding equations of motion numerically.1
It was not only better numerical algorithms that were being
developed. The hardware was evolving fast: the speed
of computers steadily increased, and their size shrank.
According to the famous Moores law an observation made
back in 1965 the number of transistors in an integrated
circuit doubled approximately every two years, and the
trend, surprisingly, holds even now. But this shrinking of
the unit elements cannot go on forever. There is a limit to
how small they can become. In principle, this would limit
the power of computers, at least those built on the same
principles as the current ones. From here followed another
important question: whether a computer can efficiently model
physical systems.
Efficiently here means that there should be a reasonable
relation between the scale of the problem and the amount of
computational resources needed to solve it. If, for example,
in order to solve the problem of three gravitating bodies you
would need a desktop, for six bodies a supercomputer, and for
nine bodies all the computing power of Earth, the modelling
would not be efficient. Usually one expects from efficiency that
the computational resources should scale no faster than some
power of the size of the problem. So an approach requiring one
desktop for a three-body problem and 2, 22 = 4 or even, say,
210 = 1024 desktops for a six-body problem could be considered
1 For example, by slicing time and space in small pieces and replacing
the differential equations by the difference equations, as we have briefly
mentioned in Chapter 2.
342
state of a many-body system is factorized, for example,
= , there will be no problem. (We denote
1 2 N
and so on). Slide rules were fast, efficient and not too expensive,
and for a hundred years there was no engineer and hardly a
physicist without a slide rule.2
More complex analogue computers were built for solving
a specific problem, or a class of problems. One area, where
one early on had to solve complex equations as accurately as
possible, is the continuum mechanics (which includes such
disciplines as fluid mechanics and theory of elasticity). Its
development was spurred by the requirements of practical
engineering. Unfortunately, the equations, which describe
such systems be it an iron strut under varying stress or an
airplane wing in a turbulent flow are notoriously hard to
solve. Take, for example, the fundamental equations of fluid
mechanics, the NavierStokes equations, which describe the
motion of a viscous fluid and are known for almost 200 years.
Not only are they very hard to solve numerically. The question
of whether there exists or not a smooth and globally defined
solution to these equations remains a major unsolved problem
of mathematical physics. But finding accurate enough solutions
to these equations for a variety of conditions is absolutely
necessary when designing ships and aircraft.
One can use wind tunnels to directly recreate the desired
conditions and simply observe what the aircraft will do under
such and such conditions. But for a large aircraft or a ship this
would not do. It was proven though (back in the 19th century)
that a scale model of a ship can be tested at a different speed,
and maybe in a different kind of medium, and the results can be
recalculated back to the full scale. The installations where such
2 In Hayao Miyazakis last animated film, The Wind Rises, the heros slide
rule plays a prominent part.
344
tests are performed (ship model basins) are essentially analogue
computers (or simulators), which use one physical system to
predict the behaviour of a different physical system.
When building analogue computers one makes use of the
fact that different physical systems are often described by
identical equations. For example, a mechanical oscillator (a
spring pendulum with friction) and an electrical oscillator
(an LC circuit with resistive loss) satisfy the same equation,
d 2q dq
= 2q dt
, even though in one case q is the position of
dt 2
a pendulum and in the other it is voltage on a capacitor, and is
determined by the friction or electrical conductance. Therefore
in order to solve a system of differential equations one can build
an appropriate electric circuit (which will be much easier to
make and measure than an equivalent system of pendulums!)
and measure the appropriate voltages.
Taking up this approach, we therefore expect (with Feynman),
that the use of quantum computers to simulate quantum
systems should save us from the tight place. Again, we will call
a quantum computer any quantum system that can be used
to simulate (and therefore predict) the behaviour of another
quantum system.
A A
A NOT
A AB
A B AND
B
A AB
A B NAND
B
A A B A B XOR
B
A A
CNOT
(reversible XOR)
B A B
Examples of logic gates include NOT, AND, XOR, NAND and CNOT.
The bits (or qubits) are represented by solid lines, which in a
representation of an algorithm as a circuit run from the left (input)
to the right (output). The Venn circles illustrate the action of the
gates: the shaded area corresponds to the value 1 (TRUE) of the
output bit.
346
Digital computers avoid these difficulties by reducing the solution
of a problem to purely symbolic manipulations. Of course, in the
final analysis these manipulations are implemented by physical
processes in some physical systems (no information without
representation!), but these elementary operations are totally
independent of the problem to be solved and can be implemented
in a universal hardware. The idea of a universal digital computer
is an old one, going at least as far back as Pascal and Leibniz. It
was almost realized by Charles Babbage,3 but had to wait for the
electron valves and semiconductor transistors.4
Taken in the most abstract way, a classical digital computer
transforms an input into an output. An input and an output
can be represented as binary numbers, i.e. sequences of bits,
each taking value zero or one. A computer performs on the
bits a series of operations (logic gates), which form a circuit. If
the output is, e.g. the numerical value of some function, there
is no difficulty (in principle) to get a more accurate result by
increasing the number of bits in the input and registers.
A logic gate takes one or more inputs and transforms them into
one or more outputs. Examples are NOT, AND, NAND and
XOR, described by the following truth tables (coming from
logic: when a bit represents the validity of some statement, the
value 1 is called TRUE and the value 0 FALSE respectively):
a nOt a
0 1
1 0
5 Any logical circuit can be implemented with reversible gates only, though
this is not how conventional computers operate.
348
Table 12.4 truth table 4 (a cnOt B)
IN IN OUT OUT
a B a=a B = a XOR B
0 0 0 0
0 1 0 1
1 0 1 1
1 1 1 0
Quantum gates
First, the set of universal quantum gates includes one-qubit
gates, which should allow arbitrary manipulations with the
quantum state of a single qubit (just imagine rotating its Bloch
vector). Examples of such gates are Pauli-X, Y and Z gates,
given by the Pauli matrices x, y ,z from Chapter 4. If a qubit is
0 1 1 0
X 0 =x 0 = = =1;Y0
1 0 0 1
0 i 1 0
= y 0 = =
i 0 0 i
1 0 1 1
= i 1 ; Z 0 = z 0 = = 0 .
0 1 0 0
A useful exercise
Find the results of action by the Pauli gates on qubit states
0 +1 1 1 0 1 1 1
1 = 0 ; + = = and = = .
1 2 2 1 2 2 1
350
We get a more clear picture of how to apply a one-qubit gate
if we represent the qubit quantum state as a Bloch vector
(Chapter 4). There we stated that by playing with the qubit
Hamiltonian (that is, turning some knobs) we can turn the
Bloch vector from any initial to any final position on the Bloch
sphere. This is not an instantaneous process: rotation of the
Bloch vector takes time depending on the kind of qubit we use,
and there are clever sequences of turns (that is, of changes to the
qubit Hamiltonian by turning some knobs), which minimize the
required time while maintaining the accuracy of the operation.
Another very useful one-qubit gate is the Hadamard6 gate,
1 1 1 0 +1 0 1
H= 1 1 . As you can see, H 0 = ; H1 = .
2 2 2
In other words, the Hadamard gate puts a qubit, which was
initially in a state zero or one (as a classical bit in a classical
computer), in a quantum superposition of these two states.
Figure 12.2 Quantum cnOt gate (Zagoskin 2011: 314, Fig. a.1.1)
1 0 0 0
0 1 0 0
CNOT = .
0 0 0 1
0 0 1 0
1 . u u
1 . v v
= = ,
0 . u 0
0 . v 0
1 0 0 0 u u
0 1 0 0 v v
CNOT = = = ,
0 0 0 1 0 0
0 0 1 0 0 0
1 0 0 0 0 0
0 1 0 0 0 0
CNOT = = = ( u v ).
0 0 0 1 u v
0 0 1 0 v u
352
A B
Figure 12.3 two flux qubits interact via their magnetic fields
1
B 1 H
354
Key idea: heads or tails?
? ? D MM
XV
A.
D MM
REX P
AD..G.
XV
REX P
IN
M
GV
D.G.
INORV
IN
M
GV
INORV
Classical
?
Quantum
0 +1 0 1 1
in
=
2
2
=
2 (
0 0 +1 0 0 1 1 1 . )
in
out
=
1
2
(
0 f (0) + 1 f (1) 0 1 f (0) 1 1 f (1) )
f (0) 1 f (0) f (1) 1 f (1)
1
= 0 + 1
2 2 2
1
1 f (x) 1 f (x)
=
2
x 2
.
x =0
Recall that the function f(x) can take only values zero
and one. Therefore the state of the qubit B in the output,
f (x) 1 f (x) 0 1 1 0 0 1
2
, is either 2
(if f(x) = 0) or 2
=
2
(if f(x) = 1). We can rewrite the state of the two-qubit system
as out
= {
1
2
1
x =0 ( x )}{ } . The plus sign in is taken
0 1
2
356
It is usually convenient to measure a qubit in one basis of states
0 +1
{0, 1}, and in this basis we cannot tell the state from
2
0 1
the state . The situation is easily remedied, if apply to the
2
qubit A the Hadamard gate. You can check that
1 1 1 1 1
H + = = = 0.
2 1 1 0 0
1 1 1 1 0
H = = = 1.
2 1 1 1 1
Exercise
A qubit is in the state +. What is the probability of finding it in
state 0? Or in state 1?
358
numbers9, no matter how large. But it is hard to find the factors,
given the product. This operation lies in the basis of public key
cryptosystems (such as the popular PGP (pretty good privacy),
those used by the banks and internet merchants to conduct
money transactions, and of course the encryption systems used
by large corporations and governments).
Hard means that in order to find the factors of an N-digit
number, all known classical algorithms10 would require
the amount of time, which grows with N faster than any
power of N (that is, e.g. as exp(N)). In other words, they are
exponential, and not polynomial, in N.
In practice, it may take decades or centuries in order to find the
factors and break the code by running the existing algorithms
on the existing classical supercomputers. What is even worse
is that as soon as somebody gets sufficient computing power
to break, say, a 128-digit key code, it is enough to switch to a
256-digit key in order to make your codes safe again (and such
switching is quite easy and cheap). On the other hand, the very
knowledge that somebody can break public key encryption in
a polynomial time would have very deep consequences for the
global economy, defence and politics.
0 0 0 0 + 0 0 0 1 + 0 0 1 0 + 0 0 1 1 + 1 1 1 1
2N
and then only keep the right one. Unlike the Deutsch algorithm,
the Shor algorithm is probabilistic. This means that if we
want to factorize some number M using this algorithm, the
quantum computer will give a correct answer Q only with a
finite probability P. But this is not really a problem. Factorizing
is a so-called NP-problem (nondeterministic polynomial time),
which essentially means that while it is hard to find an answer,
it is easy to check whether the answer is correct or not. In our
case we simply divide M by Q. If there is a remainder, we just
rerun the algorithm and check again. The probability of not
getting the right answer after n attempts is (1 P)n and will very
quickly go to zero as n grows.
Schrdingers elephants,
quantum error correction and
DiVincenzo criteria
In order to run the Shor algorithm, a quantum computer must
contain at the very least enough qubits to fit in the number to
be factorized. These hundreds of qubits must be manipulated
in a very precise way (we have seen that even a simple Deutsch
algorithm requires a lot of manipulations with qubit states
in order to extract the information), and all the time they
must be maintained in a coherent superposition of many
360
different quantum states. This is a notoriously difficult task. A
superposition of two quantum states of a single or a few qubits,
00 + 11
like , is routinely called a cat state, and as we know they
2
are already very fragile. In a useful digital quantum computer
we must deal with real Schrdingers elephant states and
somehow ensure their survival during its operation.
The decoherence time of a quantum system generally drops
fast as the size of the system increases, and the decoherence
time even of a single qubit is not long enough to perform
the necessary number of quantum logic gates. Therefore the
quantum state of a quantum computer will be disrupted long
before it will finish the assigned task, and the result will be
garbage.
The designers of classical computers met with a similar
problem. Even though classical bit states, zero or one, are much
more robust than the qubit states, they are subject to noise and
fluctuations. In order to prevent errors, each logic bit is encoded
by several physical bits. This redundancy (like the Alpha, Bravo,
Charlie, etc.) with the periodic error correction ensures the
stability of information.
The simplest example of classical error correction is the three-
bit majority rule. Each logic bit is encoded in three physical bits:
0 000,1 111. In the absence of noise all three bits must
have the same value. From time to time the bits are read out
and corrected. For example, 010 000,110 111. This scheme
protects only against single bit-flips: if two bits were flipped
between the check-ups, the majority rules will make the mistake.
No error correction method protects against all possible errors
(a complete physical destruction of a computer is also an error
of sorts), but the ones employed in modern computers are
more than adequate. Unfortunately these methods cannot work
with quantum computers, because the intermediate states of
qubits are quantum superpositions and will be destroyed by a
measurement.
The goal of error correction is to restore the state of a
computer to what it was before the error happened. As we
know, the rigorous no-cloning theorem (Chapter 11) prohibits
362
Table 12.6 three qubits, four outcomes
Syndrome: This means that:
00 no error happened
01 Qubit B flipped
10 Qubit c flipped
11 Qubit a flipped
1 well-defined qubits;
2 qubits initialization in a desired state (e.g. in the state
0 0 0 0 0 0 0 ; )
3 a long decoherence time;
4 a universal set of quantum gates;
5 qubit-specific measurements.
The requirements are quite logical, but they interfere with each
other and are generally hard to meet. For example, a universal
set of quantum gates and qubit-specific measurements mean that
control and readout circuits must reaching if not all, but many
13 To give you an idea, at the present time, the number of quantum gates
one can apply to a superconducting qubit before it decoheres is about a
hundred and growing.
14 Assuming, of course, that there is no fundamental ban on large enough
systems being in a quantum coherent state. It may be that the best way of
finding it out is to try and build a quantum computer.
364
E
En+1(t)
En(t)
E1(t)
E0(t)
t
Ground state
Figure 12.5 adiabatic evolution of energy levels
Adiabatic evolution in the case of a single qubit (for example, a
particle in a two-well potential, Chapter 4) can be explained as
follows. The Hamiltonian
1 (t )
H(t ) =
2 (t )
If initially the qubit is in the ground state and then the bias changes
slowly from minus to plus infinity, the qubit will have time to
tunnel from the right to the left potential well, and will end up in
the ground state. If the bias changes fast, the qubit will remain
trapped in the right potential well, that is, in the excited state. (This
trapping is called the Landau-Zener-Stckelberg effect and is a
major source of errors in adiabatic quantum computing).
366
landscape that is, there will be a huge number of states with
almost as low energy as the ground state, and the system will
almost certainly get stuck in one of these metastable states,16
none of which is even close to our solution. This is where we
use the adiabatic theorem. We will initially make the energy
landscape of our system very simple so that it will quickly
cool down to its ground state, 0(t) and then will change the
system very slowly. In the end we will be still in the ground state
(with a probability close to one), but this ground state 0()
is now encoding the answer to our important problem. This
approach is called the adiabatic quantum computing.
It has some clear advantages. Any quantum program can be run
on an adiabatic quantum computer: the sequence of quantum
gates can be replaced by a specific pattern of static qubit-qubit
interactions. No need to apply quantum gates means that there
is no need in a precision manipulation with quantum states of
qubits. The ground state is the lowest energy state and if the
conditions of the adiabatic theorem are satisfied, it is separated
from the other states by a finite energy gap. Therefore as long
as we maintain the temperature of our system smaller than
this energy gap, suppress the noise coming from the outside
and manipulate the system slowly (to avoid the Landau-Zener-
Stckelberg effect), the system will not be disturbed, and its
quantum coherence should be preserved.
Not all these advantages are as clear-cut as one would like. The
rewriting of a circuit based quantum programme in terms of
an adiabatic quantum computer requires some unrealistically
complex patterns of couplings between qubits. There are
reasons to believe that one cannot operate this kind of a device
indefinitely long, since its quantum coherence will be eventually
disrupted. The energy levels of such a device may not satisfy the
conditions of the adiabatic theorem. But still, it is a tempting
proposition, and there can be some short cuts.
One such short cut is a quantum optimizer or quantum
annealer. This is a limited-purpose adiabatic quantum
computer, a kind of quantum slide rule. It is indeed like a
quantum counterpart to a classical analogue computer. It would
solve not an arbitrary problem, but a specific one, or a specific
limited class of problems. It may not even solve them exactly,
but give a good enough approximation. Many important
problems either belong to this class (optimization problems) or
can be reduced to them.
A good example is the travelling salesmans problem: given
a set of towns connected by roads, find the shortest route
passing through each town at least once. This is an important
logistical problem, especially if you can solve it (or produce a
good enough approximation) in real time. And this is an ideal
problem for a quantum optimizer.
The machines produced by D-Wave Systems Inc. (Figure 12.6)
belong to this class of quantum devices. They contain hundreds
368
Figure 12.6 D-Wave systems quantum optimizer
370
to tell, whether and to what degree they are really quantum
devices. As of now the jury is still out but the very fact that
we can (actually must) discuss and test quantum devices on this
scale is in itself extremely exciting.
? Fact-check
1 The main difference between an analogue and a digital
computer is that
a a digital computer is built of electrical circuits
b an analogue computer works faster
c an analogue computer uses a physical process, which
mimics the problem to be solved
d a digital computer has an unlimited accuracy
3 A logic circuit is
a a set of electronic circuits on a chip
b a set of universal logic gates
c a set of logical operations realizing an algorithm
d a set of truth tables or Venn circles
a 1 u +v
2 u v
b 1 u v
2 u +v
c 1 u
2 v
(u + v ) 0 + (u v ) 1
d
2
7 The Shor algorithm allows us to break
a one-time pad encryption
b public key encryption
c Enigma code
d Morse code
372
Dig deeper
m. Le Bellac, A Short Introduction to Quantum Information and
Quantum Computation. cambridge university Press, 2007.
a. Zagoskin, Quantum Engineering: Theory and Design of
Quantum Coherent Structures. cambridge university Press,
2011 (chapter 6).
a. m. Zagoskin, e. ilichev, m. Grajcar, J. Betouras and F. nori,
how to test the quantumness of a quantum computer? Front.
Phys., 30 may 2014. Online at
http://dx.doi.org/10.3389/fphy.2014.00033
374
Quantum mechanics was also a very special case, because its
discovery and early development seem having been propelled
by a tidal wave of sheer luck. Looking at the history of
science, it is hard to find another such example with the
experimental evidence first literally forcing scientists to build
from scratch an utterly counterintuitive theory, going against
all previously held assumptions, and then confirming the
theory whenever and wherever it was applied, and with an
unbelievable precision. Special relativity comes distant second,
in part, because for all its paradoxes it is a very simple
and straightforward theory; since Hermann Minkowski
demonstrated that it is actually a bit of pseudo-Euclidean
geometry, its consistency was never in any doubt and it never
had to be changed in any significant way.
1 Just imagine that instead of the photoelectric effect one had to start from
superconductivity.
376
they jump from one orbit to another. The jumps were totally
unpredictable, and the motion of an electron during the jump was
impossible to describe by any trajectory in real space and time. The
size of the orbital was determined by the demand that the angular
momentum of the electron was quantized in the units of .
This was the foundation of the old quantum theory. This picture
of fixed orbits and of indescribable and unpredictable quantum
jumps, accompanied by the emission or absorption of light
quanta, flew in the face of classical physics with its Laplacian
determinism. However, not only did it resolve the problems
with atomic stability and their fixed sizes, but it was in an
excellent agreement with the spectroscopic data, expressed by
such formulas as Balmers and Rydbergs. So Bohrs theory was
accepted, not as a finished theory, but as a working hypothesis,
pointing at some unknown properties of Nature. This was
lucky for more complex atoms with more electrons, the Bohrs
model is far from precise.
The old quantum theory was further developed by many
physicists, notably by Arnold Sommerfeld (of the Bohr
Sommerfeld quantization). It was Sommerfeld a brilliant
mathematician who worked out a more general model of
hydrogen and hydrogen-like atoms, with not only circular, but
also elliptic electron orbits not necessarily confined to the same
plane, and thus discovered other quantum numbers l,m
besides the principal quantum number n.2
The old quantum theory was not a consistent theory. It relied on
the BohrSommerfeld quantization rule, the adiabatic invariants
and the correspondence principle, proposed by Bohr.
378
Fortunately, Schrdinger decided to try a non-relativistic
approximation, which turned out excellent and produced the
equation we know and use. This was the wave mechanics.
At the same time a few months in advance Werner
Heisenberg (and Paul Jordan and Max Born) worked out the
matrix mechanics, which associated with every observable
(as we call them now) a matrix. Essentially, it was a way
of directly representing operators in Hilbert space, and the
Heisenberg equations of motion were first written for these
matrices.3 The achievement is even more remarkable, since
Heisenberg rediscovered the mathematics of matrices by
himself. Obviously, the linear algebra was not a standard part
of physics curriculum at that time. Matrices generally do not
commute, and this directly led Heisenberg to the formulation
of the uncertainty principle.
The same year Paul Dirac recognized the relation between the
commutators and the Poisson brackets and demonstrated both
the direct relation between quantum and classical mechanics
and the equivalence of Heisenbergs and Schrdingers
approaches. Finally, in 1926 Max Born proposed the
probabilistic interpretation of the wave function.
That was more or less it. The latter developments built on
this structure; extending it toward the relativistic domain;
applying to more and more complex systems; making it more
elegant; looking for unexpected or unsuspected conclusions and
applications; removing the scaffolding but the structure of
quantum mechanics was there to stay.
380
which rigidly separates quantum and classical worlds without telling
you where to draw the boundary. You have certainly heard about
the many-world interpretation proposed by Hugh Everett; the pilot
wave interpretation by David Bohm goes back to Louis de Broglie
himself. There are other interpretations. But the bottom line is that
as long as one talks about interpretations, they all must lead to the
same standard equations and rules of quantum mechanics and yield
the same predictions. FAPP For All Practical Purposes they are all
the same. I would therefore wholeheartedly support the dictum of
physicist David Mermin: Shut Up and Calculate.
But where is the physical sense? Well, what is the physical
sense, if not the intuition acquired by the experience of
dealing with concrete problems based on a specific physical
theory? In classical mechanics we deal with forces, velocities,
accelerations, which are all vectors, without much fuss but
there are no vectors in the Universe you could touch or point
at. (The realization that in the material world there is no
such thing as a triangle, a sphere or a dodecahedron, goes
back to the ancient Greeks, in particular Plato.) In quantum
mechanics we deal with quantum states, which are vectors
in Hilbert spaces, and observables, which are Hermitian
operators in these spaces, and there is no reason to make
fuss about that either. The main thing is that they correctly
describe and predict the behaviour of the outside world, the
objective reality, which is independent on anyones
subjective mind.
But is this a true description? Here we must talk about science
in general, the scientific method and the philosophy of science.
382
Key idea
The Allegory of the Cave from Platos The Republic is arguably
the most powerful and beautiful piece of philosophical writing in
human history.
I see.
And do you see, I said, men passing along the wall carrying all
sorts of vessels, and statues and figures of animals made of wood
and stone and various materials, which appear over the wall?
Some of them are talking, others silent.
Like ourselves, I replied; and they see only their own shadows, or
the shadows of one another, which the fire throws on the opposite
wall of the cave?
True, he said; how could they see anything but the shadows if they
were never allowed to move their heads?
And of the objects which are being carried in like manner they
would only see the shadows?
Yes, he said.
384
The proof of the pudding
The crucial importance of logical chains binding very remote
statements for science did not become clear at once. For much
of the time that passed since Aristotle those who had time, desire
and opportunity to investigate Nature, more often than not based
their investigation on some self-evident truths, which could be
directly extracted from either everyday experience or general
ideas of what is true, proper, simple or beautiful (and more
often than not, the outcome of the investigation should better
agree with the locally preferred interpretation of this or that
revealed truth). For example, the deeply rooted Greek belief that
a circle is the perfect shape (being indeed the most symmetric and
therefore the simplest to describe) forced astronomy into fifteen
centuries of Ptolemeian contortions, only ending with Keplers
laws. The obvious facts that every inanimate object on the Earth
will not move unless pushed, and will eventually stop, and that
the celestial bodies, on the contrary, move eternally and regularly,
produced the Aristotelian physics, which was only unseated by
Galileo and finally demolished by Newton.
The rise of modern science was made possible by the
recognition that one cannot demand simple simplicity and
obvious beauty; that an observation or an experiment may
need a painstaking interpretation, eliminating the side effects,
noise, mistakes; and that the correct theory of an effect may
be very far removed from everyday experience and common
sense. In other words, modern science was made possible by the
development of the hypothetico-deductive method.
386
notions. The birth of modern science or of science in the full
meaning of this word can be counted from the moment of this
recognition.
388
number of examples finite induction does not produce a
proof. Examples, though, can direct us to a valid statement.
This is what the reasoning by induction is about. Moreover,
in mathematics, there is a formal procedure (mathematical
induction), which allows statements to be proved based on what
can be thought as an infinite number of examples.
The inherent intellectual humility of the scientific method is
sometimes taken as a sign of its weakness. Science does not
produce eternal truths! Science is tentative! Science was wrong
in the past, so it may go wrong again! In philosophy such views
are often based on the so-called Neuraths boat argument:
We are like sailors who have to rebuild their ship on the open
sea, without ever being able to dismantle it in dry-dock and
reconstruct it from the best components.
Otto Neurath, Protocol Sentences, in A.J. Ayer (ed.) Logical Positivism
(1959)
390
(To finish with another military analogy, the philosophy of
science can be compared to a baggage train. It is certainly
useful, and sometimes absolutely indispensable, but it should
follow the army of science. If the baggage train gallops ahead of
the troops, it may mean one thing only: the army is routed.)
At the end of the day, the validity of scientific theories is
confirmed by their routine everyday applications. For example,
a computer I use to write these words is the best proof that, e.g.
quantum physics of solid state (microchips!) and light (lasers!)
correctly grasps some very essential truths about Nature.
Dig deeper
h. Poincare, The Foundations of Science: Science and Hypothesis,
The Value of Science, Science and Method. createspace
independent Publishing Platform, 2014.
m. Kline, Mathematics, The Loss of Certainty. Oxford university
Press, 1980.
m. Kline, Mathematics and the Search for Knowledge. Oxford
university Press, 1985.
B. Russell, Wisdom of the West. crescent Books, 1978.
392
Fact-check Answers
CHApTEr 1 CHApTEr 3 CHApTEr 5
1 (b) 1 (b) 1 (a)
2 (d) 2 (d) 2 (a)
3 (b) and (d) 3 (c) 3 (b)
4 (d) 4 (a) 4 (b)
5 (b) 5 (a) 5 (b), (c) and (d)
6 (c) and (d) 6 (b) and (c) 6 (a) and (b)
7 (a) and (b) 7 (c) 7 (b) and (c)
8 (c) and (d) 8 (d) 8 (b) and (c)
9 (c) 9 (a), (b) and (c) 9 (b)
10 (b) 10 (b) and (c) 10 (d)
394
Taking it further
Baggott, J., The Quantum Story: A History in 40 Moments.
Oxford University Press, 2011.
Feynman, R. P. and A. R. Hibbs, Quantum Mechanics and
Path Integrals. McGraw-Hill Companies, 1965 (there are later
editions).
Feynman, R., R. Leighton and M. Sands, Feynman Lectures
on Physics, Vols I, II and III. Basic Books, 2010 (there are
earlier editions). See in particular Vol. I (Mechanics and some
mathematics, Chapters 125), Vol. III (all of it quantum
mechanics) and Vol. II (Principle of least action Chapter 19).
Gamow, G., Thirty Years that Shook Physics: The Story of
Quantum Theory. Dover Publications, 1985.
Hamill, P., A Students Guide to Lagrangians and Hamiltonians.
Cambridge University Press, 2013.
Hawking, S. (ed.), The Dreams That Stuff is Made of: The Most
Astounding Papers of Quantum Physics and How They Shook
the Scientific World. Running Press, 2011.
Kline, M., Mathematics, The Loss of Certainty. Oxford
University Press, 1980.
Kline, M., Mathematics and The Search for Knowledge. Oxford
University Press, 1985.
Lancaster, T. and S. J. Blundell, Quantum Field Theory for the
Gifted Amateur. Oxford University Press, 2014.
Le Bellac, M., A Short Introduction to Quantum Information
and Quantum Computation. Cambridge University Press, 2007.
Lipkin, H. J., Quantum Mechanics: New Approaches to Selected
Topics. Dover Publications Inc., 2007.
Mattuck, R. D., A Guide to Feynman Diagrams in the Many-
body Problem. Dover Publications Inc., 1992.
396
Figure credits
Figure 4.16 Copyright 2003, American Physical Society
Figure 5.6 A. M. Zagoskin 2011, published by Cambridge
University Press, reproduced with permission.
Figure 8.5 Quantum Theory of Many-Body Systems,
Figure 2.1, 2014, p. 55, A. Zagoskin Springer
International Publishing. With permission of
Springer Science+Business Media.
Figure 8.8 T.-S. Choy, J. Naset, J. Chen, S. Hershfield and
C. Stanton. A database of fermi surface in virtual
reality modeling language (vrml). Bulletin of The
American Physical Society, 45(1): L36 42, 2000.
Figure 8.10 Quantum Theory of Many-Body Systems,
Figure 2.3, 2014, p58, A. Zagoskin Springer
International Publishing. With permission of
Springer Science+Business Media.
Figure 10.5 Quantum Theory of Many-Body Systems,
Figure 4.5, 2014, p167, A. Zagoskin Springer
International Publishing. With permission of
Springer Science+Business Media.
Figure 11.6 A. M. Zagoskin 2011, published by Cambridge
University Press, reproduced with permission.
Figure 12.2 A. M. Zagoskin 2011, published by Cambridge
University Press, reproduced with permission.
Figure 12.6 Copyright 2011 D-Wave Systems Inc.
398
Greens functions of 245 compound bosons 205, 301, 302
occupation numbers 298300 computers 3412
see also photons analogue 3436
bra vectors 72, 73 digital 343, 3479
Brattain, Walter 298 see also quantum computing
Brillouin zones 234 conduction band 236
conduction electrons
calculus 6 in crystals 2307
calculus of variations 261 dressed 238, 24950
canonical momentum 26 in jellium model of metals 2229,
carbon atom 201 2378, 245
carroll, Lewis 324 in superconductors 298, 3024
cartesian coordinates 2, 46, 98 configuration space 256
cathode ray tube 88 conjugate variables 22, 423, 61,
celestial mechanics 3401 62, 689
charge qubits 102, 1301 cooper, Leon 298
superconducting 102, 103, 1269, cooper pairs 102, 126, 1289,
13949 3024, 306
charge screening 2245, 226, coordinate systems 23
2378, 2489 cartesian 2, 46, 98
charged quantum particles in generalized 256
electromagnetic field 2708 spherical 989, 1934
circuits 347, 3489 coordinate transformation 23
clarke, arthur c. 390 correspondence principle 376,
classical bits 978, 111 3889
classical mechanics 230, 166 cosine 1314
newtons laws of motion 910, creation operators 20617
15, 2558, 261, 3401 cryogenic technology 293
perturbation theory 242 cubic oscillators 242
quantum-classical transition 139 curie point 282, 2856, 288
statistical mechanics 279, 312 cyclic frequency 16
variational approach to 25862
cnOt gates D-Wave systems quantum
classical 348, 349 optimizers 36870
quantum 3513 Davisson, clinton 186
code-breaking 126, 35860 de Broglie, Louis 185, 186, 377
see also encryption de Broglie wavelength 1867, 188
collapse of the wave function 1379 de maupertuis, Pierre-Louis
communications, faster-than-light moreau 261
3224 decoherence 3323
commutators 55, 601, 378 decoherence time 361, 364
complex conjugates 46 degrees of freedom 22
complex numbers 46, 489 density matrix 8991
Index 399
of qubits 1216, 1435 in atoms 192201
density of electron states 176, 229 charge screening 2245, 226,
dephasing of qubits 1235, 148 2378, 2489
derivatives 610, 55 and chemical properties of
partial 212 elements 2001
Descartes, Rene 2, 4, 17, 3867 in crystals 2307
detector back-action 148 density calculation 2289
Deutsch algorithm 3537, 362 density of electron states 176,
Dewdney, alexander 346 229
differential equations 1011, 21, distribution functions of 2259
434 dressed 238, 24950
differentiation see derivatives effective mass of 230, 238
differentiation operator 55 Greens functions of 2445,
diffraction of massive particles 2468, 24950
1847 interactions with phonons 245
digital computers 343, 3479 in jellium model of metals 2229,
dilution fridges 127, 369 2378, 245
Dirac delta function 67, 228 photoelectric effect 17981, 375
Dirac, Paul 67, 70, 77, 225, 269, 378 relativistic equation for 269
discrete energy spectrum 83 scanning tunnelling microscope
distribution functions 28, 29, 30, 1757
912 in superconductors 298, 3024
of electrons 2259 elements
dressed particles 238, 24950 chemical properties of 2001
dummy indices 74 transmutation of 1667
ellipses 17, 18
ehrenfest, Paul 41, 204 empty lattice model 2335
eigenstates 57, 589, 612 encryption 126, 32430, 35860
projection postulate 1367, 1556 energy bands 2347
of qubits 11215 energy conservation law 19
see also measurement energy eigenstates see eigenstates
eigenvalues 58 energy quanta 35, 378, 213
einstein, albert 36, 39, 45, 181, 319, energy quantization condition see
375, 379 quantization condition
einstein-Podolsky-Rosen energytime uncertainty relation
experiment 31220 1578
and Bells inequality 3335 entanglement 3203
and encryption 32430 epicurus 60
and entanglement 3203 ePR see einstein-Podolsky-Rosen
electronelectron attraction 298, experiment
3024 equilibrium states 28990
electron gas 2259 error correction 3613
electrons 88, 100, 102, 103 euler, Leonhard 49, 258, 261
400
eulers formula 49 monoatomic ideal 279
eulers identity 47 Gedankenexperiment 76, 1503,
everett, hugh 380 3312
expectation values 568, 59, 62 see also einstein-Podolsky-
of qubit observables 120 Rosen experiment
extremal action, principle of 25960 Geigernuttall rule 170, 173
extremal paths 2634 generalized coordinates 256
Germer, Lester 186
factorized quantum states 322 Gibbs factor 168, 170
faster-than-light communications Gibbs, Josiah Willard 28
3224 GinzburgLandau theory 2947,
FermiDirac distribution 2259 303, 305
Fermi energy 176, 2268, 2357 Ginzburg, Vitaly 294
Fermi, enrico 225 Goudsmit, samuel 204
Fermi momentum 228 gravity 10, 311
Fermi operators 2079 Greens functions 23940, 2448,
Fermi sphere 2289, 247 24950
Fermi surface 237, 302
fermions 2015, 2079, 210, 218 h-theorem 312
Greens functions of 2445, hadamard gate 351, 357
2468, 24950 hamilton, W.R. 20
see also electrons hamiltonian 20, 21, 26, 70, 78
ferromagnetic coupling 285 of Josephson junctions 306
Feynman diagrams 2429, 303 of magnets 283, 285, 292
Feynman, Richard 255, 265, 342 in perturbation theory 2413,
field operators 21416 2445
hermitian 21718 of quantum oscillators 21113
first order phase transitions 2867 of qubits 10912, 118, 1235, 128
Fock space 2089, 213, 216 time dependent 365, 366
Fock states 208 hamiltons equations 212, 26,
Fourier series 501 689
free energy 28990 harmonic oscillators 1420, 212,
Fresnel, augustin-Jean 80 35, 3641, 445, 21114
functional integrals 268 harmonic tones 51
functions 4 heaviside, Oliver 54
scalar product of 512 heaviside step function 226,
see also distribution functions; 2278
hilbert space heisenberg equations of motion
6971, 745, 767, 213, 378
Galileo Galilei 2, 152, 3856 heisenberg magnet 2912
Gamow, George 16970, 171, 173 heisenberg microscope 1502
gases heisenberg representation 71, 76,
electron 2259 77, 92
Index 401
heisenberg uncertainty relations Kamerlingh Onnes, heike 293, 296
62, 14952 Kapitsa, Leonidovich 3001
heisenberg, Werner 68, 69, 378 ket vectors 72, 73
helium-4, superfluid 294, 301 kinetic energy 17, 19
helium liquefaction 293 kinetic energy operator 216, 295
heller, michal 3867 KleinGordon equations 377
hermitian conjugates 108, 207, 210 Kronecker delta 208
hermitian field operators 21718 Kronig, Ralph 204
hermitian matrices 108, 109,
11920, 122 Lagrange equations 24, 262
higgs boson 205, 304 Lagrange, Joseph-Louis 245, 258
hilbert space 503, 567, 723 Lagrangian 256, 2589, 262, 266,
Fock space 2089, 213, 216 273, 274
of qubits 1045, 109 Landau, Lev 89, 91, 255, 289, 294
hilbert, William 282 Landau theory of second order
history of quantum mechanics phase transitions 28990
3748 Landau-Zener-stckelberg effect
hooke, Robert 14 366
hookes law 14, 15 Laplace, Pierre-simon 241
huygens, christiaan 80, 81 Laplacian operator 193, 194
hydrogen atoms 1927, 3756 Large hadron collider (Lhc) 167
hypervolumes 423 Larmor frequency 124
hypothetico-deductive method Larmor precession 124
38490 Leibniz, Gottfried Wilhelm von 6,
18
independent variables 21 Lenard, Philipp 35, 39, 180, 181
indistinguishability of elementary light
particles, principle of 2025 photoelectric effect 17981, 375
instantaneous eigenvectors 365 single-photon double slit
integrals of motion 157 experiment 1823
integration 18 wave theory of 802, 1812
interference linear frequency 16
of massive particles 1847 Liouville equation 30, 912
of photons 1823 logic gates 346, 3479
interference patterns 87 see also quantum gates
interpretation of quantum Lorentz force 284
mechanics 37880 Lorentz invariant 269
ionization energy 195, 196 Loschmidt, Johann 312
402
magnetic resonance imaging metals, jellium model of 2229,
(mRi) see nuclear magnetic 2378, 245
resonance (nmR) method of stationary phase 264
magnetism 2829, 2912 microscope
master equation 93, 123 heisenberg 1502
material points 46 scanning tunnelling 1738
matrices 715 microscopic states of systems 27,
hermitian 108, 109, 11920, 122 29
Pauli 11920, 121, 122 minkowski, hermann 269, 374
traces of 90 misener, Don 300
transposition 108 mixed state 90
see also density matrix modular arithmetic 3256
matrix mechanics 75, 378 momentum 10
matrix products 712, 734 angular 99100
maximally entangled states 322 canonical 26
maxwells equations 179 Fermi 228
measurement 56, 579, 13461 quantum angular 2034
collapse of the wave function momentum quantization 835
1379 momentum space 302
energytime uncertainty relation monoatomic ideal gas 279
1578 moores law 341
and heisenberg uncertainty multiplication operator 54
relation 14952
projection postulate 1367, 1556 n-particle probability distribution
quantum non-demolition (QnD) function see distribution
1558 functions
standard quantum limit 1534 natures intrinsic randomness 60,
strong 143, 1445, 146 137
strong continuous 15960 navierstokes equations 344
of superconducting charge qubits neurath, Otto 388
13949 neutrinos 121, 225
uncertainty 602 newton, sir isaac 1, 6, 8, 18, 386
weak continuous 146, 149 newtons laws of motion 910, 15,
Zeno paradox 15961 2558, 261, 3401
measurement problem 139 nmR see nuclear magnetic
mechanical energy 17 resonance (nmR)
mechanical energy conservation no-cloning theorem 3289, 3612
law 19 nobel Prize in Physics 1767
meissnerOchsenfeld effect 294, non-locality of quantum mechanics
297 31213, 333, 335
mendeleev, Dmitri ivanovich 200 normal flux quantum 277
mermin, David 380 nuclear magnetic resonance (nmR)
metallic bonding 2223 101, 111, 1246
Index 403
observables 56, 134, 166 phenomenological theories 2956
qubit 11920 philosophy of science 38090
see also measurement phonons 231, 245, 3034
occupation numbers 206, 298300 photoelectric effect 17981, 375
old quantum theory 3767 photons 17984, 299
one-time pads 32430 Planck constant 19, 35, 38, 423
operators 547, 601, 166 Planck, max 35, 36, 374, 379
annihilation 20617 Plato 3813
creation 20617 Poisson-arago-Fresnel spot 80
Laplacian 193, 194 Poisson brackets 689, 701, 92,
velocity 297 378
orbital quantum number 1967 Poisson, Denis 80
order parameter 288, 28990 Poissons equation 226
superconducting 296, 304, 305 polar angle 989
polonium 331
pairing potential 304 Popper, Karl 387
partial derivatives 212 potential energy 17, 1819
particle density operator 215 potential energy operator 216
particle number operator 20910 principal quantum number 195,
particle-wave duality 87 1967
particles 46 principle of extremal action 25960
in a box 824 principle of indistinguishability of
like waves 1848 elementary particles 2025
in a ring 845 probability distribution functions
waves like 17984 see distribution functions
path integrals 255, 26570, 2726 projection postulate 1367, 1556
Pauli exclusion priniciple 121, protons 1001
2012, 209, 225 public key encryption 126, 35860
Pauli gates 34950 pure state 90
Pauli matrices 11920, 121, 122
Pauli, Wolfgang 121 QnD see quantum non-demolition
pendulums 36, 3941, 2567 (QnD) measurement
see also harmonic oscillators quantization condition 35, 36,
periodic boundary conditions 845 3741, 60
Periodic table 199, 200, 222 Bohrsommerfeld 389, 41, 44,
perturbation theory 2403, 3401 45, 48, 376
phase space 1113, 413 quantum angular momentum
6N-dimensional 23, 25, 423 2034
phase trajectories 1620, 368, 48 quantum beats 11415, 118,
phase transitions 282, 2867 12931
Landau theory of second order quantum bits see qubits
28990 quantum-classical transition 139
superconductors 2934, 3023 quantum compass needles 3678
404
quantum computing 99, 101, quasiaverages 2912
12631, 345, 349 quasimomentum 2324, 235, 237
adiabatic 3678 quasiparticles 231, 24950
Deutsch algorithm 3537, 362 see also conduction electrons
DiVincenzo criteria 3634 qubits 97131
error correction 3613 ancilla 362, 364
and public key encryption 126, Bloch equations 11516, 118,
35860 1236
quantum gates 34953 charge 102, 1301
quantum optimizers 36870 controlling quantum state of
quantum parallelism 3537 11819
shor algorithm 35860 density matrix of 1216, 1435
quantum dots 102, 1301 dephasing of 1235, 148
quantum electrodynamics 184, 255 eigenstates of 11215
quantum ensembles 8991 encoding 3623
quantum equations of motion hamiltonian of 10912, 118,
see heisenberg equations of 1235, 128
motion measurement of 13949
quantum error correction 3613 observables 11920
quantum gates 34953 quantum gates 34953
quantum non-demolition (QnD) relaxation of 1235
measurement 1558 state vector of 1048
quantum numbers superconducting charge 102,
magnetic 1967 103, 1269, 13949
orbital 1967 superconducting flux 103
principal 195, 1967 see also quantum computing
spin 99101, 202, 204
quantum optimizers 36870 radioactivity 1667, 16972, 331
quantum oscillators 3641, 45, radius vectors 56, 98
21114 randomness 60, 137
quantum parallelism 3537 RayleighJeans law 374
quantum propagation 802 recollection theory 3813
quantum statistical mechanics reference frames 6
8891 relativistic quantum field theory
quantum superposition principle 184
867 relaxation of qubits 1235
quantum tomography 329 renormalization 238
quantum tunnelling 111, 173 Rohrer, heinrich 176
-decay 16772 rotation, vector 70, 11617
scanning tunnelling microscope running waves 85
1738 Rutherford, ernest 1667, 169, 283,
quantum uncertainty principle 375
423, 48, 60, 613, 82, 378 Rydberg 195
Index 405
Rydberg atoms 201 special relativity 374
Rydberg formula 196, 197 spherical coordinates 989, 1934
Rydberg, Johannes 197 spherical harmonics 194
spin 99101, 202, 204
scanning tunnelling microscope spin-statistics theorem 121, 202
1738 spontaneous symmetry breaking
schrieffer, John Robert 298 2901
schrdinger elephant 361 spring pendulums see harmonic
schrdinger equations 7780, 86, oscillators
269, 3778 standard deviation 61
Bloch waves 2323 standard quantum limit 1534
of electrons 1925, 200 state vector projection 1357
of Josephson junctions 307 state vectors 23, 4950, 534, 557,
quantum tunnelling 171, 173 166
for qubits 11112, 115 of qubits 1048
schrdinger, erwin 76, 269, 320, stationary states 80
330, 3778 statistical ensembles 289
schrdinger representation 76, 77 statistical mechanics 279, 312
schrdingers cat 76, 3312 quantum 8891
science, philosophy of 38090 stern, Otto 36, 45
scientific theories 385, 38790 stm see scanning tunnelling
second derivatives 9 microscope
second order differential equations stoletovs law 179, 181
1011, 21 strong continuous measurement
second order phase transitions 15960
286, 287 strong force 169
Landau theory of 28990 strong measurement 143, 1445,
superconductors 2934, 3023 146
second quantization 21617 subjectivist point of view 3312
self-energy 24950 superconducting charge qubits 102,
shockley, William 298 103, 1269, 13949
shor algorithm 35860 superconducting energy gap 304
sine 1314 superconducting flux qubits 103
single-photon double slit superconducting order parameter
experiment 1823 296, 304, 305
slide rules 3434 superconducting penetration depth
snowden, edward 330 304
soddy, Frederick 166, 169 superconducting phase 296
solid state theory 110 superconducting phase transition
sommerfeld, arnold 376 2934, 3023
sound waves 179 superconductors 2927
see also phonons Bcs (Bardeen-cooper-
spacetime 26970 schrieffer) theory 298, 3024
406
Josephson effect 3058 vector potential 2712, 274, 276
superfluid helium-4 294, 301 vector rotation 71, 11617
superposition principle 867 vectors
Bloch 1058, 11516, 11819,
tangents 4 1224, 1447
taylor expansion 260 cross product of 11718
taylor, Geoffrey ingram 183 ket and bra 72, 73
temperature, and magnetization radius 56, 98
282, 2856, 288 scalar product of 51, 73
thomasFermi length 226 see also state vectors
thomson, George 186 velocity 79, 10
thought experiments see angular 11617
Gedankenexperiment velocity operators 297
three-bit majority rule 361 VenOna project 32930
trajectories 6 virtual particles 2478
phase 1620, 368, 48 von neumann equation 913, 121
transistors 298 von neumann, John 136
transition amplitude 2656, 268
translational symmetry 232 wave equations 78
transmutation of elements 1667 wave function 7880
transpositions 108 wave packets 82
truth tables 3479 wave theory of light 802, 1812
tunnelling see quantum tunnelling waves
tunnelling matrix element 110, 111 like particles 17984
two-level quantum systems see particles like 1848
qubits running 85
two-well potential 110, 111, 11314 superposition of 87
weak continuous measurement
uhlenbeck, George 204 146, 149
uncertainty principle 423, 48, 60, Wiens law 374
613, 82, 378 Wigner, eugene 332
universal gravity constant 10 work functions 1801
universal quantum computers
3634 Young, thomas 182
Index 407