Quantum Information & Quantum Optics With Superconducting Circuits
Quantum Information & Quantum Optics With Superconducting Circuits
Superconducting quantum circuits are among the most promising solutions for the devel-
opment of scalable quantum computers. Built with sizes that range from microns to
tens of meters using superconducting fabrication techniques and microwave technology,
superconducting circuits demonstrate distinctive quantum properties such as superposition
and entanglement at cryogenic temperatures. This book provides a comprehensive and
self-contained introduction to the world of superconducting quantum circuits and how
they are used in current quantum technology. Beginning with a description of their basic
superconducting properties, the author then explores their use in quantum systems, showing
how they can emulate individual photons and atoms and ultimately behave as qubits within
highly connected quantum systems. Particular attention is paid to cutting-edge applications
of these superconducting circuits in quantum computing and quantum simulation. Written
for graduate students and junior researchers, this accessible text includes numerous
homework problems and worked examples.
J UA N J O S É G A R C Í A R I P O L L
Institute of Fundamental Physics (IFF), CSIC
www.cambridge.org
Information on this title: www.cambridge.org/9781107172913
DOI: 10.1017/9781316779460
© Juan José García Ripoll 2022
This publication is in copyright. Subject to statutory exception
and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.
First published 2022
A catalogue record for this publication is available from the British Library.
Library of Congress Cataloging-in-Publication Data
Names: García Ripoll, Juan José, author.
Title: Quantum information and quantum optics with superconducting circuits /
Juan José García Ripoll.
Description: New York : Cambridge University Press, 2022. |
Includes bibliographical references and index. |
Summary: “The dawn of the 20th century brought us the birth of quantum mechanics
and a deeper understanding of the miscroscopic world. The new theory describing photons,
atomic spectra and many other physical processes, postulates that the microscopic world is, in its
truest essence, probabilistic. Particles such as electrons or photons move or “propagate” as probability
waves to be detected at a given position, or in a given state. However, those waves or wavefunctions are
very different from a mere representation of our ignorance about the world”– Provided by publisher.
Identifiers: LCCN 2021061918 (print) | LCCN 2021061919 (ebook) |
ISBN 9781107172913 (hardback) | ISBN 9781316779460 (epub)
Subjects: LCSH: Quantum theory. | Quantum optics. | BISAC: SCIENCE / Physics / Quantum Theory
Classification: LCC QC174.12 .G359 2022 (print) | LCC QC174.12 (ebook) |
DDC 535/.15–dc23/eng20220521
LC record available at https://lccn.loc.gov/2021061918
LC ebook record available at https://lccn.loc.gov/2021061919
ISBN 978-1-107-17291-3 Hardback
Cambridge University Press has no responsibility for the persistence or accuracy of
URLs for external or third-party internet websites referred to in this publication
and does not guarantee that any content on such websites is, or will remain,
accurate or appropriate.
xi
xiii
2
Ô Uncertainty of operator, (Ô)2 = 12 Ô Ô † + Ô † Ô − Ô .
Variance of observable, Ô † = Ô, as (Ô)2 = (Ô − Ô)2 .
C Capacitance
EJ Josephson energy
H.c. Hermitian conjugate, as in a + H.c. = a + a †
L Inductance
L Lagrangian or Lindblad operator
x, s . . . Vectors of numbers such as x = (x1,x2, . . . ,xN )
h Planck constant, 6.626070040(81) × 10−34 J · s
h̄ = 2π h
Reduced Planck constant
φ Electric flux on a node or a branch of a circuit (Section 4.3)
0 = 2e h
Magnetic flux quantum, 2.067833831(13) × 10−15 Wb
ϕ0 = 0
2π
= h̄
2e
Flux-to-phase conversion
σ (H ) Spectrum or collection of eigenvalues of an operator H
σ x, σ y, σ z Pauli matrices
σ Vector of Pauli matrices σ = (σ x ,σ y ,σ z )
Linear capacitor
Linear inductor
Nonlinear inductance associated to a Josephson junction
Josephson junction: nonlinear inductor and capacitor
in parallel
Constant voltage source
xiv
The dawn of the twentieth century brought us the birth of quantum mechanics and a
deeper understanding of the microscopic world. The new theory describing photons,
atomic spectra, and many other physical processes postulates that the microscopic
world is, in its truest essence, probabilistic. Particles such as electrons or photons
move or “propagate” as probability waves to be detected at a given position, or in a
given state. However, those waves or wavefunctions are very different from a mere
representation of our ignorance about the world.
Quantum mechanics is the mathematical language that we use to describe the
microscopic world. Quantum mechanical states – our wavefunctions – go beyond
our classical understanding of the world, accepting the possibility of a system to
coexist in a quantum superposition of two distinguishable configurations – i.e., an
atom in two positions, a photon in two polarizations, or a neutron moving in two
opposite directions. Only when we measure the state of the quantum mechanical
object, it collapses to a well-defined configuration, which may be different on each
realization of the experiment. Even more dramatically, multiple particles may be
in a collective superposition or entangled state, allowing arbitrary measurements
on these particle to produce random but perfectly correlated outcomes, irrespective
of the space and time separation between those measurements. These and other
predictions of quantum mechanics made some physicists such as Albert Einstein
deem the theory as “incomplete” or even inconsistent.
The end of the twentieth century and the beginning of the twenty-first have
witnessed an incredible collective effort to challenge the wildest predictions of
quantum mechanics in the broadest variety of experimental systems possible, from
photons, to atoms, all the way to macroscopic solid-state devices and circuits. Great
experimentalists of our time have revealed quantum properties in objects that we
can see with the naked eye, created complex quantum states of tens to thousands of
particles, and entangled photons and spread them to different points of the planet,
to test their quantum mechanical correlations. In the process, not only has quantum
mechanics been validated, but we have become incredibly good at controlling
quantum systems.
Trapped ions were one of the first systems to enter this controlled quantum regime
(Leibfried et al., 2003). Using electromagnetic traps and laser cooling, experimen-
talists may isolate one or more charged particles, cooling some of their properties
close to the absolute. This makes it possible to observe the discrete or quantized
nature of the atomic excitations, not only at the level of the electronic states, but also
with respect to quanta of energy stored in the motional states of the trapped atom.
If instead of charged particles we use neutral atoms, we can trap more particles
and enter the regime of mesoscopic quantum systems. Experiments in 1995 showed
how to to trap and evaporatively cool a cloud of 106 alkali atoms in an almost perfect
vacuum down to few nanokelvins. At these temperatures, the ultracold atoms form
a singular state of matter called Bose–Einstein condensate (BEC) (Townsend et al.,
1997), in which all bosonic atoms are described by the same wave function, φ(x),
whose dynamics and quantum properties can be engineered at will.
Starting from a perfectly controlled state of matter, such as the condensate, we can
build extremely sophisticated quantum states and physical devices. We can split the
atoms of a BEC into the pockets of an optical lattice, to recreate a crystal of bosons
and simulate quantum magnetism and even Luttinger liquid physics. Bose–Einstein
condensates can be used to cool down fermionic atoms and create Fermi seas,
study Bardeen–Cooper–Schrieffer (BCS) superconductivity or even implement the
Hubbard model for electrons in a solid. And using time-dependent controls, we can
create large amounts of entanglement and squeezing that can be used for sensing or
for foundational experiments in quantum information.
The beauty of atomic, molecular and optical (AMO) physics’ bottom-up approach
can hardly be exported to solid-state devices. Condensed-matter objects are simply
too large. We cannot control all atoms in a piece of metal and convince them
to adopt a perfect state. Moreover, solid devices rarely exist in perfectly isolated
environments. They are always in contact with other elements – contacts, substrates,
measurement apparatuses – at higher temperatures than the AMO setups we just
discussed.
In a way, once we overcome the barriers associated to cooling and trapping, it is
not surprising that we can reveal quantum mechanical phenomena in large atomic
setups – it just confirms that the rules of quantum mechanics extend to very large
composite systems. The million dollar question, however, was whether those rules
extend all the way to solid-state and condensed matter systems!
This was indeed the great challenge for condensed matter systems around the
1990s. Can we prepare a perfect solid-state quantum system? Can we build a
simple, atomic-like device, with a perfectly controlled state? The approach to
this problem is subtly different from AMO. Starting from a macroscopic object –
a gated or self-assembled quantum dot, impurities in diamond, nanoresonators,
or superconducting circuits – we select one or two degrees of freedom in which
we seek evidence of quantum phenomena. The candidate for a quantum degree of
freedom may be the charge of a small superconducting island, the electrical current
along a metallic loop, or the wavefunction of electrons in a dot or a color center,
for example. Experimentalists focus on one property and work toward isolating
it from their environments, cooling them down, and engineering better and better
quantum states.
This challenge, the quest for an artificial atom – what we now call a qubit; see
Chapter 6 – was still ongoing when I first learned about the field of superconduct-
ing circuits. The review article by Makhlin et al. (2001) showed some promising
candidates in the forms of charge qubits (Section 6.2) or three-Josephson junction
qubits (Section 4.8), but quantum superpositions were fragile and easily killed by
the environment.
The surprise came in year 2004, when the Yale team (Wallraff et al., 2004) rad-
ically improved the lifetime of charge qubits by placing them inside microwave
resonators. That work, and subsequent works at various groups in the University of
California, Santa Barbara (UCSB), Saclay, Karlsruhe Institute of Technology (KIT),
etc., sparked the beginning of a productive intersection between superconducting
circuit technology, quantum optics, and quantum information, which we now call
circuit-QED.
We live in the second quantum revolution (Dowling and Milburn, 2003), in which
quantum mechanics becomes a tool for industrial applications and practical devices.
These applications are generally known as quantum technologies, because they use
the unique properties of quantum mechanics – quantum superpositions, entangle-
ment, noncontextuality, etc. – to create actual technologies for quantum communi-
cation, quantum cryptography, quantum sensing, quantum simulation, and quantum
computing.
Superconducting circuits are at the center the quantum revolution. Supercon-
ductors are the basis for the most accurate single-photon detectors, which are
used in advanced setups for quantum communication. We rely on superconducting
qubits for some of the most powerful and accurate quantum computers to date
(see Chapter 8), and also to simulate large and complex problems from quantum
magnetism (Chapter 9). Superconducting circuits have also found their way,
as controls and interfaces for other microwave-based quantum technologies,
from quantum dots to nanomechanical resonators. Understanding how these
circuits work, how they are designed, and how they are operated is a natural
path for both theoreticians and experimentalists alike. This book is a self-contained
undergraduate manual designed to help you gain this understanding.
In the second part of the book, Chapter 5 focuses on the linear models of super-
conducting circuits, such as microwave LC resonators and waveguides (the equiva-
lent of coaxial cables), and showing how the quantization of the circuit gives birth
to microwave photons that can be created, manipulated, and measured. Chapter 6
studies superconducting circuits in the highly nonlinear regime, in which the circuit
acts as an artificial atom or qubit. We review the most popular qubit designs, such as
the transmon or the flux qubit, and develop tools to prepare and characterize those
qubits. Chapter 7 combines artificial atoms with microwave photons, developing
the theory of light–matter interaction. This chapter is a primer on concepts from
quantum optics – models for low-dimensional atom–light interaction, the physics
of spontaneous emission, photon absorption, and open quantum systems. We then
focus on the simpler setup of qubit–resonator interactions, introducing the Rabi and
Jaynes–Cummings models, and showing how qubits can control the state of light
and vice versa, cavities can be used to control and measure qubits.
The third part of the book focuses on the applications of superconducting circuits
to various quantum technologies. We begin in Chapter 8 studying universal quantum
computers in the circuit model. We offer a checklist of ingredients that are needed
to build a practical quantum computer, and investigate how those elements are built
and characterized in the superconducting platform. We also discuss the roadmap
toward fault-tolerant, error-corrected, and fully scalable quantum computers, and
what we can do with their small-scale, faulty versions in the near term. Finally,
Chapter 9 builds on the previous formal developments to introduce a design for an
adiabatic quantum computer, also known as quantum annealer. This chapter starts
with a formal description of this computational model, where a physical system is
adiabatically coerced into a configuration that represents the solution to a mathe-
matical problem. It then moves on to the actual superconducting architecture for a
quantum annealer using superconducting flux qubits, along the line of the D-Wave
machine, but referencing other later designs.
If the book is to be used with a strong focus on quantum optics or quantum
circuits, the student or teacher should at least cover Chapters 1–7. This should give
students an overview of the most relevant experiments in the field from 2004 to this
day, providing them with the language and tools to access more complex literature
and explore their own ideas. Chapters 8 and 9 in the third part of the book are
independent from each other, but require a good understanding of the first part of
the book. A minimal set of exercises is provided within each chapter. Solutions and
errata will be posted on the author’s webpage.1
1 http://juanjose.garciaripoll.com.
1.2 Acknowledgments
This book would not be possible without the help of many colleagues and friends
who supported me both before and during the writing of the text. I was introduced
to the exciting field of superconducting circuits by my friend and colleague Enrique
Solano, who himself laid out the foundations to many ideas that are explored in the
realm of circuit-QED and photon measurements.
I have learned a lot interacting with fellow experimentalists, including Frank
Deppe, Achim Marx, Adrian Lupascu, Chris Wilson, Gerhard Kirchmair,
Mathieu Juan, and Aleksei Sharafiev. I am indebted to Pol Forn-Díaz, whose
curiosity and perseverance has strongly driven our interest in ultrastrong qubit-
microwave interactions and quantum annealing.
This book would not be born without the spark and motivation to write it, which
was ignited by my friend and outstanding physicist Oriol Romero-Isart. My visits to
Innsbruck – always a place for inspiration at various inflection points in my career –
and the lectures I gave there helped me move on and were the seed for this book.
I obviously owe many thanks to the students who have suffered various iterations of
this work, and also my former PhD student Borja Peropadre, whose thesis inspired
the first iterations of this project.
Special thanks must go to Carlos Navarrete, Guillermo Romero, Manuel Pino,
Paula García-Molina, Pol Forn Díaz, and Alp Sipahigil for carefully reading various
iterations of this manuscript.
None of this would be possible without the support of a family that has always
been there, even when I was not. I owe to them, not this book, but the privilege of a
life as scientist.
And finally, this book exists because of you, the reader. I hope it will help you
approach this field, learn about its beauty and its challenges, and inspire you to be
a better “quantum mechanic.”
physicist Max Planck showed that experiments measuring the radiation from black
bodies could be explained by assuming that these objects – which at the time were
just perfect cavities with a tiny hole – could only exchange energy in fixed amounts
or quanta, determined by the frequency ν of the light emitted or absorbed by the
black cavity.
In his treatment, Planck models the excitations of the black body as a collection of
harmonic oscillators with frequencies that cover the measured spectrum – i.e., ν(k),
labeled by wave vectors k. The energy of those oscillators is quantized, which means
that the oscillators equilibrate to the same temperature by exchanging discrete units
of energy or quanta with the environment. If we could measure the state of the
black body, its energy would be a sum of the quanta nk that are stored in each
electromagnetic mode k:
E= hν(k) × nk, nk ∈ {0,1,2, . . .}. (2.1)
k
In this model, each oscillator has associated a quantum of energy hν(k) determined
by the frequency1 ν(k) and Planck’s constant h 6.62607004(81) × 10−34 J/Hz.
The collection of all integers |nk1 ,nk2 , . . . is a unique configuration of the black
body, which we call the quantum state, and the collection of all states is used to
develop a statistical model of the black body’s spectrum.
Despite his success, Planck was very wary of extending the idea of quanta to
the actual electromagnetic field. It was Einstein who made the connection between
Planck’s quanta and the existence of a particle of light, the photon. With this
particle, Einstein could explain in 1905 the photoelectric effect: Some materials
may convert light into an electrical current, but when the intensity of light is
lowered enough, this current becomes a series of discrete random bursts, which
Einstein associated with the absorption of photons. This successful explanation was
shortly followed by Bohr and Rutherford’s model of the atom, based on quantized
electronic orbits that explained the discrete spectra of light-emitting atoms. Barely
a decade later, Schrödinger (1926) and Heisenberg (1925) replaced all ad hoc
quantization ideas with two equivalent formulations of quantum mechanics based
on wave and matrix equations. In both theories, the discretization of energies is
a mathematical consequence of the discrete spectra of the operators that govern
the evolution of light and matter. Put to work, the newly born theory provided
quantitative explanations for the spectra of atoms, molecules, solids, and the
electromagnetic field itself, consolidating nonrelativistic quantum mechanics as
1 Ordinary frequencies are typically denoted by the letter ν and are measured in the S.I. unit of Hertz (Hz). Quite
often we will also use angular frequencies ω = 2π ν, sometimes denoted in rad/s or s−1 . In the first case, the
quantum of energy is given by hν, while in the second case it is given by Planck’s reduced constant h̄ω with
h̄ = h/2π .
2 As we will see in Chapter 4, this is not a futile exercise, because the harmonic potential V (x) = 1 mω2 x2 is
2
formally analogous to the simplest electrical circuit, an LC resonator, and describes how this circuit is actually
quantized.
1
L(ẋ,x) = mẋ2 − V (x). (2.3)
2
According to the stationary principle, a small perturbation of the particle’s true tra-
jectory xε (t) = x(t) + ε(t) should leave the action unperturbed up to second-order
corrections S[xε ] = S[x] + O(ε 2 ). This stationary principle produces Lagrange’s
equations
d ∂L ∂L
= , (2.4)
dt ∂ ẋn ∂xn
which are equivalent to the original Newtonian equations (2.2).
We now introduce a Hamiltonian formulation, where the functional that generates
the dynamical equations is a function of two canonically conjugate variables, x and
p, with the prescription
∂L
H (x,p) = pẋ − L(ẋ,x), . with p =(2.5)
∂ ẋ
This Legendre transform establishes a link between the particle’s velocity ẋ and
its canonical momentum p, which now replaces the former in all equations. The
transform also produces an object, the Hamiltonian H (x,p), governing the orbits
of the particle. More precisely, any observable O(x,p,t) that we can construct as a
function of the canonical variables and time evolves according to the Hamiltonian
equation
d ∂O
O = {O,H } + , (2.6)
dt ∂t
with the classical Poisson brackets
∂A ∂B ∂B ∂A
{A,B} = − . (2.7)
j
∂xi ∂pi ∂xi ∂pi
In particular, since {xi ,pj } = δij this prescription trivially recovers Newton’s equa-
tions, but now expressed as a set of first-order differential equations:
d ∂H 1 d ∂H
x= = p, p=− = −∇V (x). (2.8)
dt ∂p m dt ∂x
linear Hermitian operators x̂, p̂, Ô acting on this Hilbert space. Each observable will
have a spectrum of eigenvalues and eigenstates determining all possible measure-
ment outcomes.
In our toy model, the quantum state of a particle that is at position r ∈ RN
is associated a vector |r in the Hilbert space, with the property x̂n |r = rn |r.
Generic states are constructed as quantum superpositions of different measurement
outcomes, such as the wavefunction ψ(r):
|ψ = ψ(r) |r dN r. (2.9)
The weights of the wavefunction ψ(r) ∈ C are complex numbers whose modulus
gives the probability distribution P (r) = |ψ(r)|2 that the particle is found at the
position r, if the observable x̂ is ever measured. States are normalized, so that the
total probability adds up to one, 1ψ = ψ|ψ = |ψ(r)| dr = 1, and we can
2
Note how, by using the orthogonality of position eigenstates r0 |x̂|r1 = r1 r0 |r1 =
r1 δ(r0 − r1 ), we recovered the formula for the average over the probability
distribution P (r).
Canonical quantization includes one final prescription that makes the algebra
of operators and states consistent with the classical limit of these equations. We
replace everywhere the Poisson brackets for classical variables with the commutator
between the respective observables {A,B} → −i[Â, B̂]/h̄, where [Â, B̂] = ÂB̂ −
B̂ Â. For the isolated particle, this prescription transforms {xn,pm } = δnm into
dÔ i ∂ Ô
= − [Ô, Ĥ ] + . (2.13)
dt h̄ ∂t
In this model, the dynamics is generated by a Hamiltonian operator that results from
replacing the canonical variables with the corresponding observables
Ĥ = 2m p̂ + V (x̂). In the Heisenberg picture, observables change in time starting
1 2
from a well-known initial condition O(t0 ) = O0 . The states |ψ0 remain stationary,
and they are regarded as objects that map the changing observables to their
expectation values Ō(t) = Ô(t) = ψ0 |Ô(t)|ψ0 .
The Heisenberg equation is rather inconvenient: We have to work with big
and complex operators, and extracting the measurement statistics becomes a very
convoluted process. In many situations, we would rather work with an equation
that determines how states evolve from, say, an initially localized configura-
tion ψ0 (x) = δ(x − x0 ), spreading to other meaurement outcomes. This infor-
mation is provided by the Schrödinger equation or Schrödinger picture, whereby
observables have an immutable representation, but wavefunctions change in time
Ō(t) = Ôψ(t) = ψ(t)|Ô0 |ψ(t), with
For our isolated particle in an external potential V (x), using the position represen-
tation, where p̂ = −i h̄∇, this results into a simple wave equation for the complex
amplitude of probability ψ(x):
1
i h̄∂t ψ(x) = (−i h̄∇)2 + V (x) ψ(x). (2.15)
2m
Even if they look very different, the Schrödinger and Heisenberg representa-
tions give the same predictions because they are both solved by a common unitary
Note that in this representation, |0 and |1 are the two eigenstates of σz , and we write
σz = |11| − |00|. If we enlarge the set of Pauli matrices to include the identity,
σ α |3α=0 = {1,σ x ,σ y ,σ z }, we have a basis where we can expand any observable in
this Hilbert space:
1
3
Ô = tr Ô σ̂ α σ̂ α . (2.20)
4 α=0
Ĥ = E + B n · σ̂, (2.21)
We will use this formula when studying superconducting qubits and the implemen-
tation of single-qubit gates in Section 6.1.4.
classical ensembles of pure states created with different classical probabilities p(χ )
(Ballentine, 1970):
ρ̂ = |ψχ ψχ | p(χ )dχ. (2.23)
This equation (2.23) describes the output of an experiment where parameters have
some uncertainty. Mixed states can also arise when a system enters in contact with
another system, called the environment. In principle, we should consider the global
wavefunction of the system plus its environment, allowing both to be correlated
|global = s,E s,E |s ⊗ |E ∈ Hsystem ⊗ Henvironment . However, since we will not
have access to all degrees of freedom of the environment, we must trace out all the
information that we ignore, obtaining a much smaller ensemble that only describes
our system:
ρ̂sys = trenvironment |global global | = s,E s∗,E |ss | . (2.24)
E s,s
Density matrices are particularly simple in the case of two-level systems, where
they can be expanded in the qubit basis ρij := i|ρ̂|j , or as a combination of Pauli
operators (2.20):
1
1
ρ00 ρ01 1
ρ̂ = = ρij |ij | = 1 + Sα σ̂ α . (2.25)
ρ10 ρ11 2 2
i,j =0 α=x,y,z
The set of all physical states with S = (Sx ,Sy ,Sz ) falls inside the Bloch sphere
|S| ≤ 1. The surface of this sphere is formed by pure state (|S| = 1), and its center
is the completely depolarized state S = 0 or ρ̂ = 12 1.
The completely depolarized state is an example of classical state, diagonal den-
∗
sity matrices – ρ10 = ρ01 = 0 – which may be constructed as a convex combination
ρ̂ = ρ00 |00| + ρ11 |11| of preparing state |0 with probability P0 = ρ00 and
preparing state |1 with probability P1 = 1 − P0 . Compare this now with the super-
position state, |+ = √12 (|0 + |1). This state exists only in the quantum model for
a two-level system. This particular superposition maximizes the off-diagonal ele-
ments ρ01 and ρ10 , also known as coherences, recognized as a true signature of
quantumness in states.
Sometimes a quantum system is sufficiently isolated and the timescales of study
are short enough that we can approximate its evolution with a Hamiltonian that
involves just that system. In that case, the unitary evolution of vectors in the Hilbert
space dictates a recipe for updating the density matrix ρ̂(t) = Û (t,t0 )ρ̂0 Û (t,t0 )† .
More generally, a quantum system will get entangled with its environment, suffering
2.4 Measurements
The axioms of quantum mechanics prescribe the behavior of a quantum system
under a complete measurement of any observable O, also known as Von Neu-
mann or projective measurements. Each observable is associated to a different
Hermitian operator Ô. The eigenvalues that result from diagonalizing this operator
on correspond to the possible measurement outcomes of the measurement. Let
P̂n = m |on,mon,m| be the projector onto all quantum states for which the
observable Ô has the value on . This projector is built using eigenstates of the
observable Ô |on,m = on |on,m, understanding that one measurement outcome
may be given by many different quantum states, which differ in other generic
quantum properties, here denoted as m. According to quantum mechanics, an ideal
projective measurement of the observable Ô onto a state ρ will produce the outcome
on with probability p(on ) = tr(P̂n ρ̂). If the measurement is also nondestructive, the
quantum state after the measurement will be projected onto a new density matrix:
1
ρ̂ → P̂n ρ̂ P̂n . (2.27)
p(on )
Figure 2.1 A resonator or cavity stores microwave photons whose state we wish to
measure. Instead of connecting the measurement apparatus to the cavity, we make
a weak connection between the cavity and an outgoing superconducting cable that,
after passing through an amplifier, brings the signal to the measurement apparatus.
The signal bout is proportional to the cavity signal acavity , but contains vacuum
noise. The measured signal, Gbout +ε, is amplified with a gain G > 1, but contains
additional noise ε from the amplifier.
There are two consequences to using such a setup. First of all, we no longer have
a single-shot measurement. Amplification introduces noise in the signal, meaning
that our oscillator will retrieve a measurement on + ε that is affected by random
fluctuating photon noise ε. In absence of systematic errors, those errors average
out ε = 0, and we can still produce meaningful estimates Ō M1 m om , but
the presence of large noise prevents us from determining the quantum state of our
system after the measurement – i.e., the measurement is no longer quantum limited.
The second consequence of indirect measurements is that we need a broader
framework to understand both the measurement statistics and the state of the quan-
tum system after a given measurement outcome. This framework is provided by
generalized quantum measurements, depicted in Figure 2.2.
(1) The system ψ is put in contact with the auxiliary quantum object φ.
(2) Both systems interact through some unitary evolution, Û .
(3) We measure the auxiliary object using a projective measurement Ô.
The generalized measurement or positive operator valued measurement (POVM)
associates measurement outcomes om to operators M̂n that are no longer projectors,
but still satisfy some completeness relation:
M̂n† M̂n = 1. (2.29)
n
19
the probability of collisions with electrons as well as the amount of energy that they
can exchange. When we combine both effects, we typically find an experimental fit
of the form
ρ ∼ ρ0 [1 + α max{T − T0,0}], (3.2)
with a temperature-dependent contribution and an intrinsic plateau ρ0 that depends
on the material and even on the specific sample.
The study of the temperature-dependent resistance in metals experienced a break-
through with the discovery of liquid helium by Kamerlingh Onnes in 1908. One of
the first applications of liquified helium was to study the conductivity of metals
(mercury, tin, and lead) under temperatures ∼ 4.2 K where the thermal contribution
to the resistivity should be negligible. In 1911, Kamerlingh Onnes submerged a
wire of mercury in helium and observed that the resistivity of the metal suddenly
dropped to zero. In other words, the previous plateau disappeared, ρ0 0, even
though nothing special had been done to improve the material’s purity and lattice
perfection. After these initial discoveries, further materials were shown to enter
this new superconducting phase at sufficiently low temperatures, summarized in
Table 3.1.
Later studies have shown that a superconducting material cannot be just charac-
terized as a conductor without resistance. As we will soon see, superconductivity is
an intrinsically quantum effect that has other unexpected consequences:
Fermi theory breaks down, and electrons join into stable bound particles called
Cooper pairs (Cooper, 1956). Because of the spin-statistics connection, the pair
of two bound electrons is a particle with a bosonic statistics, which condense into a
superfluid state.
The BCS theory provides a very elegant and also very straightforward frame-
work for studying the superconductor, with only a few parameters that account for
all the physics. The main parameter is the BCS or superconducting gap, usually
denoted by (k). The gap is the binding energy of every two electrons that form
a pair. It depends mildly on the electron’s momenta, and the smallest value (0)
explains many quantiative properties of the superconducting phase and the phase
transtion. For instance, as we increase the temperature, we can expect that processes
in which pairs are broken become relevant, destroying superconductivity. The quan-
titative answer is a bit more complicated, as the gap itself depends on temperature
√
(T ) 1.74(0) 1 − T /Tc , and the critical temperature Tc (0)/1.76kB is the
point at which pairing becomes energetically trivial.
The superconducting gap also explains some features in the interaction of the
superconductor with electromagnetic fields. A superconductor can absorb photons
through two different mechanisms. Low-energy microwaves in the range of 1–20
GHz can excite plasmons of the charged superfluid. These processes create quantum
excitations that behave very much like photons (see Chapter 5) or like artificial
atoms (see Chapter 6), and which we can route, confine, and operate using super-
conducting circuits. The second mechanism involves stealing a Cooper pair from
the condensate and breaking it into two separate electrons. The energy required
for this is h̄ω 2(0) 3.52kB T . For the case of the widely used material in
superconducting circuits, aluminum, this energy lays in the range of 100 GHz.
Therefore, we can suppress this type of event by sufficiently cooling our circuits
and isolating them from the environment, with filters that prevent the injection of
highly energetic photons.
The BCS theory has other important consequences, including studies of heat
capacity, impurities, quasiparticle excitations, Andreev states and normal super-
conductor interfaces. Overall, this theory applies very well to type I supercon-
ductors, and to some extent to type II, but it does not explain high-temperature
superconductivity.
Fortunately, we are not so much interested in complex superconducting mate-
rials or sophisticated excitations. Rather, we would like an effective model of
the superfluid condensate in the simple materials, Al or Nb, which are used
in the quantum circuit experiments. As explained by Gor’kov (1959), the BCS
model of superconductivity predicts an effective nonlinear theory for the con-
densate order parameter. This is the Ginzburg–Landau model or, in the simpli-
fied linear version that we introduce here, the macroscopic wavefunction model
(Orlando, 1991). Without many complications, this intuitive model will provide a
solid and approachable foundation to the engineering of quantum circuit in later
chapters.
This type of macroscopic accumulation of particles into the same quantum state is
what we expect from a Bose–Einstein condensate well below its critical temper-
ature. However, we are also allowing for this accumulation, which is typically a
property of a ground state, to also describe the dynamics of the collective system
in time, as it reacts to external perturbations from electromagnetic fields, currents,
etc. This is a conceptual extension that is only justified by the agreement with
experiments and the exact simulations of small systems.
The macroscopic wavefunction theory leads us to introduce new fields ns and θ ,
which respectively describe the charge density and, as we will soon see, the flow of
particles:
√
ψ(x,t) = N ξ (x,t) ns (x,t)eiθ(x,t) . (3.4)
We typically assume a constant and approximately uniform density of carriers
throughout most of the material, ns (x,t) n̄s . This assumption is approximately
valid in many situations because matter tends toward charge neutrality. It will not
apply when we consider the charge trapped on a capacitor or in a superconducting
1
i h̄∂t ψ = (−i h̄∇ − qs A)2 + qs v(x,t) ψ. (3.5)
2ms
This model is inspired by the Schrödinger equation for a charged particle moving
in an electromagnetic field with scalar and vector potentials v(x,t) and A(x,t),
respectively. The model introduces two effective parameters ms and qs to describe
the mass and charge of the Cooper pair. As we have seen, the charge is precisely
known:
qs = −2e = −2 × 1.60217662 × 10−19 C. (3.6)
However, ms = 2m∗e contains the effective mass of the electrons moving through
the solid lattice m∗e , which depends on the band structure and has to be determined
experimentally for each material.
This superfluid charge includes a very large background that compensates the
charge of the ions structuring the lattice of the metal or alloy. From the point of
view of circuit theory, it is more interesting to work with the superfluid current,
a vector field J(x,t) describing the flow of charges. The evolution of the electric
charge Q confined in a volume , is related to the supercurrent flowing across the
boundary of that volume ∂:
d
Q= ∂t ρd x = −
3
J · dn. (3.8)
dt ∂
Here n is the unit vector normal to the surface ∂ at each point of the boundary.
The same physics is described by the continuity equation:
∂t ρ = −∇ · J. (3.9)
As Fritz London conjectured, the superfluid current may be derived from the
Schrödinger equation (3.5) as a combination of the macroscopic wavefunction cur-
rent and the electromagnetic field:
∗ h̄ qs
J = qs × Re ψ −i ∇ − A ψ . (3.10)
ms ms
This combination explain the fluxoid quantization and the Meissner effect, as
described by London et al. (1935).
Note that we can obtain the electric current intensity of a circuit I by integrating
the charge current J(x,t) across any section S of any cable in the circuit:
I (t) =
J(x,t) · dS. (3.11)
S
1
− h̄∂t θ J2 + qs v. (3.14)
2ns
Notice the new constant, the isotropic London coefficient1 = ms /qs2 ns .
In the absence of currents, J = 0, (3.14) becomes the so-called phase-voltage
relation
qs
∂t θ − v. (3.15)
h̄
This relation is an obvious consequence of unitary evolution. For quasi-stationary
states, the wavefunction remains constant up to a global phase, determined by
the energy of the system ψ(x,t) = exp(−iEt/h̄)ψ(x,0). Since the energy of the
charged particle in a potential is E = qs v, we obtain ∂t θ = −E/h̄ = −qs v/h̄. The
problem with relation (3.15) is that it is not gauge invariant – it is only valid in the
Coulomb gauge – and it has been derived under the condition of no persistent
currents, J = 0. We have to complete our derivation to regard more general
conditions!
1 As mentioned before, superconducting currents shield magnetic fields out of the material. The London
coefficient is related to the penetration depth of magnetic fields into the superconductor, a fact that can also be
derived from this theory (Orlando, 1991).
2 This choice of path is one of the essential steps in working with any superconducting circuit, as we will see in
Chapter 4. However, in that chapter we will assume quasi-one-dimensional structures, where it is easy to
believe that such paths do exist, at least in the form xl = xl + x0 (1 − l) for l ∈ [0,1].
3 V and v are not the same observable. Only the former is gauge independent.
We use Stokes’s theorem to transform the second line integral into a surface integral
over the region S. The contour integral of the phase must produce a multiple of
2π × h̄/qs = −0 , as otherwise the wavefunction ψ would be discontinuous. This
results in the quantization equation
(J) · dl + B · dS + 0 × m = 0, with m ∈ Z. (3.25)
C S
plus the flux due to the induced supercurrents must be an integer multiple of the
magnetic flux quantum (3.21). This is a very powerful result, used by Deaver and
Fairbank (1961) to estimate 0 and demonstrate that the superconducting particles
qs = −2e are formed by two electrons.
There is a simpler derivation of fluxoid quantization that is of more interest to
us, and which is based on the gauge-invariant phase (3.16). This definition can be
reformulated as
qs
∇θ = ∇ϕ + A. (3.27)
h̄
Integrating around a closed loop, we find
2π
∇ϕ · dl = 2π × m + loop . (3.28)
C 0
Using the phase-voltage relation, this can be rewritten as a condition for the
flux differences along different segments of a superconducting circuit, or fluxoid
quantization:
∇φ · dl = 0 × m + loop . (3.29)
C
Figure 3.1 (a) Circuit schematics for a Josephson junction connected to an inten-
sity source. (b) The Josephson junction is made of two superconducting leads
(white) separated by a thin insulating barrier (gray). (c) We model the junction as a
barrier energy U0 , which is thin enough, d, that it allows some quantum tunneling
of Cooper pairs.
4 We can obtain a similar physics through other physical devices, such as constrictions, point contacts, and
normal interfaces. However, the theoretical picture is simpler in this case.
5 The one-dimensional model is sufficient for describing the type of small junctions that appears in typical
superconducting circuits. A more detailed model that takes into account the transverse dimensions is found in
Orlando (1991).
where the proportionality constant A is the area perpendicular to the junction – i.e.,
the cross-section size.
What do the solutions of the Schrödinger equation look like with these condi-
tions? Since far away from the insulator the current is fixed, we expect plane-wave
solutions such that ∂x ϕ ∝ J . More precisely, we write
√
ψGI (x) ∝ neikx , |x| > d/2. (3.33)
The sign and magnitude of the superconduction current may be derived from
the momentum k of the wavefunction as J = qs ns h̄k/ms . The momentum also k
determines the energy of the solution
h̄2 k 2 1 2 2
E= = J . (3.34)
2ms 2n
When the current is very small J 0, this energy will lay well below the barrier
U0 . Inside the insulator, we will have an equation of the form
ms 1
∂x2 ψGI = 2
(U0 − E)ψGI = 2 ψGI, |x| ≤ d/2, (3.35)
h̄ ξ
which can only be satisfied with exponentially decreasing or increasing solutions,
ψGI ∝ exp(±x/ξ ). The final solution then reads
The superfluid current is uniform across the circuit. Along the insulator
qs ∗
J = Re(−ψGI i h̄∂x ψGI ) = Jc sin[ϕR (d/2) − ϕL (−d/2)], (3.38)
ms
it is proportional to the Josephson junction’s critical current
qs h̄ ns
Jc = − ≥ 0. (3.39)
ms ξ sinh(d/ξ )
Currents above this value cannot be captured by the exponential solution, but instead
correspond to plane waves with E > U0 .
The previous derivation has shown us that small values of the current have a
nonlinear relation with the gauge-invariant phase jump δϕ = ϕR − ϕL across the
junction:
I = Ic sin(δϕ). (3.40)
As clearly explained by Orlando (1991), the first Josephson relation still holds in
presence of magnetic and electric fields, provided we still work with the gauge-
invariant phase.
Assume now that we establish a voltage difference V among the junction’s
leads. Since we work with metals, the voltage will be approximately uniform
along each lead. We then expect t a flux difference on both sides of the insulator
δφ(t) = φ(d/2) − φ(−d/2) 0 V (τ )dτ . Using the connection between flux and
phase (3.22), we obtain the second Josephson relation:
2π dV
δϕ = . (3.41)
0 dt
The Josephson relations combine into an equation connecting flux and current:
2π
I = Ic sin δφ . (3.42)
0
The dc Josephson effect is a consequence of this equation: a constant voltage bias V
induces an oscillating current due to the linearly growing flux:
2π
I (t) = Ic sin V t + δϕ(0) . (3.43)
0
However, (3.42) has more general implications, as a constituent equation, allow-
ing us to include Josephson junctions in the general theory of superconducting
circuits from Chapter 4. For instance, we can derive the inductive energy stored
in the junction as we switch on the voltage – and the flux – to a finite value:
t
E= I (t)V (t)dt. (3.44)
0
With the change of variables V (t)dt = dφ, we can substitute the expression (3.42)
and integrate between the initial and final value of the flux
δφ
2π Ic 0 2π
E= Ic sin φ dφ = − cos δφ . (3.45)
0 0 2π 0
From the point of view of superconducting circuit theory, the junction behaves as
a nonlinear inductive element. The inductance may be derived from the current–
voltage relation or from an expansion of the inductive energy just derived. In both
cases, we obtain a similar expression:
V 0 1
LJ = dI
= . (3.46)
dt
2π Ic cos(2πδφ/0 )
This nonlinear inductance and the constituent equation (3.42) will be repeat-
edly used in Chapter 4 when studying superconducting qubits, dc-SQUIDs, and
rf-SQUIDs.
4.1 Introduction
This chapter introduces a quantum mechanical theory of electrical circuits, devel-
oped in the spirit of the original work by Yurke and Denker (1984), and building
on the more pedagogical explanations by Devoret (1995). The chapter begins
discussing the temperature and power requirements needed to observe quantum
phenomena in electrical circuits. Section 4.2 recalls the three lumped circuit
elements – the capacitor, the inductor, and the Josephson junction that we use to
build most superconducting circuits, describing the classical equations that govern
them. Section 4.3 introduces an effective quantum theory of circuits, formulated
as a set of rules to transform a circuit network into the quantum Hamiltonian that
describes its dynamics. Sections 4.4–4.8 apply this theory to the most important
circuits in this book – a simple LC resonator, a transmission waveguide, a nonlinear
resonator, and different types of Josephson junction loops. You are strongly advised
to work out these examples and exercises on your own, to build a solid foundation
for later chapters where we use these circuits to engineer photons (Chapter 5),
superconducting qubits (Chapter 6), and light–matter interfaces (Chapter 7).
34
for pure states, or the master equations for mixed states. All this affects the way we
design, control, and operate superconducting quantum circuits.
Like all solid-state devices, electrical circuits are built from myriads of quantum
systems – a sea of electrons floating through a lattice of ions, covalently bound
molecules, oxidized metal, etc. A circuit will exhibit quantum phenomena depend-
ing both on how it is built and how it is operated. Common wisdom states that there
are three causes for a system to act in a purely classical way, hiding its quantum
nature:
(1) When we combine very large number of quantum objects into a many-body
system, the properties of the macroscopic object tend to be well defined, with
negligible quantum uncertainties. These objects can be safely described with
classical equations.
(2) Phenomena such as superpositions and entanglement are destroyed by inter-
action with the environment – e.g., by surrounding electromagnetic fields and
particles. The typical decoherence time in which this happens tends to be shorter
than the duration of common experiments.
(3) Preparing good quantum states requires cooling the system to a point in which
thermal fluctuations do not obscure quantum fluctuations. The temperature at
which this happens varies with the physical model.
Let us discuss these challenges in greater detail:
1 There is always a reservoir in contact with our system, even if that reservoir is just the electromagnetic field at
the temperature of the cosmic background radiation!
2 A good rule of thumb to remember this calculation is the equivalence between 20 GHz and 1 K, which makes
the excitation probability e−1 ∼ 36%. Every time we halve the temperature, we divide this probability by ∼3.
100
50
10
5
1
0.5
0 1 2 3 4 5 6
Figure 4.1 Thermal fluctuations in a quantum device. We study a device with either
two possible configurations separated by an energy h̄ω (dashed), or a ladder of
states with energies {0, h̄ω,2h̄ω, . . .} (solid). Introducing the ratio n̄ = h̄ω/kB T ,
we plot the probability that the system is found in an excited state. The solid line
follows the law exp(−n̄).
Figure 4.2 Circuit elements, together with the usual convention for current and
voltage. From left to right: generic element, capacitor, inductor, and battery. Note
that particles flow from the point of the circuit with the largest potential energy to
the area with the lowest potential energy. Hence, V and I have opposite directions.
Note that these conventions still work when the current is formed by negatively
charged particles, such as our Cooper pairs. Negatively charged particles move
along the direction of growing electric potential, from A to B, because their total
energy E = −2e × V is therefore decreased. However, if we understand the current
intensity as the time derivative of the charge I = dQ/dt, a flow of negative particles
from A to B is equivalent to a flow of positively charged particles from B to A.
Using these conventions, we will now analyze two linear elements, the capacitor
and the inductor, and then move on to the nonlinear Josephson junction. For all
three circuits, we will establish the voltage-current relations that will be used in
Section 4.3 to build the circuit’s dynamical equations.
4.2.1 Capacitor
We model a generic capacitor using two parallel plates, as shown in Figure 4.2. We
charge the capacitor by establishing a (positive) current that drains charges from
plate A and induces a positive charge in the opposite plate. As we know from our
physics textbooks, the charge imbalance creates a uniform electric field E directed
from the positive to the negative plate. This electric field is the gradient of a poten-
tial E = −∂V /∂x that decreases from the positively charged plate B down to
the plate A. Since E is proportional to the charge stored in the plates, we have a
proportionality relation:
t
Q 1
V = VB − VA = = I (τ )dτ, (4.7)
C C 0
where the constant C ∝ (area×plate distance) is the total capacitance of the circuit.
We frequently use this relation in its differential form:
dV
I =C = C φ̈. (4.8)
dt
4.2.2 Inductor
The inductance is the inertia that a conductor exhibits when we try to change the
current that traverses it. As sketched in Figure 4.2, a change in the current I passing
through a cable induces a change in the magnetic field B around it. By Maxwell’s
equations, an electric field will appear that opposes the change in magnetic field
∇ × E = −∂B/∂t. This electric field is equivalent to a potential drop parallel and
proportional to the time variation of the electric current:
dI
, or φ = LI .
V =L (4.11)
dt
The proportionality constant L is called the inductance, and (4.11) is the constitutive
relation for this lumped-element circuit.
Similar to the capacitance, the inductance relates the magnetic energy stored in
a circuit to the current intensity that passes through it. In an linear inductor L, we
need to spend an energy
1 1 2
Eind = LI 2 = φ (4.12)
2 2L
to raise the current intensity – or the electric flux – from zero to I .
3 Remember the factor 4 in the definition of E , which originates from the fact that the charging energy is a
C
concept that is also found in quantum dots and other electrical objects where the charge unit is an electron, not
a Cooper pair.
The inductance of a circuit element depends on its geometry and the material with
which it is constructed. In the case of a superconducting material, we may also need
to consider the kinetic inductance. This accounts for how the current flow is affected
by microscopic details, such as the penetration depth of the material and the actual
thickness of the superconductor film (Orlando, 1991). This source of inductance is
particularly interesting when we wish to enhance the magnetic interaction between
circuit elements (see Section 7.3.4).
convenience: The capacitor and the nonlinear inductor are just one entity, and the
loop formed by these two lumped elements does not exist.
The values of the Josephson critical current and of the Josephson energy depend
very much on the actual application. For instance, the work by van der Wal et al.
(2000) uses three junctions with critical currents in the range Ic ∼ 400−600 nA.
This implies Josephson energies EJ ∼ 13.2−19.8 nJ or EJ /h̄ ∼ 1 250−1 870 ×
2π GHz. However, you will also find many qubits with Josephson energies that
have lower values, in the 100s of GHz.
Figure 4.3 Quantization steps. (a) Open tree and node fluxes; (b) identify all
oriented branch fluxes; (c) define loop fluxes and use flux quantization to find
independent branches. The direction of the loop defines the orientation of the
enclosed surface S with which we compute the flux (3.26).
quanta trapped in the superconducting ring4 and the magentic flux induced
by external fields loop, computed with the right orientation of the enclosed
surface (cf. Figure 4.3c):
φx = n0 + loop . (4.17)
x∈loop
When all branches belong to the main tree, except for branch x = i → j, we
have
φx = n0 + loop + (φj − φi ), x = i → j ∈ C, (4.18)
to be compared with (4.16). Note that capacitors actually implement a physical
discontinuity in the conductor. When capacitive elements close a loop, n = 0,
and there will be no trapped fluxes.
(5) Branch currents: Compute the currents traversing each circuit element. For a
circuit on a branch b = i → j , the current will flow from j to i with a value I :
Capacitor Ij →i = Cb V̇b = Cb φ̈b See Section 4.2.1
Inductor I˙j →i = Vb /Lb ⇒ I = φb /Lb See Section 4.2.2
Junction Ij →i = IJ,b sin(φb /ϕ0 ) See Section 4.2.3
The branch flux φb is either (4.16) or (4.17), depending on whether b ∈ T or
b ∈ C. The capacitance Cb , inductance Lb , or critical current IJ,b are deter-
mined by the circuit element and do not depend on the branch orientation.
(6) Current conservation: Write down one equation per node, establishing that
the sum of currents coming in and out of the node add up to zero:
Ij →i = 0. (4.19)
j
This will produce one ordinary differential equation per node i in the tree, with
a general form
b 1 φb
si Cb φ̈b +
b
si φb + b
si Ib sin = 0, (4.20)
b∈capacitor b∈inductor
Lb b∈junction
ϕ0
4 This value should be ideally zero, if the system is properly annealed. Note that discrete changes in the trapped
fluxes along an experiment may cause decoherence in the quantum state of the circuit.
We keep the branch fluxes φb because they make the expression more com-
pact, but these values should ultimately be replaced with the independent node
variables φi and φ̇i . Note also that, in order to have a proper identification of
energies, we cannot rescale (4.20) or introduce new variables before deriving
the Lagrangian.
(9) Canonical variables: Define the node charges as the canonically conjugate
momenta of the node fluxes:
∂L
qi = . (4.24)
∂ φ̇i
Use these variables and a Legendre transform to construct a Hamiltonian
H (q,φ) = qi φ̇i − L(φ, φ̇), (4.25)
i
where we use (4.24) to write flux derivatives as functions of charges and fluxes,
φ̇(q,φ). With this Hamiltonian, the evolution equations read
dφ ∂H dq ∂H
= , =− . (4.26)
dt ∂q dt ∂φ
(10) Quantization: As in Section 2.1, replace the canonical variables (φi ,qi ) with
operators (φ̂i , q̂i ) that satisfy commutation relations
[φ̂i , q̂j ] = i h̄δij . (4.27)
Define a Hamiltonian operator Ĥ using (4.25) with the flux and charge
operators as variables. Due to the separation between capacitive and inductive
energies, this operator will be Weyl-ordered and Hermitian. As explained in
Section 2.1.3, this operator can be used to derive the Heisenberg equations
for charge and flux, or to study the Schrödinger equation for the circuit
wavefunction (4.6).
In this whole treatment, we have worked with fluxes φi on the nodes of the tree
because they are the canonical variables and because one can derive the general form
of the Lagrangian (4.22) from the current equations on these nodes (4.20). However,
this is not a unique way to work with the circuit: Any change of coordinates that
satisfies the constraints of our system will lead to a valid quantization procedure.
We can work with node charges, branch charges,5 or branch fluxes as variables
instead. We will discuss this last case because it illustrates the structure of the
resulting Hamiltonians and because it is convenient for some Josephson junction–
based circuits.
Nodes and tree branches are largely interchangeable. On every connected tree
there will be N + 1 nodes and N branches. When the tree includes voltage sources
or ground planes, the number of independent variables is identical in both repre-
sentations. Each node flux can then be written as a reference potential plus various
branch fluxes from on the tree. The change of coordinates from branch variables
φ b to node fluxes φ = Aφ b is performed by a matrix of ±1’s and 0’s. We use this
matrix to express the Lagrangian (4.22) as a function of the branch fluxes L(φ b, φ̇ b ).
In doing so, we must remember that closing branches are sums of branch variables
on the tree, due to flux quantization φc = n0 + − b φb . We may also need to
reduce the number of variables in φ b to account for voltage or current constraints.
The inductive terms remain similar in structure and do not need to be discussed.
The capacitive terms adopt a quadratic form:
1 T
Ecapacitive (φ̇ b ) = φ̇ b C φ̇ b + Cg VT φ̇. (4.28)
2
This includes a capacitance matrix Cbb and a voltage source term V connected
to the circuit through the coupling capacitances Cg . The capacitive term fixes the
definition of branch charges Qb = ∂L/∂ φ̇ b = C φ̇−Cg V, and gives the Hamiltonian
1
H (Qb,φ b ) = (Qb − Cg V)T C −1 (Qb − Cg V) (4.29)
2
φ2
+ b
− EJ,b cos(φb /ϕ0 ).
b∈inductor
2Lb b∈junction
5 These are, for instance, the variables used in the works by Yurke and Denker (1984) or Blais et al. (2004) to
derive equations for transmission lines and resonators.
4.4 LC Resonator
Let us apply the quantization formalism to the LC circuit shown in Figure 4.4a,
which consists of a superconducting resonator coupled to a constant voltage source.
The four lumped elements in this circuit are the resonator’s (i) capacitor C and
(ii) inductor L, (iii) the voltage source, and (iv) the gate capacitor that connects
it to the resonator, Cg . Using this information, please try to develop the quantum
Hamiltonian yourself, checking the result with the following derivation – one of
many equivalent ones.
In Figure 4.4b, we have constructed one possible circuit tree. The illustration
marks the three different nodes {φ0,φ1,φ2 }, three tree branches {φa ,φb,φc }, and a
closing branch φd . From left to right, the nodes include a ground plane – common
to the inductor, the capacitor, and the battery – the intersection between the inductor
and the two capacitors φ1 , and the cathode φ2 . We have also chosen directions for
all branches, defining the sign of potential differences:
φa = φ1 − φ0, (4.30)
φb = φ2 − φ1, (4.31)
φc = φ2 − φ0 . (4.32)
1 1
Ci φ̈i = (φi+1 − φi ) − (φi − φi−1 ). (4.38)
Li Li−1
N
1
N−1
1
L= Ci φ̇i2 − (φi+1 − φi )2 . (4.39)
i=1
2 i=1
2Li
The charges follow the expected definition qi = ∂L/∂ φ̇i = Ci φ̇i . The Legendre
transform produces the Hamiltonian:
N N
1 2 1
N−1
H = qi φ̇i − L = qi + (φi+1 − φi )2 . (4.40)
i=1 i=1
2Ci i=1
2Li
We can quantize the theory replacing (φi ,qi ) with operators satisfying [φ̂i , q̂j ] =
i h̄δij . The continuum limit of this theory is obtained by inserting the actual size of
the discretization:
2
N
x q̂i 2 x
N−1
φ̂i+1 − φ̂i
H = + , (4.41)
i=1
2c(xi ) x i=1
2l(xi ) x
defining the charge density ρ̂(xi ) = q̂i /x, and replacing sums with integrals to
obtain a quadratic field theory:
1 1
H = dx ρ̂(x) +
2 2
[∂x φ̂(x)] . (4.42)
2c(x) 2l(x)
These fields satisfy new commutation relations [φ̂(x), ρ̂(y)] = i h̄δ(x − y).
Figure 4.6 (a) Equivalent circuit for the charge qubit, with (b) the actual network
on the right. The superconducting island is the region contained between the two
capacitors, CJ and Cg . The nonlinear element EJ represents the channel by which
pairs can tunnel into the ground plane.
plane – both direct and mediated by the junction; (iii) a gate capacitor Cg ; and (iv)
voltage source V , which accounts for any external control on the island.
The quantization procedure copies the one in Section 4.4, using the same inde-
pendent variables and constraints. We simply replace the linear inductor with the
junction’s nonlinear current-voltage relation
d Cg
C φ̇1 − V̇ = −Ic sin(φ1 /ϕ0 ). (4.43)
dt C
The Lagrangian’s potential energy is now the nonlinear inductive energy of the
junction
1
L = C φ̇12 − Cg φ̇1 V + EJ cos(φ1 /ϕ0 ), (4.44)
2
with EJ = Ic ϕ0 . The charge operator has the same expression and the Hamiltonian
becomes a nonlinear function
1
H = (q − qg )2 − EJ cos(φ1 /ϕ0 ), (4.45)
2C
where the external potential sets the equilibrium charge qg = −Cg V in the ground
state of the qubit.
4.7 SQUIDs
We will now discuss the circuits for two different superconducting quantum inter-
ference devices (SQUIDs). These are superconducting circuits that make use of flux
quantization and Josephson tunneling. This feature makes the SQUID so sensitive
to magnetic fields that they can be used as high-precision magnetometers in biology,
medicine, or geology, among other fields. In a way, SQUIDs are the peak of super-
conducting circuit technology, a vast field of research6 that predates and enables
modern-day quantum circuits. In this book, we do not have space to go deep into
this field, but will frequently rely on the SQUIDs to implement superconducting
qubits, add tuneability to circuits, and build quantum amplifiers.
4.7.1 rf-SQUID
The rf-SQUID consists of a Josephson junction that is shunted by a linear inductor,
as shown in Figure 4.7. The device is called a radio-frequency SQUID because
it relies on the ac Josephson junction effect. It is a cheap magnetometer and also
6 See, for instance, The SQUID Handbook by Clarke and Braginski (2004).
one of the earliest superconducting qubits, which still finds applications in modern
quantum annealers (see Section 9.4.1).
Compared to previous devices, the rf-SQUID introduces new formal challenges
that were not discussed in circuit quantization: (i) We have a loop without reference
potential; (ii) the SQUID couples to external current sources via a mutual induc-
tance M; and (iii) the external currents induce a net magnetic flux in the loop = 0,
which affects the dynamical equations.
The equation for the node variable φ0 balances the currents on the junction and the
linear inductor. The current in the inductor is partially caused by the self-inductance
of the branch φL /L, and partially by the mutual inductance. According to our sign
conventions, the downward external current I induces an upward moving current
on the SQUID −MI , so as to reduce the total magnetic interaction between both
circuits. Hence,
1
CJ φ̈J + Ic sin(φJ /ϕ0 ) =
φL − MI . (4.46)
L
We replace the branch fluxes φJ and φL with their definitions in terms of flux nodes,
using the fluxoid quantization. Assuming no trapped vortices and with our choice of
tree, we have φL = φ1 −φ0 and φJ = − + (φ0 − φ1 ) = −( + φL ). Note that the
total magnetic flux is created both by the external currents I and by the current
supported by the loop itself φL /L. Far away from the regime of bistability, we can
neglect this last contribution and approximate ˙ = 0. The equation on the lower
node
d φ0 − φ1 − φ1 − φ 0
CJ (φ̇0 − φ̇1 ) = −Ic sin + − MI (4.47)
dt ϕ0 L
is identical to that of the one on the upper node, up to a global sign. Both equations
have the form 4.21 with the Lagrangian
Figure 4.7 Equivalent circuit for the rf-SQUID qubit in full version (a) or restricted
to the SQUID (b). A Josephson junction (CJ ,EJ ) is shorted by a large inductor L
that may couple to an external magnetic field generated by a current source. The
mutual inductance M is introduced in the text.
1 1 φ0 − φ1 −
L = CJ (φ̇1 − φ̇0 )2 − (φ1 − φ0 )2 − M(φ1 − φ0 )I − EJ cos ,
2 2L ϕ0
(4.48)
1 1 2 1
H Q2L + φL − MI φL + EJ cos (φL + ) . (4.49)
2CJ 2L ϕ0
If we look at the inductive terms, the role of MI is to shift the flux φL at which the
minimum of energy is reached. However, this shift MI L will have to compete with
the tendency of the nonlinear potential to trap the flux around φL ∝ 2πϕ0 + . In
a regime of maximum frustration, where /ϕ0 ∼ π , we can obtain an energy land-
scape with two local minima that can be applied to create a qubit, as discussed in
Section 6.4.
4.7.2 dc-SQUID
The “interference” in the word SQUID is revealed by circuits with multiple junc-
tions, such as the dc-SQUID from Figure 4.8a. In this lumped-element model, there
are two junctions connected in parallel, forming a loop. The incoming supercurrents
split into two distinct paths, interfering at the exit port. The outcome of this inter-
ference and the total dc current that passes through depend on the magnetic flux
injected through the SQUID.
in out
Using flux quantization and our choice of main tree, the upper and lower branch
fluxes are φu = φ1 − φ0 and φd = + φ0 − φ1 , with the total external and trapped
fluxes . We assume identical junctions and use the identity sin(a) ± sin(b) =
2 sin a±b
2
cos a∓b2
to write
φd − φu φd + φu
CJ (φ̈d − φ̈u ) = I − 2Ic sin cos x. (4.50)
2ϕ0 2ϕ0
This element is used to make other quantum circuits tunable. For instance, we
can upgrade a charge qubit (Section 6.2), replacing its junction with a SQUID to
adjust the qubit’s spectrum of energies. We can also use the SQUID in place of a
linear inductor, such as in an LC resonator or transmission line, to modulate the LC
resonance in time and create a parametric amplifier (see Section 5.5.6).
dc-SQUIDs can be used to measure magnetic fields with high precision. The crit-
ical current Ic () can change from 0 to hundreds of nanoamperes over the span of a
flux quantum 0 ∼ 2 × 10−15 T · m−2 . This means that if we can detect changes in
Ic (), we can use a SQUID to measure a magnetic field (or flux) with sensitivities
close to attoteslas. In order to measure Ic , we can rely on the dynamics of a current-
biased SQUID:
Figure 4.9 Washboard potential of the dc-SQUID (4.54) (a) below and (b) on the
verge of switching to a voltage state.
φ−
Eind = −EJ () cos − I φ− (4.54)
ϕ0
For small biases, this function has the shape of a washboard potential (cf.
Figure 4.9a). The infinitely many local minima are metastable configurations of
the flux φ− with V ∝ φ̇− = 0. When the bias current I exceeds the critical value
Ic () = 2Ic cos(/2ϕ0 )/ϕ0 , all such minima disappear (cf. Figure 4.9a), and the
flux “flows” down the slope, linearly accelerating in time.7 This translates into a
measurable potential difference between the SQUID’s ports, which can be detected.
In the world of quantum circuits, SQUIDs are used to measure the current state
of other circuits. For instance, a flux qubit has two states |0 and |1 that correspond
to different orientations of the supercurrent. These states can inject different values
of the flux 0 and 1 into a nearby SQUID. We can bias the SQUID to a point in
between the critical currents for those two states, say Ic (0 ) < I < Ic (1 ). When
doing so, the SQUID will either remain the same or switch to a voltage state, depend-
ing on whether the qubit is in the |1 or |0 states. These are almost projective
measurements, limited in precision by noise in the SQUID and the finite duration
of the measurement.
7 This is a formally ill-defined problem, which stabilizes to an asymptotically stable solution once we consider
the resistance of the junctions, or add a resistor in parallel to the circuit.
Figure 4.10 (a) The qubit is a loop with three Josephson junctions, two identical
(1,2) and a smaller one (3). (b) The equivalent circuit where we focus only on the
branch fluxes.
expands over states with a well-defined value of the node phases |ϕi . This
continuous representation transforms the Schrödinger equation into a partial-
differential equation for (ϕ1, . . . ,ϕN ), which can be solved analytically in
some important cases, such as the transmon qubit (Section 6.3).
Number representation: We construct the Hilbert space using the eigenstates of
the charge operator q̂, which are states with a given number of excess Cooper
pairs q̂ = −2en̂. These states form a discrete basis {|n , n ∈ Z}. Operators such
as n̂, q̂, or the Hamiltonian H are represented as infinitely large matrices, which in
most cases can be truncated to a reasonable cutoff |n| ≤ nmax , either for analytical
approximation – see the charge qubit in Section 6.2 – or for some numerical
diagonalization method – see the three-junction flux qubit in Section 6.4.3.
In each representation, we have one operator that is diagonal and simple, but
we ignore the conjugate observable. In the number representation, the charge
9 Our notation is consistent with Pegg and Barnett (1989) and with the original definition of the phase operator
by Susskind and Glogower (1964).
Notice that we have safely replaced the ϕ̂ operator with its representation, the phase
degree of freedom ϕ. The previous equation is solved by the identification n̂ = i∂ϕ
and we can derive the phase eigenstates from the equation
m|ei ϕ̂ |ϕc = m + 1|ϕc = eiϕc m|ϕc , (4.65)
which implies
1 iϕc m
|ϕ = √ e |m . (4.66)
2π m∈Z
This discussion completes a dictionary for working with flux and charge, sum-
marized in Table 4.2. Using these tools, we can take the Hamiltonian of a Josephson
Continuous Discrete
Variable ϕ∈R n∈Z
Charge q = −i2e∂ϕ q = −2e n n |nn| = −2e n̂
0
Flux φ̂ = 2π ×ϕ ei2π φ̂/0 = n |n − 1n|.
junction (4.45) and write it in different representations. In the phase variables, the
state is given by a periodic wavefunction (ϕ,t) that evolves in time with a partial-
derivative equation:
∂ ∂ 2
i h̄ = −4EC 2 − EJ cos(ϕ). (4.67)
∂t ∂ϕ
In Chapter 6, we solve analytically a stationary version of this equation, but in
practical applications with two or more variables, or with additional terms, it is not
possible to do it. We could in principle address those complicated examples using a
method such as finite differences. However, a simpler approach is to write the same
Hamiltonian in the number representation. For instance, in the Josephson junction
case (4.67) we have
H = [4EC n |nn| − EJ (|n + 1n| + |nn + 1|)] . (4.68)
n
In many real-world applications, where EJ /EC ≤ 100, the probability that a state
with large number of Cooper pairs is occupied is very low. This means we can
reduce the complexity of the Hamiltonian, truncating the matrices and vectors to
a smaller subspace, of |n| ≤ nmax ∼ 10 or less. In some cases, such as the charge
qubit 6.2, this truncation is even more dramatic, and we end up in a two-dimensional
subspace, such as n ∈ {0,1}.
Exercises
4.1 Discuss the differences between the thermal state ρ∞ = 12 |00| + 12 |11|,
and the quantum superposition |ψ = √12 (|0 + |1). What is the density
matrix ρq associated to the latter state? How can we distinguish both states
experimentally if we have access to measurements of the σ z and σ x operators?
4.2 Using the commutation relation [φn,qm ] = i h̄δnm, prove (4.59). Hint: Write
down the Taylor expansion of the exponential, and use the commutator on
each term.
4.3 Reproduce the derivations of all the circuit models and effective Hamiltonians
from Sections 4.4 to 4.8.
4.4 Using (4.1), show that the energy of the LC circuit H = 12 CV 2 + 12 LI 2 is
conserved, that is, dH /dt = 0. What happens if you add a resistor in parallel
to the circuit?
4.5 Draw a circuit of two capacitively coupled LC resonators. The circuit should
be the result of replacing the battery in Figure 4.4 with another resonator. Find
out the quantum model for this circuit. Find out the canonical variables Q̃n and
φ̃n that diagonalize the Hamiltonian, and bring it to a form
Q̃2 1
Heff = n
+ Ln φ̃n2 . (4.69)
n
2Cn 2
4.6 Similar to Problem 4.5, derive the circuit and the Hamiltonian for two identical
charge qubits connected via a capacitor. Show that, in the limit of weak cou-
pling, the interaction between both qubits is modeled by the product of charges
∝ q1 q2 from both qubits.
4.7 Compute analytically the value of the bias current at which the washboard
potential from Figure 4.9 loses all its local minima.
4.8 Think how you would write the Hamiltonian for the 3-JJ flux qubit (4.58) as
a matrix in the number-phase representation. Remember to use the decompo-
sition of trigonometric functions into exponentials that have a simple matrix
representation cos(ϕ) = 12 (eiϕ + e−iϕ ). Also realize that eiϕ+ eiϕ− will translate
into a Kronecker or tensor product of two matrices A ⊗ B, each representing
the action of the exponential on a different Hilbert space – such as A for eiϕ+
and B for eiϕ− .
5.1 LC Resonator
5.1.1 Energy Quantization and Photons
Figure 5.1a shows a very small LC circuit made of two superconducting elements: a
very visible interdigitated capacitor with many fingers, shunted by a linear inductor
in the form of a small superconducting loop. The LC circuit is so small that the
lumped-element circuit approximation is well justified. To be more precise, we can
tune global properties, such as the total charge, flux, voltage, or intensity, but we
cannot control these properties’ distribution across the circuit, because this “struc-
ture” or mode is fixed in a state of minimum energy. If we want to alter the mode
structure, we need to introduce changes over distances d ∼ 100 μm. The energies
associated to these tiny details is hundreds of GHz, well above the superconducting
gap and far from our experiments’ operation regime.
63
0.2 mm
The resonator in Figure 5.1a not only a very good example of the lumped-
element approximation, it even resembles the graphical representation of the LC
circuit, shown in Figure 5.1b, which we introduced in Section 4.4. As discussed
then, charge, flux, voltage, and current all follow a similar oscillatory equation,
φ̈ = −ωLC 2
φ, solved as φ(t) ∼ cos(ωt + ϕ) with a single frequency that depends
on the electrical properties of the circuit:
1
ωLC = √ . (5.1)
LC
Because this circuit only supports a fixed spatial distribution of charge and flux,
it only has a single mode or frequency, which is why it is also called a zero mode
resonator.
The LC resonator is conceptually similar to the ideal oscillators that Planck used
in his model of the black-body’s electromagnetic radiation (see Section 2.1). Fol-
lowing Planck’s and Einstein’s ideas, we could propose that our superconducting
resonator is a quantum mechanical object that only exchanges discrete units of
energy with the environment – our quanta or microwave photons. The states of the
resonator would then be labeled by the number of photons that it stores |n, and the
total energy would be a sum of quanta En ∼ h̄ω × n, n ∈ {0,1,2, . . .}.
This qualitative reasoning is essentially correct. It can be justified using Dirac’s
quantization and the quantum circuit models from Section 4.3. The “photons” that
we find in the superconductor have a mixed nature: They are excitations of the
electromagnetic field that are accompanied by waves of charge and current on the
surface of the metal. This hybrid nature would make it more appropriate to call
them microwave plasmons, but we keep the name “photon” to distinguish these
plasmonic excitations from others that have a more anharmonic spectrum, which
we use to build qubits.
The microwave photons stored in a resonator do not interact with each other,
at least not until we reach high powers that excite nonlinear contributions to the
inductance and capacitance. Also, when placed on extended larger systems, such
as a superconducting waveguides (cf. Figure 5.1c), the photonic nature of these
plasmons evidences in a propagation without dispersion, at a fraction of the speed
of light.
However, the microwave photons do have a matter-like component. This com-
ponent allows us to control the properties of the photon – frequency, speed, wave-
length, etc. – by a proper design of the circuit: capacitance, inductance, thickness
of the material, etc. Moreover, the matter component in the wave enables the strong
interaction between these microwave photons and other circuits, atoms, molecules,
NV-centers, etc., opening regimes of extreme light–matter interactions that were
previously unreachable in quantum optics.
Unfortunately, working with microwave photons has some drawbacks. In particu-
lar, the small energy contained in a microwave photon makes it very difficult to build
a microwave photodetector or photon counter. The lack of such devices prevents
us from directly verifying the quantization of the cavity’s excitations En = h̄ωn, a
fact that only becomes evident when we introduce other, more matter-like nonlinear
circuit elements, such as qubits.
This method no longer works when the Hamiltonian depends on time, such as when
the resonator is connected to an oscillating potential V (t) in (5.2). This problem is
best analyzed in the Heisenberg picture (2.13).
The Hamiltonian for a resonator coupled to an oscillating potential V (t) reads
Ĥ = h̄ωa † a + i h̄(t)(a † − a) + Eglobal (t). (5.8)
√
The linear displacement h̄(t) = h̄/2ZCg V (t)/C contains the external poten-
tial. The energy shift Eglobal (t) ∝ V (t)2 will be ignored, since it commutes with all
operators. The Heisenberg equations for the canonical and Fock operators are linear,
first-order differential equations:
d
φ̂ = 1
[ q̂ − Cg V (t)], dâ
dt C
⇔ = −iωa + (t). (5.9)
d
dt
q̂ = − L φ̂,
1 dt
The Fock operators experience a phase and a complex displacement α(t):
t
−iωt
â(t) = e â(0) + e−iω(t−τ ) (τ )dτ = e−iωt â(0) + α(t), (5.10)
0
and provide the solution for flux and charge:
h̄Z +iωt † h̄
φ̂(t) = (e a + H.c.) + φext (t), q̂(t) = (ie+iωt a † + H.c) + qext (t),
2 2Z
√ (5.11)
with α = [φext (t) + iZqext (t)]/ h̄Z.
In absence of external drive, the solutions in the phase space (φ̂, q̂) are elliptical
orbits with angular frequency ω:
φ̂(t) = cos(ωt)φ̂(0) + sin(ωt)Z q̂(0), (5.12)
1
q̂(t) = cos(ωt)q̂(0) + sin(ωt) φ̂(0). (5.13)
Z
As in the classical theory, the impedance is the ratio between the positive (or nega-
tive) frequency components of the flux and charge, or voltage and current intensity.
This can be seen from a Fourier transform, or directly, taking the positive frequency
terms from (5.11):
√
φ̃(+ω) h̄Z/2 a † (0) Ṽ (+ω)
=√ = = −iZ. (5.14)
q̃(+ω) h̄/2Z a (0)
†
I˜(+ω)
Here and throughout this book, we often write the expected value of a non-
Hermitian operator a † . Even though a † is not an observable, it can be decomposed
as the addition of two measurable quantities:
2 2Z
a † (t) = φ̂ − i q̂. (5.15)
h̄Z h̄
Therefore, the expectation value of a † , a, a † a, or even higher moments a 2 a † , etc.,
can be inferred from the composition of two separate measurements. This will be
relevant in later sections, where we study the amplification of light and the simul-
taneous measurement of quadratures (Section 5.5.7).
4 Note the word conventional. A microwave photon can still trigger other elements, such as superconducting
circuits Romero et al. (2009a,b), as a basis for microwave photodetection.
Figure 5.2 (a) A coplanar waveguide. The left and right superconducting planes
are grounded to a fixed potential while the charge and electromagnetic waves run
through the middle section. (b) Other planar waveguides, such as the stripline
(top), differ from the coplanar waveguide (bottom) on the position of ground
planes relative to the main conductor. (c) Equivalent circuit for the waveguide (see
Section 4.5).
inductive energy describing the “inertia” of the conductors to change the current.
Introducing the capacitance and inductance per unit length, c(x) and l(x), we
derived two alternative models: a discrete theory (4.40) and a continuous field
theory (4.42). We will now diagonalize and study these models in three different
scenarios with different boundary conditions and line terminations.
The line has length d and is divided into N segments with the same capacitance and
inductance, C = cx and L = lx. We assume that the line closes onto itself, so
that φ̂N+1 = φ̂1 .
The second term in this model is a quadratic form, φ T Bφ/2L, with a nonnegative,
symmetric matrix B (cf. Section A.1.1). Following the mathematical tools from
Appendix A.2, we diagonalize the matrix B and obtain N normal modes to write
the Hamiltonian as a sum of N independent oscillators:
† 1
Ĥ = h̄ωk b̂k b̂k + . (5.17)
k
2
There are a total of N modes, which we identify with left- and right-moving waves,
k < 0 and k > 0, labeled by the integer numbers n ∈ (−N/2,N/2]. We can recon-
struct the flux and charge fields along the waveguide’s positions xm = x × m as
h̄
φ̂(xm ) = uk (x)b̂k + uk (x)∗ b̂k† , (5.19)
k
2cωk
h̄cωk
q̂(xm ) = x iuk (x)∗ b̂k† − iuk (x)b̂k ,
k
2
√
using the normalized eigenmode wavefunctions ux (x) = eikxm / d.
For small momenta, the dispersion relation is approximately linear:
1 1
ωk v|k|, v=√ , |k| . (5.20)
c0 l0 x
The dispersion relation, the speed of light v, and the operators φ̂ and q̂/x are all
defined in terms of intensive quantities and scale-independent normal modes. As
anticipated in Section 4.5, this facilitates taking the limit x → 0, where flux and
charge are replaced with continuous fields [φ̂(x), ρ̂(y)] = i h̄δ(x − y). The new
Hamiltonian
d 2
1 1
H = dx ρ̂(x) +
2
∂x φ̂(x) (5.21)
0 2c 2l
corresponds to a massless Klein–Gordon theory. These fields satisfy a wave equa-
tion with periodic boundary conditions φ̂(x + d,t) = φ̂(x,t):
1
∂t −
2
∂ φ̂(x,t) = 0,
2
(5.22)
c(x)l(x) x
whose stationary solutions (ωk2 − ∂x2 )uk (x) = 0 are the continuous version of the
discretized eigenmodes that we found before.
(a) (b)
k xd
Figure 5.3 (a) Dispersion relations for a transmission line. The dashed line plots
the relations for x = 1,1/2,1/4, and the solid line plots the limiting relation for
x = 0. When the transmission line has a finite length d = 6, the momenta and
frequencies are discrete (circles). (b) First flux modes for an open trasmission line
of finite length.
isolated line. In this limit, no charge can escape the line and the currents must vanish
at the boundaries. The zero-current property translates into Neumann-type boundary
conditions for the flux ∂x φ|x=0,d = 0. The Hamiltonian (5.21) with these boundary
conditions is diagonalized by an infinite set of standing waves:
πn
2
u (x) =
(n)
cos x , n ∈ N = {1,2,3, . . .} (5.23)
d d
As show in Figure 5.1c, we can use this fact to couple capacitively the resonator
to a semi-infinite transmission line. The coupling, whose strength depends on the
capacitors C1 or C2 used, is our means to control the resonator, injecting photons or
measuring its internal state (see Section 5.5).
If we only need one control access, we may want to connect the other boundary to
a ground plane, as shown in Figure 5.4b. This connection replaces one of the Neu-
mann boundary conditions with a fixed value for the voltage at that point φ̂(0) = 0,
(a)
(b)
(c)
Figure 5.4 First φ0 (x) and second φ1 (x) eigenmodes of (a) λ/2 and (b) λ/4 trans-
mission line resonators. (c) A transmission line resonator with a SQUID becomes
an adjustable resonator. The external flux affects the SQUID’s inductance, the
lowest energy eigenmodes, and the effective length of the resonator d(). Note that
the drawing is not on scale: A typical SQUID is much smaller than the waveguide.
while the other condition remains the same ∂x φ̂(d) = 0. The new set of modes that
satisfies these boundary conditions are the sinusoidal functions
2 π (2n − 1)
u =
(n)
sin x . (5.25)
d 2d
The length of the waveguide is a quarter of the wavelength of the fundamental mode,
d = λ1 /4. Such λ/4 resonators are interesting because close to the edge x = 0 the
current ∂x φ/ l is maximal. This is beneficial for coupling inductive elements to the
resonator and to place inductive elements that tune the properties of the cavity, as
we will now see.
5 Does anyone remember old radio receivers with a frequency selection wheel that adjusted a big capacitor? In
this device, the tuning wheel changed the opposing surface between two metallic plates, adjusting the
capacitance.
If we do not want to use a zero-mode resonator, we can place the dc-SQUID inside
a transmission line, as shown in Figure 5.4c. When the SQUID is much smaller than
the wavelength of the photons that we excite, we can treat it as a local perturbation
in the Klein–Gordon model (5.21):
d
1
Ĥind φ̂(x)2 δ(x − x0 )dx + O(φ 4 ). (5.27)
0 2LJ ()
This term affects the eigenfrequencies and eigenmodes of the guide. In particular,
the normal modes become discontinuous around the SQUID: The flux jumps by an
amount δφ ∼ ∂x φ × δx, which would correspond to extending the waveguide by an
additional length δx = LJ / l0 – a ficticious length that remains “folded” inside the
SQUID.
The inductance of a junction can be much larger than that of typical coplanar
waveguide. The change in the length δx/λ can be comparably large, affecting the
location of the nodes and antinodes of the resonator. If we connect other circuits –
superconducting qubits, resonators, etc. – to those points, we can use a dc-SQUID
to bring them in and out of resonance, coupling and decoupling the resonator from
them (Forn-Díaz et al., 2017).
The change in resonator length can also be much faster than the time d/v for a
photon traverse the waveguide (Sandberg et al., 2008). In this regime of driving,
it is as if the “mirrors” of the resonator moved at close to the speed of light. Such
conditions simulate a gedanken experiment known as the dynamical Casimir effect,
which results in the production of entangled photon pairs out of the vacuum. This
experiment would be impractical with real mirrors and cavities, but has been repeat-
edly demonstrated by Wilson et al. (2011) and later works.
Finally, in Section 5.5.6 we will see that a periodic modulation of the inductance
can be used to achieve an effect known as parametric amplification. When we send
a quantum signal to this resonator, it is reflected back with one or two quadratures
that are enlarged or shrunk. Quantum amplification is a very strong and robust
mechanism, instrumental for verifying the quantum properties of a photonic field
(see Section 5.5.7).
quantum states – we ignore high-order moments that are hard to measure – as well
as the theoretical simulations of practical circuits.
Fock states are simple and useful, but difficult to prepare. If we drive a linear
resonator with an external field, the coupling (t)(â † + â) spreads the wavefunction
over all Fock states. To create a state of one photon, we need a single photon source:
a saturable quantum device that is always excited to the same state, and which
always decays emitting just one photon. The word “saturable” means that the device
can take a finite amount of energy: typically, it will jump from some ground state |g
to the same excited state |e, and it will stop there. When we stop driving the source,
the photon source decays back to |g, emitting a single photon. These nonlinear two-
level circuits are the superconducting qubits that we introduce in Chapter 6.
Using qubits, we can play some magic tricks. For instance, we can prepare the
qubit in a superposition of ground and excited state α |0 + β |1, so that when
it exchanges energy with the microwave resonator it implements the operation
(α + β â † ). Multiple iterations of this trick lead to arbitrary superpositions of Fock
states, as demonstrated by Hofheinz et al. (2009). Moreover, the same qubit can also
be used as photon counter, to reconstruct the full wavefunction of the microwave
resonator – see Section 7.3.2 and works by Schuster et al. (2007) and Hofheinz
et al. (2009).
∞
e−βH
ρ= = e−β h̄ωn (1 − e−β h̄ω ) |nn| . (5.30)
Z(β) n=0
This ensemble gives a nonzero probability of finding photons in the resonator pn>0
and a nonzero average occupation n̄:
pn>0 = cρnn = 1 − ρ00 = e−h̄ω/kB T , (5.31)
n>0
1
n̄ = a † a = . (5.32)
eh̄ω/kB T − 1
7 Also on other circuits, such as superconducting current loops (van der Wal et al., 2000), and on trapped ions,
optical lattices, etc.
Introducing the rotated quadrature x̂φ = √1 (â † eiφ + âe−iφ ), we find two observables
2
with opposite behavior
x̂θ (r) = e−r x̂θ (0), p̂θ (r) := x̂θ+π /2 = e+r p̂θ (0). (5.39)
The operator p̂θ and its uncertainty p̂θ are both amplified with a gain factor
√
G = e+r , while the operator √ x̂θ (r) and its variance x̂θ (r) are both contracted
−r
by the inverse factor e = 1/ G.
Experimentally, this means that measuring x̂θ (0) on the squeezed vacuum
Usq-1 (r) |0 produces
√ results with an uncertainty below the standard quantum limit
x̂θ = e−r / 2. Squeezing is measured in decibels rdB := 10 × log10 (Onew /Oold ).
8 Instead of studying how the states are transformed, we study how the observables that we want to measure are
transformed.
A value rdB ∼ 3 means a 50% reduction in uncertainty e−r ∼ 0.5. Not too long ago,
such values were considered the state of the art in quantum optics – for instance,
the Laser Interferometer Gravitational-Wave Observatory (LIGO) uses between 2.7
and 3.2 dB of squeezing in their interferometer. These days, much larger values of
10−12 dB (0.1–0.06 reductions) are available in the superconducting lab, thanks to
the parametric amplifiers described in Section 5.5.6.
The second motivation for engineering squeezed states is to create entanglement
and correlations between two or more bosonic modes (Laurat et al., 2005). In the
typical two-mode squeezed state, we compress a quadrature that is a linear com-
bination of two oscillator modes b1 + b2 , using the general two-mode squeezing
Gaussian operation:
Usq-2 (r) = exp rei2θ bj† bi† − r −i2θ bi bj . (5.40)
Applying this transformation onto a thermal state, we achieve squeezing and ampli-
fication of the joint quadratures x̂θ± = 12 (b̂1† ± b̂2† )e−iθ + H.c. and p̂θ± = 2i (b̂1† ± b̂2† ) +
H.c. similar to before:
x̂θ± (r) = e∓r x̂θ± (0), p̂θ± (r) = e±r x̂θ± (0). (5.42)
We will discuss the creation of two-mode squeezing with linear amplifiers and these
gain/contraction relations in Section 5.5.6.
The two-mode squeezed state is an entangled state because it cannot be written
as product of two independent states for each of the modes ρsq-2 = ρ1 ⊗ ρ2 . This is
qualitatively appreciated when we write the Schmidt decomposition of the state in
√ n/2
the Fock basis |0r := Usq-2 (r) |0 = 1 − λr n λr |n,n, with λr = tanh(r)2 .
We can also compute the reduced density matrix that results from tracing out or
ignoring one of the modes ρ1 = tr2 (|0r 0r |) = (1 − λr ) n λnr |nn|. The von
Neumann entropy of this matrix is a measure of the entanglement between the
modes S = −tr(S log(S)) ∼ 2r and diverges with the amount of squeezing r.
It is possible to create entangled states of multiple degrees of freedom. In particu-
lar, the transmission line is an example of multimode squeezed state. From (5.19), we
see that the fields {b̂k, b̂k† } are macroscopic superpositions of quadratures at different
positions. Thus, while the vacuum state may seem a rather trivial state with no
photons b̂k | = 0, it is in fact a highly correlated state from the point of view of
the field operators along the line, {φ̂(x), q̂(x)}. If we try to estimate the uncertainty
of the local observables, we will find strong divergences:
h̄eik(x−y)
φ̂(x)φ̂(y) − φ̂(x) φ̂(y) = bk bk† , (5.43)
k
2c0 xω k
both in the infrared and ultraviolet limit.9 This manifestation of entanglement in the
1D field theory is but one example of other quantum field theory concepts that can
be explored using superconducting circuits – particle localization, propagation of
entanglement and causality, vacuum energy and Casimir effect, etc.
9 Infrared because the denominator approaches zero for k = 0, and ultraviolet because the fraction is
proportional to 1/n for |n| = 1, . . . ,N/2, and the sum diverges logarithmically with the discretization size.
10 Notation follows Adesso et al. (2014) and Olivares (2012).
ij := {R̂i − ri , R̂j − rj }ρ = R̂i R̂j + R̂j R̂i ρ − 2 R̂u R̂j ρ , (5.49)
are uniquely determined by the first and second moments of the quadratures.
As we anticipated, all single- and multimode coherent states |α are Gaussian
states.
√ Their Wigner functions look like displaced vacua: They are centered on
d = 2(αre,αim ) and – because they are minimal uncertainty states – they have the
same width on all directions. Figure 5.5a and b illustrate the Wigner function for a
vacuum state and for a coherent state at α = 5/2.
A squeezed state is a Gaussian state with unequal widths along two or more
mutually orthogonal directions. Figure 5.5e shows the Wigner function of a single-
mode squeezed vacuum state – state U (r) |0 with θ = 0 in (5.37). As expected,
the quadrature x̂ is reduced by a factor e−r = 1/2, while the canonically conjugate
momentum is enlarged by the inverse factor. The result is a Gaussian √ with elliptical
equiprobability contours, and a minimum width x̂ = 1/2 2 below the Heisen-
berg limit.
The family of Gaussian states includes also thermal states. Moreover, all Gaus-
sian states can be written as a finite temperature state of one quadratic model,
ρ = e−βHeff /tr e−β Ĥeff with Ĥeff = 12 Hij (R̂i − ri )(R̂j − rj ). The process of determin-
ing the model Ĥeff and the temperature β = 1/kB T is equivalent to finding the
simplectic transformation that diagonalizes .
Gaussian operations are those that preserve the Gaussian nature of a state. This
includes all linear transformations between modes R = U † RU = OR generated
by quadratic Hamiltonians U = exp(−iHeff t). Examples of those coherent transfor-
mations are the single-mode and two-mode squeezing operations from Section 5.4.5.
However, Gaussian states are also preserved by incoherent operations that change
the area covered by the Wigner function, enlarging the uncertainty of one or more
quadrature. We will see examples of those operations when we study the coupling
of a Gaussian model to a linear environment in Section 5.5.4.
(a) (b)
(c) (d)
(e)
Figure 5.5 Wigner functions of (a) the vacuum state |0, (b) a coherent state
|α = 5/2, (c) a Fock state |1, (d) a Schrödinger cat ∝ |5/2 + |−5/2, and (e) a
squeezed state with e−r = 1/2. Note how the Wigner function for the |0 and |α
state look identical, up to a displacement. The Wigner function becomes negative
for the “nonclassical” states (c) and (d).
The Wigner function is a useful tool to reconstruct and identify bosonic quantum
states, even if they are not Gaussian. Figure 5.5c and d show the Wigner function
of a Fock state and of a Schrödinger cat (5.36). The Fock state spreads over a ring
of radius |α|2 ∼ n, while the Schrödinger cat shows the interference between to
coherent states |±α.
The Wigner function is an infinite-dimensional object that can only be approx-
imately reconstructed under realistic assumptions. The following are main recon-
struction methods:
Gaussian state verification. This method only works for Gaussian states. One mea-
sures moments of the quadratures (a † )m a n up to a certain order n + m = Nmax > 2.
These moments are used to verify that the state is Gaussian, showing that higher-
order moments n + m > 2 are related to the first and second moments by Wick’s
theorem (Menzel et al., 2012). Once the Gaussian nature is verified, the covariance
matrix and the mean values r can be derived from the moments with n + m ≤ 2.
Parity measurements. There is a formula that relates the Wigner function to the
†
parity operator = (−1)a a :
2 2
Wρ (α) =
tr(ρα ) = D(−α)† D(−α)ρ . (5.50)
π π
This formula enables the computation of the Wigner function at every point of phase
space, through a three step process: (i) prepare the state ρ, (ii) displace the bosonic
mode using some external microwave drive to construct ρα := D(α)† ρD(α), and
(iii) measure the parity over this new state tr(ρα ), for instance with the help of a
qubit (Hofheinz et al., 2009).
Truncated moments. For states where there is a maximum number of photons,
higher-order moments above Nmax can be vanishingly small. This allows us to recon-
struct the Wigner function from the smaller set of nonzero moments (Eichler et al.,
2011):
(−λ∗ )m λn 1 ∗ ∗
(â † )m â n ρ e− 2 |λ| +αλ −α λ .
2
Wρ (α = x + ip) = d2 λ 2 (5.51)
n,m C
π n! m!
quantization produces the linear, driven harmonic oscillator from (5.8), with the
√
external field (t) = h̄/ZCg V (t)/C . This external drive induces a displacement
transformation onto the field â → D(α)† âD(α) = â + α. If we switch from the
Heisenberg to the Schrödinger picture, the same operation interpreted as preparing
the resonator, from a vacuum state, into a coherent state with adjustable displace-
ment and phase |α = D(α) |0 (see Problems 5.4 and 5.5).
Let us study an oscillating microwave field (t) = 0 cos(ωd t + ϕ) that res-
onates with the circuit for a brief period of time. The solution of (5.10) predicts a
displacement of the resonator with two parts:
â(t) = e−iωt â(0) + αRWA (t) + αnon-RWA (t) . (5.52)
(a)
(b)
created by the transmission line at the interface with the LC circuit. If the coupling
is inductive, we would instead have
1 2 1
Ĥ = Ĥt-line +Q̂ + ( ˆ − MLIˆ)2, (5.57)
2C 2L
where M is the mutual inductive coupling between the LC and the line
(cf. Section 4.7.1), and Iˆ = ∂x φ̂(x0 )/ l0 is the transmission line’s current close
to the place of mutual interaction.
When we quantize either circuit, we obtain a Caldeira–Leggett model for an
oscillator {â, â † } interacting with a bath of bosonic modes, corresponding to the
microwaves in the transmission line {b̂k, b̂k† }. If the coupling is so weak that it does
not change the eigenmodes of the transmission line, we may write
H = h̄ωâ † â + h̄ â + â † gk b̂k† + gk∗ b̂k + h̄ωk b̂k† b̂k . (5.58)
k k
(3) We study the state of the radiation that leaks from the circuit into the antenna.
We derive Heisenberg equations relating the field on the line to the field in the
LC circuit, in what is known as input–output theory. This approach is used to
model the spectroscopy of a resonator in Section 5.5.3 and to develop techniques
for measuring the quantum state of the circuit based on the leaked radiation in
Section 5.5.7.
Input–Output Theory
Our goal is to analyze how the quantum fluctuations in the open transmission line
affect the resonator. We assume that the exchange of energy between both elements
happens at a slow pace, slower than the speed at which the LC circuit evolves, and
the speed at which waves propagate through the line. In this limit, the line can
control and monitor the resonator, without affecting too much its evolution. This
separation of time scales justifies a RWA effective Hamiltonian:
HRWA = h̄ωâ † â + h̄gk â b̂k† + h̄gk∗ â † b̂k + h̄ωk b̂k† b̂k . (5.60)
k k
The Heisenberg equations for this model describe a continuous exchange of pho-
tons between the oscillator and the bath (cf. Appendix B.3). Integrating out the bath,
we get a Langevin equation for the resonator:
t
d
â = −iωâ(t) − i ξ̂ (t) − K(t − τ )â(τ )dτ . (5.61)
dt t0
The second term is the input field ξ̂ (t) = k gk∗ e−iωk (t−t0 ) b̂k (t0 ) transported by the
line. The third term is the field emitted by the resonator into the line in the past â(τ ),
which may be partially reabsorbed.
The memory function K(τ ) is the Fourier transform of the spectral function:
∗ −iωk (t−τ ) 1
K(t) = gk gk e = J QO (ω̄)e−i ω̄t dω̄. (5.62)
k
2π
The spectral function of an ideal transmission line grows linearly with the frequency,
acting as an Ohmic bath:
J QO (ω̄) = 2π |gk |2 δ(ω̄ − ωk ) παω1 . (5.63)
k
Whenever the spectral function is a such broad and smooth function, the Fourier
transform K(τ ) becomes extremely concentrated around τ = 0. The Markovian
limit is a regime in which we approximate the memory function with a Dirac delta,
K(t) (iδLamb − κ2 ) × δ(t). In the Markovian limit, the bath has no memory of the
oscillator’s past, and (5.61) becomes local in time:
d κ √
â(t) = −iω − â(t) − i κ b̂in (t). (5.64)
dt 2
We have to read this equation as follows. First, the environment introduces a Lamb
shift of the oscillator’s resonance, ω = ω − δLamb . This is a slowdown of the res-
onator caused by “dragging along” the modes of the line as it evolves. However, this
slowdown is rarely discussed, because once the resonator is connected to the line
we can only measure ω , not ω.
The second term −(κ/2)â is an exponential attenuation of the resonator field. It
is caused by photons leaking into the bath at the cavity decay rate κ = J QO (ω )
παω . The ratio κ/ω is uniform across most of the spectrum: low- and high-
frequency resonators decay at the same relative speed. Since κ is the rate at which
the bath and the resonator exchange energy, consistency with the RWA in (5.60)
imposes the limitation κ/ω 1 for this whole treatment to be justified.
The last term in the equation represents the injection of photons coming from the
√
line into the resonator, −i κ b̂in . When the transmission line contains a coherent
microwave drive, such as the (t) used to control the resonator, it is customary to
perform a displacement of the input operators, separating this “classical” contribu-
tion b̂in (t) → (t) + b̂in (t), from a truly quantum noise operator b̂in (t) that accounts
for the quantum fluctuations that the bath injects into the resonator.
The Heisenberg equation for the cavity is accompanied by an input–output rela-
tion that connects the field reflected by the cavity b̂out to the input field b̂in and the
field that leaks from the resonator:
√
b̂out (t) = b̂in (t) − i κ â(t). (5.65)
Together with (5.64), this can be used to study cavity spectroscopy and quadrature
measurements, as explained in Sections 5.5.3 and 5.5.7.
As anticipated in Section 5.5.1, the losses stabilize a coherent state with a number
of photons around 4|0 |2 /κ 2 . The reemitted signal and the incident microwave
interfere, producing the reflected field
i(ω − ωd ) − κ/2
b̂out (t → ∞) = 0 e−iωd t = r(ωd )(t). (5.67)
i(ω − ωd ) + κ/2
The cavity reflects the field (t) with a phase shift r(ωd ) = e−i2 arctan(2δ/κ) . This
shift can be experimentally determined (cf. Section 5.5.7) and used to estimate the
resonance ω – not the bare frequency ω! – with very high precision.
Measuring the reflected signal from a circuit is inconvenient, because the signal
has to be separated from the original drive using circulators. A common alternative is
to work with transmitted microwaves, using the configuration in Figure 5.4a, where
we inject the signal into one side of the resonator, and measure the output signal
on the other side. We can extend the input–output theory to consider the two decay
channels of the λ/2 resonator: to the left âL and to the right âR , with the total decay
κ = κL + κR :
d √ √
â = (−iω − κ/2)â − i κL b̂Lin − i κR b̂Rin, (5.68)
dt
√
out
b̂L,R = b̂L,R
in
− i κL,R â.
A signal 0 e−iωd t coming from the right port b̂Rin results in a transmitted signal b̂Lout
and a reflected signal âRout . For instance:
√
κL κR
b̂Lout = 0 e−iωd t , (5.69)
i(ω − ωd ) + (κL + κR )/2
i(ω − ωd ) − (κR − κL )/2
b̂Rout = 0 e−iωd t . (5.70)
i(ω − ωd ) + (κL + κR )/2
The reflected signal is now zero when on resonance ωd = ω , and the transmitted
signal only contains the field emitted by the cavity:
|0 |2
n(ωd ) ∝ . (5.71)
(κ/2)2 + (ωd − ω)2
In the transmission spectroscopy setup, the LC circuit is a filter that transmits
frequencies on a narrow band around ωd = ω . The full-half-width (FHW) of this
filter – the separation between the two points at 50% maximum transmitted power –
is given by the decay rate κ and by the inverse of the quality factor Q = ω/κ.
A Q ∼ 105 means that the resonator filters out frequencies that are 0.002% outside
the central frequency ω . Such a cavity or resonator is a very good isolator in which
to embed other circuits.11 When those circuits are off-resonant with the cavity, they
will be effectively shielded from the environment. We can undo this shielding and
allow the circuit talk to the b̂k modes by shifting its resonances, bringing it closer
to the frequencies ω allowed by the resonator. These ideas will be explored in later
chapters about cavity-QED and quantum computing devices.
11 Note the superconducting qubit attached to the transmission line resonator in Figure 5.1c. In other montages,
the resonator sits in between a transmission line and the quantum register, as seen in Figure 8.2.
12 In the work by Vlastakis et al. (2013), ω ∼ 2π × 8.18 GHz, κ ∼ 2π × 7.2 kHz and Q ∼ 106 . This supports
extremely long-lived Schrödinger cat states and error correction!
Figure 5.7 (a) Microwave beam-splitter with 90◦ hybrid design. (b) Scheme of a
three-port circulator.
port into two different ports. In visible optics, beam splitters are partially reflective
mirrors, but in microwaves we have to use sophisticated designs of waveguides
with different lengths and impedances.
Figure 5.7a shows a 90◦ hybrid, a microwave beam splitter with resonance fre-
quency λ/2. The device is a square with four segments of coplanar waveguides √with
the same length λ/2, and different impedances – Z0 on the vertical arms, Z0 / 2 on
the horizontal segments. The hybrid connects to four waveguides at its four corners,
which we can use to inject and extract microwaves. Right at the frequency resonant
with the hybrid’s wavelength λ, the input and output modes are related by the simple
unitary transformation of a 50–50 beam splitter:
⎛ ⎞ ⎛ ⎞
b̂1out ⎛ ⎞ in
⎜ out ⎟ 0 0 i 1 ⎜b̂1 ⎟
⎜b̂2 ⎟ −1 ⎜0 i⎟ ⎜ in ⎟
⎜ ⎟= √ ⎜ 0 1 ⎟ ⎜b̂2 ⎟ . (5.75)
⎜b̂out ⎟ 2 ⎝i 1 0 0⎠ ⎜ in ⎟
⎝ 3 ⎠ ⎝b̂3 ⎠
b̂4out 1 i 0 0 b̂in
4
If we inject a signal through port 1 or through port 2, it splits into an equal super-
position of ports 3 and 4. For instance, S † b̂1in † S = − √i 2 b̂3out † + − √12 b̂4out † . The scat-
tering matrix of the hybrid depends on the frequency of the incoming microwave.
Away from resonance, not only is the splitter unbalanced, but some the energy
is spread over all four ports – including some undesired back-reflection. How-
ever, these effects are negligible within a broad band of 10% the central resonance
(Schneider, 2014).
Beam splitters, in this or other designs, are used to divide and to merge signals.
For instance, in Section 5.5.7 we divide a signal into two different ports b̂3,4 and
process those ports separately. This allows us to measure two different quadratures
(b̂1† + b̂1 ) and i(b̂1† − b̂1 ) of the original input field. Beam splitters can also be used
combine light from two different bosonic modes and create entangled microwaves
(Menzel et al., 2012).
The circulator maps an input signal from one port straight into the following port,
with a circular order denoted in circuits by an arrow (see Figure 5.7b). A circulator
breaks time reversal symmetry: If we reflect back a signal that exits from port 2, it
does not return to the input port 1, but exits from port 3. To break this symmetry,
circulators are built with permanent magnets, as bulky devices that cannot be placed
on chip – although there is ongoing research in implementing such circulators using
2D electron gases or periodically driven quantum circuits.
Circulators are used to isolate systems or separate signals. In the first type of
applications, we connect our circuit to port 1, ground port 3, and connect a measure-
ment device on port 2. This way our circuit is not affected by the noise generated by
the measurement device – which sinks into port 2. In the second type of application,
the setup is identical but we connect a microwave source to port 3. This way, the
measurement device only collects the light reflected by the circuit at port 1 and is
not affected by the microwave source.13
5.5.6 Amplification
The Gaussian set of operations includes the amplification of one or more quadra-
tures to make them more easily measured. Amplification appears naturally when
look at the single-mode (5.38) and two-mode (5.41) squeezing transformations in
terms of the gain factor G = cosh(r)2 ≥ 1:
√ √
â(r) = Gâ(0) − G − 1ei2θ â † (0), (5.77)
√ √
b̂1 (r) = Gb̂1 (0) − G − 1e2iθ b̂2† (0). (5.78)
The first equation is an example of a degenerate, phase-sensitive linear amplifier.
It is degenerate because it operates using a single mode â and just one frequency
of photons. It is phase sensitive because amplification only takes place along one
direction in phase space, p̂φ (r) = √i 2 (â † eiθ − âe−iθ ), which is the one we should
use in the measurements. This type of process is also called noiseless amplification
because it preserves the relative strength of fluctuations p̂/p̂, without any increase
in noise due to the amplification.
13 See, for instance, the work by Hoi et al. (2011, 2012), where a superconducting qubit operates on microwave
light, sorting out photons and creating non-Gaussian states.
The last term is the number of photons added by the amplification process. A quan-
tum limited amplifier is one which adds the minimum amount of photons allowed
by quantum mechanics n̄amp = b̂2† b̂2 = 0.
We will assume that this oscillator is connected to a waveguide that brings the
signal and collects the amplified output. Using the input–output theory for a single
lead produces two linear equations:
d √
â = (−iωamp − κ/2)â − igei(t+θ) â † − i κ b̂in, (5.81)
dt
√
b̂out = b̂in − i κ â. (5.82)
We can solve these equations and establish a linear transformation from input b̂in
to output b̂out modes in frequency space. Following Roy and Devoret (2016), we
√
express â = (b̂in − b̂out )/ i κ and construct an equation that couples annihilation
and creation operators at different frequencies:
K †2 2
H0 = h̄ωa â † â − â â . (5.88)
2
The Kerr nonlinearity K can be traced back to the Josephson junction potential,
cos(φ̂/ϕ0 ) = 1 − 12 (φ̂/ϕ0 )2 + 14 (φ̂/ϕ0 )4 + · · · . The input–output theory for the JPA
with a signal b̂Sin and a pump b̂pump
in
is
d κ1 + κ2 √ √ in
â = −iωa − â + iK â † â 2 − i κ1 b̂Sin − i κ1 b̂pump . (5.89)
dt 2
This model is equivalent to the linear parametric amplifier in (5.81), with the pump
2ωp = , phase θ = 2ξ , intensity-dependent resonator frequency ωamp = ωa −
2Kα02 , and intensity-dependent coupling g = Kα0 . Using these identities, we can
find the regimes of nondegenerate parametric amplification ωp ωa , where the
JPA acts as a phase-insensitive amplifier, as well as the phase-sensitive degenerate
amplifier regime, ωp ωS .
JPAs are fantastic amplifiers that power many of today’s experiments. Unfor-
tunately, the stronger the amplification, the narrower the bandwidth over which
the JPA operates. Typical designs of JPA amplify signals of gigahertz with large
gains GD,I ∼ 22 dB, over a relatively narrow frequency band of around megahertz
(Castellanos-Beltran et al., 2009).
There is an alternative to the JPA that supports comparable gains over a broader
range of frequencies. These devices are known as traveling wave parametric
amplifiers (TWPAs). TWPAs were introduced by Cullen (1960), but they have
been adapted to work with a long, open string of chained Josephson junctions or
SQUIDs. The chain provides a 1D model with a Kerr nonlinearity that is capable of
amplifying the waves that propagate through it, with comparable gains of ∼ 20 dB
over a bandwidth of 3 GHz (Macklin et al., 2015). A single amplifier can work
over multiple control signals for a superconducting circuit, and can be used to
amplify the output from many different measurement devices, simultaneously, and
without interference between signals. This is ideal for large and complex quantum
computing setups, where tens and probably hundreds of qubits must be manipulated.
Unfortunately, the theory and details of TWPAs fall outside the scope of this book,
but you are encouraged to dig into the literature to learn about this incredibly useful
technology.
with a prefactor Ḡ that includes all gains and losses – also the attenuation at the
cavity port κ – and an effective field ĥ summing all classical and quantum noise. This
includes noise from the amplifiers, beam splitters, detectors, and cables themselves.
The added noise is quantified in the number of photons it adds prior to the gain
itself (5.79). This noise can be as low as 1/2 in ideal quantum-limited amplifiers, or
even lower in noiseless phase-sensitive amplifiers. However, when further connec-
tions and mixers are accounted for,14 we find more common values of 1−10 pho-
tons in Josephson-junction-based amplifiers, and 30−50 in cyrogenic high-electron
mobility transistor (HEMT) amplifiers.
Power Measurements
We can feed the signal b̂amp (t) into a device that measures the integrated or average
power. A broadband power meter uses diodes to rectify the microwave current,
creating a dc signal that is calibrated to determine the average, peak, or integrated
power in the signal. In practice, we need to very accurately calibrate the average
power Pnoise of the background noise – i.e., noise introduced by amplification stages,
input lines, etc. – to separate the power of the actual signal Psignal = P −Pnoise , from
the measured power P .
Since the dynamical range of a detector can span multiple orders of magnitude,
the signal is measured using a logarithmic scale. The dBm unit is a log-10 scale
referenced to a standard power of 1 milliwatt. A power P in watts is converted to
this scale as PdBm = 10 log10 (P /10−3 ). A very good broadband power detector
operating in the gigahertz range can have a minimum sensitivity of −70 dBm. This
sensitivity can be improved using spectrometers that work on narrow bands around
the desired frequency. A good device can have a detection threshold of around
−140 dBm/Hz, relative to the bandwidth. If we wish to detect photons generated
by a cavity with a bandwidth κ from 1 kHz to 1 MHz, this means the threshold for
detection will lay around −110 and −80 dBm, respectively.
An important problem is that even in this scale, quantum microwave signals
can take ridiculously small values. When a cavity releases photons of frequency
ω ∼ 2π × 10 GHz at a rate κ = 1 MHz, the output power will be Pphoton = h̄ωκ ∼
6.626 × 10−18 W, or PdBm ∼ −171 dBm. This is ∼ 10−9 or ∼ −90 dB to below
the threshold of commercial power detectors! Hence the need for various stages of
amplification that bridge this gap.
Another problem is obtaining a good signal-to-noise ratio for Psignal , estimated as
the difference of two large quantities, the measurement with signal P , and a calibra-
tion of the noise in the empty line Pnoise . Both quantities have to be measured with
high precision, over long integration times, and fluctuating experimental conditions.
This challenge in stability and calibration makes this measurement technique less
desirable than quadrature measurement techniques.
Quadrature Measurements
The use of power meters has been superseded by the use of digital techniques.
An analog-to-digital converter (ADC) can periodically sample the voltage of an
incoming quantum signal, producing a high-resolution stream of data from which a
computer – or a field programmable gate array (FPGA) card (Eichler et al., 2012) –
estimates moments â , â † â, etc. The benefit of this technique is that it provides
more information than just a power measurement, but actual implementations face
two obstacles. First, the ADC will only sample one quadrature, the voltage â(t) +
â † (t). Second, we must slow down the signal, because the ADC cannot sample fields
b̂amp (t) ∼ â(t) that oscillate at frequencies of gigahertz.
Theoretically, both problems have easy solutions. The first obstacle is avoided
by dividing the signal b̂amp into two copies – one in phase, and another one with a
90◦ rotation eiπ /2 b̂amp – and feeding each copy to a different ADC. This provides
us with two streams of values, associated to the two quadratures √12 (b̂amp + b̂amp† )
and √i 2 (b̂amp − b̂amp† ).
The second problem is solved by looking at how the cavity operator evolves.
From input–output theory (cf. Section B.3.2), we know that most of the oscillation
in the field is due to “trivial” phases, â(t) = âslow (t)e−iωt . The actual information
is encoded an operator that evolves at a much slower pace âslow (t) = eiωt â(t). We
just need to fabricate this operator by mixing the original signal with the opposite
field eiωt .
Both solutions are jointly solved by a type of microwave circuit called mixers.
A simple mixer is a nonlinear circuit that takes two signals f (t) and i(t), and returns
their product f (t)i(t), as sketched in Figure 5.8a. If we feed our signal â(t) and a
classical oscillating field i(t) = cos(ωm t +θ) into a mixer, it will produce two copies
of the signal at two sidebands. The first sideband oscillates with a lower frequency
and is close to our slow modulation ei(ωm t) â(t) ei(ωm −ω)tx âslow (t). The second side-
band oscillates at higher frequency ω + ωm and can be filtered out. In addition to
this, note how the sidebands will be phase-shifted an angle θ determined by the
input field. We can use this angle to select which quadrature we wish to explore.
The complete setup is shown in Figure 5.8c. We feed the signal to an amplifier
and chain the output into an IQ mixer. The IQ mixer uses two mixers to combine
the signal with two reference microwaves, cos(ωm t) and sin(ωm t), derived from a
stable local oscillator. The two outputs of the IQ mixer are two quadratures, X̂ and
P̂ , that are sampled by two ADCs. This produces two classical traces X(tn ) and
Figure 5.8 (a) A mixer is a nonlinear device that multiplies a signal f (t) with a
modulation i(t). (b) An IQ mixer divides the input signal, combining it with two
signals in 90◦ opposition cos(ωt) and sin(ωt). The output of the mixer contains
the two orthogonal quadratures of the signal. (c) The combination of a mixer
with a local oscillator, filters, and amplifiers allows us to measure any quadrature
Q̂ω,θ = √1 (aω e−iθ + aω† e−iθ ) in a quantum field φ(t).ˆ The classical signal q(t)
2
is a random sequence of voltages associated to one realization. Averages of many
realizations of q(t) estimate the expectation value Q̂ω,θ (t).
P (tn ) that combine into a complex number Z(tn ) = X(tn ) + iP (tn ). In practical
applications, ωm − ω is never zero, but a finite value around MHz. This is slow
enough that the ADC can reconstruct the signal, but large enough that the signal
is not masked by slow 1/f noise in the electronics. If we take into account this
residual oscillation and calibrate
√ the gains of our circuit, we can build a sequence
S(tn ) = Z(tn )e−i(ωm −ω)t / Ḡ, that is a sampled measurement of our slow field with
some noise ŝ(t) = âslow (t) + ĥ† (t).
The classical values S(tn ) can be used to reconstruct a measurement of resonator
quadratures â. Let us assume that we prepare the LC circuit and stop all external
controls. As soon as we do, the cavity field begins to decay exponentially âslow (t) =
â(0)e−κt/2 . Using the weight function w(t) = κe−κt/2 , we recover the cavity oper-
ator from traces of our ADC chain:
w(t)ŝ(t)dt = â(0) + ĥ†w ∼ Sw := w(tn )S(tn )t. (5.92)
n
ĥ†w = 0, (ii) the statistics of the noise are independent of the signal ĥ†w â
ĥw â 0, and (iii) noise moments – ĥw ĥ†w or higher – can be calibrated from
experiments with an empty cavity. This allows us approximate arbitrary moments
of the cavity field as â(0) Sw , â † (0)â(0) Sw∗ Sw − hw ĥ†w , etc. First,
second, third, and higher moments are used to test whether a state is Gaussian, and
to do Wigner function tomography as explained in Section 5.4.6.
This analysis of bosonic quadratures is a powerful method that has been used to
reconstruct the wavefunctions of squeezed states (Mallet et al., 2011) and of single
photons (Eichler et al., 2011). The quadratures themselves allow the reconstruction
of other properties, such as the energy of a wavepacket with less than one pho-
ton – i.e., values around 10−26 Joules! – or the spectral distribution of a photon
wavepacket (Menzel et al., 2010).
5.6 Conclusion
We close here a long chapter on microwave photons. The chapter began discussing
how linear circuits – and nonlinear circuits operating in a linear regime – store
energy in quantized units, which we called photons. These bosonic excitations
behave as optical photons in all respects, from the linear spectrum of equispaced
eigenenergies, h̄ω,2h̄ω,3h̄ω . . ., to the quantum description of modes and wave-
functions.
The noninteracting nature of photonic excitations poses both advantadges as
well as problems. Among the advantages, we find a relatively simple mathematical
description, access to amplification, and real-time tomography of wavefunctions.
Photons can also be confined in high-quality environments, for long times of
hundreds of microseconds and support long-lived entangled states and quantum
superpositions.
The problems begin when we want to create those interesting states or think about
developing a quantum computer. We cannot use linear circuits to generate or detect
individual photons, and we are constrained to a family of states (Gaussian states)
that can be simulated classically and pose no computational interest. To escape these
constraints, we must introduce nonlinearities. This is the focus of the next chapter,
where we study the design and operation of superconducting qubits.
Exercises
5.1 Compute the expected values and uncertainties of q̂, φ̂, V̂ , and Iˆ for the vac-
uum state |0 of an LC resonator.
5.2 Plot the average occupation of a microwave cavity as a function of the cavity
resonance ω and the temperature of the cryostat, for temperatures ranging from
5–100 mK. What is the minimum frequency at which the occupation number
lays below 10%? And 1%?
5.3 Show that the number of photons in a coherent state |α is n̂ = |α|2 . Compute
the uncertainty of all quadratures, {φ̂, q̂}, on such a state.
5.4 Show that the ground state of Hamiltonian (5.8), for constant drive (t) = 0 ,
is a coherent state with α ∝ . Hint: Use the definition of the coherent state
in terms of displacement operators, and the commutation relations between a,
a † , and D(α), such as âD(α) = D(α)(â + α).
5.5 If we start with the vacuum as initial state |(0) = |0, the coherent drive (t)
in (5.8) will produce a coherent state |(t) ∝ |α(t) with the displacement
from (5.10). Relate the Schrödinger and Heisenberg pictures, showing that
U (t)† â |(t) = â(t) |(0). Use the explicit formula for â(t) to demonstrate
that |(t) is a coherent state. Bonus points if you get the phase of the coherent
state right.
5.6 Compute the Wigner function for a single photonic mode in (i) a coherent
state, (ii) a Fock state with one photon, and (iii) a Schrödinger cat with generic
displacement α. Which of these states is Gaussian? Which of them has a
negative Wigner function? Which of them is broader in phase space and which
of them saturates the Heisenberg uncertainty relation?
5.7 We want to place a magnetic molecule inside a λ/2 and a λ/4 transmission
line resonator. The molecule couples to the magnetic field generated by the
nearby currents in the line.
(1) Which position optimizes the coupling between the molecule and the fun-
damental mode, ω0 ?
(2) Which position optimizes the coupling to the first harmonic of the cavity,
ω1 ?
(3) Where should we place the molecule so that it couples equally well to both
modes, ω0 and ω1 ?
5.8 Cavity spectrum. In Section 5.5.3, we studied how the driven-dissipative res-
onator stabilizes to an average occupation number. We want to solve the same
problem but using master equations:
(a) Starting from the driven model H = a † a + (∗0 e−iωd t a + 0 eiωd t a † ), and
following Section B.2, show that the master equation in the interaction
picture becomes a function of the detuning, δ = ωd − ω:
κ
∂t ρ = −i δ â † â + 0 a + H.c.,ρ + 2âρ â † − â † âρ − ρ â † â . (5.93)
2
(b) Write and solve the evolution equations for â = tr(âρ) and for the
number of photons â † â. Hint: Use the cyclic property of the trace to
convert terms of the form tr(AρB) = tr(BAρ) = BA.
(c) Show that the FHW linewidth is δ = κ.
5.9 Assume a cavity without driving, coupled to an environment at zero temper-
ature. Use the master equation (5.73) to solve analytically the density matrix
of the cavity, using as initial conditions the pure states ρ(0) = |ψ ψ| in a
superposition of one photon and the vacuum, |ψ = cos(θ ) |0 + sin(θ ) |1.
Show that the average number of photons â † â = tr(ρ(t)â † â) in this state
follows (5.74). Study the decay of the coherence 0|ρ(t)|1. How fast is it,
and how does it relate to κ?
5.10 A microwave signal passes sequentially through two amplifiers with gains
G1,2 and added number of photons n̄1,2 . Each stage is described by an equation
of the form (5.91), with noise operators ĥ†1,2 that are independent from each
other – e.g. ĥn† m† r† k n† m† r† k
1 ĥ1 ĥ2 ĥ2 = ĥ1 ĥ1 ĥ2 ĥ2 – and from the signal. Use this
property to estimate the total gain and total number of added photons of the
device.
We have already introduced the qubit as the smallest quantum mechanical system
that we can study. In this chapter, we further explore qubits, examining how they
behave, both formally – through Schrödinger and master equations – and through
their implementation in various superconducting circuits. Each design has unique
characteristics, but we can develop a common framework for describing the qubit’s
evolution, its coupling to the environment and to external controls, the qubit’s qual-
ity and coherence properties, etc. We also discuss in detail the most used qubits
in present and past state-of-the-art experiments – charge qubits, transmons, and
flux qubits. This chapter intentionally drops some interesting circuits – fluxoniums,
g-mons, and other curious animals out there – but it should be enough to approach
any qubit architecture critically, deducing their Hamiltonians and coherence prop-
erties, and understanding the pros and cons of each those designs critically.
106
The word bit applies to both the unit of information and to the physical system
that encodes it. Similarly, a qubit would be the smallest unit of quantum information,
as well as any physical system that embodies those two-dimensional superpositions.
Qubits as such do not exist in Nature, because all known physical systems require
more than just two quantum states to be described. A photon has a polarization
degree of freedom, but also frequency and momentum; an electron has a spin,
but also a spatial wavefunction. Creating a physical qubit requires controlling the
dynamics of a quantum system, constraining it to a subset of all physically available
quantum states.
In addition to this simplified dynamics, there other practical requirements, includ-
ing preparation, read-out, and measurement. We can group them in a “shopping
list,” a variation of DiVincenzo’s famous requirements for a quantum computer
(DiVincenzo, 1995):
(1) The physical qubit has two accessible orthogonal eigenstates |0 and |1.
(2) It supports arbitrary, long-lived quantum superpositions α |0 + β |1.
(3) We can reset the qubit to some state, typically |0.
(4) We can do a projective measurent on the computational basis, typically through
the observable σ z = |11| − |00|.
(5) We can perform (or approximate) arbitrary unitary rotations in the qubit’s
Hilbert space.
(6) We can implement at least one type of universal two-qubit operation among
pairs of qubits.
(7) We can engineer devices with medium to very large numbers of qubits.
(8) Tolerances in all operations – unitary gates, measurements, qubit reset – are
tight enough to implement fault-tolerant quantum computation.
Out of this list, items (1)–(5) suffice for many applications in quantum communi-
cation and quantum optics. Combined with some type of qubit–qubit interaction –
which unlike in term (6) does not need to be very precise (1) this even allows us
to build some types of quantum simulators. Conditions (6) and (7) were recently
achieved together in setups with more than 50 qubits (Arute et al., 2019; Wu et al.,
2021). This enables near-term intermediate scale quantum computers – also known
as noisy intermediate-scale quantum (NISQ) computers thanks to Preskill (2018) –
to perform tasks, such as the simulation of random quantum circuits, which are
arguably difficult in the classical computing world. It also opens the door to other
useful applications of imperfect near-term quantum computers, especially in the
study of quantum physical systems.
However, the ultimate goal in the field is to achieve fault-tolerant quantum com-
putation as a means to build quantum computers that run arbitrarily long quantum
algorithms. Fault-tolerant devices need extremely large numbers of qubits to encode
information redundantly. These qubits are operated with arbitrary numbers of mea-
surements and quantum gates to implement quantum algorithms that are resilient
to environmental and operational errors. We will discuss this requirement further
in Section 8.5. However, as of the writing of this book this is a regime that is far
from being achieved, even if we have promising results and proofs of principle
demonstrators Chen et al. (2021).
The quest for physical qubits has ran in parallel to – and often been the motivation
for – the search for controllable quantum systems in the lab. Table 6.1 enumer-
ates the most successful qubit systems, together with the degrees of freedom that
encode the information and the energy scales involved. Indeed, the energy scales
of the qubit degrees of freedom are extremely important, because they underlay the
general conditions for doing quantum experiments regarding temperature, isolation
requirements, cooling, and preparation times (see Section 4.1.1).
However, our shopping list introduces a new and extremely important considera-
tion: the need of single out two states from an experimental device to encode a qubit.
This is not as easy as it sounds. Take, for instance, the photonic superconducting
circuits from Section 5.4. We could identify the qubit with the states of zero and
one photon, |0 and |1. However, as discussed in Problem 6.2, we cannot rely on
microwaves to implement transitions between |0 and |1 without involving other
photon number states. Linear LC-type circuits therefore do not satisfy the require-
ments for qubit preparation, arbitrary rotations in the qubit space, or even projective
measurements.
Figure 6.1 (a) A general qubit uses two metastable states |0 and |1, out of
a spectrum that is anharmonic – that is, the energy spacing is not uniform. In
particular, the qubit ω01 is not an integer multiple or fraction of the energy
difference to other neighboring states of the system, ω0n or ω1n . (b) Energy levels
of the Ca+ ion, including the two qubit states |0 = |S1/2 and |1 = |D5/2 .
Note that in addition to the quantum information bit and the physical qubit, there
exists the idea of logical qubit, analyzed later in this book. In this case, the quantum
bit is stored formally in the Hilbert space of a complex quantum system. The qubit is
no longer identified with the eigenstates of that model, but is prepared, operated, and
measured in a sophisticated, error-correcting, or error-suppressing way. Typically,
logical qubits can be synthesized out of systems from many imperfect physical
qubits. However, as demonstrated by Ofek et al. (2016), one can also develop logical
qubits using complex quantum systems, such as Schödinger cats created by the inter-
action between superconducting qubit circuits and superconducting cavities. Here
the superconducting qubit is just a nonlinear circuit that enables the preparation,
operation, measurement, and error correction of the logical qubit, which is no longer
an associated to an eigenstate of the cavity but to a very long-lived quantum state.
The levels |n are the eigenstates of the circuit that makes our atom. The lowest
energy states |0 and |1 are identified with our qubit. As sketched in Figure 6.1a,
the spectrum is anharmonic and the gap h̄ω01 = E1 − E0 is not an integer fraction
of the spacing to other levels, ω02,ω12 . . .
There will be typically two direct controls on any qubit. Tuning an external
parameter such as a magnetic flux , we can change the energy levels En ()
preserving their population. We can also “illuminate” the qubit with microwaves
(t) that couple to the off-diagonal dipolar moment operator n|d̂|n = 0 and
induce transitions between eigenstates.
By using weak or slow controls, and engineering the spacing between levels,
we can constrain the dynamics of the artificial atom to just two lowest energy
states. This effective two-level system lives in a reduced Hilbert space H2 created
by arbitrary linear superpositions of |0 and |1. Following Section 2.2, we model
the dynamics in this space using the most general real-valued1 qubit Hamiltonian:
h̄(t) z h̄ε(t) x
H = σ + σ + E0 1 = E0 + B(t) · σ . (6.2)
2 2
1 It will become evident, when we introduce actual qubit designs, that we can write all models using two
observables, one diagonal – the qubit energies – and one off-diagonal – the dipole moment operator.
A consistent selection of phases in the definition of states |n allows us to represent those observables as σ x
and σ z (see Problem 6.1).
We can get rid of this “free evolution” and eliminate the dynamical phases by work-
ing in a new rotating frame, called the interaction picture. This involves redefining
the wavefunction and the observables:
so that expectation values and predictions remain unaltered ψI |ÔI |ψI = ψ|Ô|ψ.
In this frame, diagonal observables such as σ z remain the same and the evolution
of the new wavefunction ψI is dictated by an interaction picture Hamiltonian:
d
i h̄∂t |ψI = HI (t) |ψI = U0† (t)H (t)U0 (t) − i h̄U0† (t) U0 (t) |ψI (t) . (6.5)
dt
The diagonal part is proportional to the detuning between the qubit frequency and
the drive δ = − ω0 . Out of the off-diagonal terms, we can single out the static one
e−iϕ σ + , using a rotating wave approximation to neglect the interaction that rotates
with twice the frequency ei2ω0 t+iϕ σ + .
2 In recent years, a resource theory has been put forward to quantify coherence (Baumgratz et al., 2014) as a
function of the off-diagonal elements in the density matrix. From this theory, it follows that decoherence is the
destruction of such a resource. We are not so rigorous in our description.
The map predicts an exponential decay of the coherences – the off-diagonal ele-
ments of the density matrix – with a dephasing rate γφ = σ 2 /2, and a dephasing
time T2 ∼ 1/γφ .
The dephasing channel (6.14) is reproduced by the dephasing master equation:
d i γφ
ρ = − [H,ρ] + (σ z ρσ z − ρ). (6.15)
dt h̄ 2
The Hamiltonian H = 12 σ z accounts for the noise-free evolution, while the Lind-
blad operator L[ρ] = σ z ρσ z − ρ models the destruction of coherences.
The dephasing master equation can be derived – as we did for the cavity environ-
ment coupling – from a microscopic model where the energy perturbations δEn (t)
arise from a diagonal coupling with a bath, provided the spectrum of fluctuations
in the bath’s Markovian limit reproduces the desired white noise model over all fre-
quencies. However, the white noise model of dephasing is a bit simplistic. Electrical
circuits – and mesoscopic systems in general – are affected by what is known as
1/f -noise or slow noise. When we study the power spectrum of these fluctuations,
the contribution of noise diverges at low frequencies. This implies a slow decay of
noise correlations – e.g., δE(t )δE(t) ∼ exp(−(t − t)/Tc ) with a large Tc that
prevents self-averaging. The limit in which the correlation time diverges Tc → ∞
describes a quasistatic noise, which remains more or less constant throughout each
experimental realization, but fluctuates between runs. Such models are described
with variants of (6.13), with other decay forms that depend on the spectral properties
of the noise (Ramos and García-Ripoll, 2018), and do not have a Markovian master
equation (see Exercise 6.3).
However, even if slow fluctuations are theoretically inconvenient, they may be
beneficial, because we can design controls to suppress them. Assume for instance,
that the noise δEn has long time correlations. There is a technique, called spin-echo,
that exactly cancels this fluctuation, and which consists in introducing a spin flip
halfway through the experiment:
If the noise is not quasistatic, but has a finite-time correlation T , we can still apply
spin-echo at regular intervals with a spacing shorter than the fluctuations, time scale
δt T . This leads to a total or partial cancelation of noise, extending the lifetime
The Coulomb blockade is an energy scale that depends on properties of the con-
ducting material, such as the density of charges and the geometry of the piece.
It predates the notion of qubits and the study of quantum information: Coulomb
blockade is observed in quantum dots, in transport experiments with molecules,
and, of course, also in the superconducting world.
Fortunately, we do not need to worry about microscopic models to describe a
charge qubit. Since we work with macroscopic devices, with a humongous number
of particles, we can treat the superconducting island as a lumped-element circuit,
describing the total energy as a quadratic function of the total charge, without caring
for how this charge distributes across the island. The work or energy required
to charge the island is given by the self-capacitance of this metallic object, as
expected:
1
E= Q2 . (6.21)
2Cself
If our excitations are electrons with undivisible charge e, we can define a constant
called the charging energy EC = e2 /2Cself and express the total cost in term of the
number of carriers n = Q/e, as in E = EC n2 .
Note the anharmonicity in this expression: The cost of adding one particle is
E1 − E0 = EC , but the second charge involves a larger penalty E2 − E1 = 3EC .
We could define our qubit using the Hilbert space of the neutral state and the state
with a single Cooper pair, |0 and |1. The total Hamiltonian then would read
EC z
H = EC |11| =σ + constants. (6.22)
2
This situation is inconvenient for several reasons: The splitting EC is too large for
a small island, we have no means to modify this splitting, and we also do not yet
know how to add other control terms to the Hamiltonian, such as a dipolar coupling
with σ x = |10| + |01| (see (6.2)). All this requires a slightly more sophisticated
circuit model.
E
C
C V
10 μm 500 nm
100 μm
Figure 6.2 (a) Superconducting island coupled to the ground plane through two
Josephson junctions. (b) Zoomed-in picture of one junction. Pictures courtesy of
Andreas Wallraff, ETH, Zurich. (c) Reminder of the equivalent circuit discussed
in Section 4.6, with the superconducting island marked in gray. The two junctions
add up to a single effective junction with nonlinear inductance that depends on the
flux enclosed by the loop. (d) Image of a transmon qubit, a charge qubit where
the capacitance has been dramatically enlarged.
Figure 6.3 Energy levels of the charge qubit without the Josephson junction
(dashed) and with the tunneling amplitude EJ 0.1EC (solid). At zero voltage,
the lowest eigenstates are |0 and |1. At ng = 1/2, both states become almost
degenerate but the tunneling creates new eigenstates that are superpositions of the
former |± = √1 (|0 ± |1).
2
Figure 6.3 shows the energy levels En of the charge states in dashed lines. Around
the point ng = 12 , states |0 and |1 become degenerate3 and well separated from all
other charge states, n = 2,3, . . . . This makes them good candidates for qubit states,
provided we find ways to reproduce the qubit Hamiltonian (6.2) using the external
controls and V .
The first control is given by the external potential. If we perturb the value of the
offset charge slightly away from ng = 12 , we obtain an energy splitting between
the 0 and 1 charge states that grows as h̄ε = 8EC (ng − 12 ). The second control is
due to the nonlinear inductance of the SQUID. This effective Josephson junction
enables the tunneling of charges in and out of the island (4.63), cos(φ̂/ϕ0 ) =
1
2 n (|n + 1 n| + |n n + 1|). Assuming that the capacitive interaction is the
dominant energy scale EC > EJ , and that we are close to degeneracy ng 1/2,
the Josephson energy may be truncated to the 0 and 1 charge states, as −EJ ()σ x
with the tunneling operator σ x = |10| + |01|. Altogether, we obtain the desired
qubit Hamiltonian in a 2 × 2 matrix form:
(
h̄ε z h̄ x h̄ε = 8EC ng − 12 ,
H = σ − σ , with (6.25)
2 2 h̄ = EJ ().
3 There are other points n = 3 , 5 , . . . , where different pairs of states become degenerate. However, the
g 2 2
physics of those points is similar to that which we discuss later and presents no formal advantage.
The energy level structure of the qubit is usually referred to as the qubit hyper-
bola, because E0 and E1 satisfy the equation of a hyperbola with asymptotes at
E0,1 ± 12 |ε| when the applied potential ε is very large.
The qubit hyperbola is symmetric with respect to the symmetry point ε = 0,
where the separation between energies is minimal E ≥ . In absence of Joseph-
son energy, = 0 and the two charge states |0 and |1 would have the same
energy (cf. Figure 6.3 and dashed line at ng = 1/2). However, the introduction
of quantum tunneling has activated quantum fluctuations of the charge and it has
broken the degeneracy, introducing a minimal separation . This separation causes
the new eigenstates of the problem to be the symmetric and antisymmetric charge
superpositions:
1 1
|± = √ |0 ± √ |1 . (6.28)
2 2
The symmetry point ε = 0 is experimentally relevant, as the configuration in
which the qubit experiences the least dephasing. Indeed, if the qubit is surrounded
by stray electric fields that introduce random contributions δV , these contributions
cancel to first order ∂E/∂V = 0 at ε = 0. Thus, the energy levels fluctuate the
least around this point, and the lifetime of our quantum states is maximized. The
notion of a symmetry point for safe operation is a very general concept that appears
also in the flux qubit. In Section 6.3 we will push this idea further, ensuring that the
energy levels become insensitive to perturbations in the electrostatic field, which
gives us the transmon qubit design.
Finally, a remark on the notational discrepancy between our charge qubit Hamil-
tonian (6.25) and the general qubit model (6.2). This difference does not affect the
physics, as both models are related by a unitary transformation that swaps the σ x
and σ z operators (see also Problem 6.1). This transformation is a relabeling with
new qubit states:
|0̃ = |− h̄ z h̄ε x
⇒ H = σ̃ + σ̃ . (6.29)
|1̃ = |+ 2 2
The diagonal Pauli matrix is now σ̃ z = |1̃1̃| − |0̃1̃|. Some works favor this nota-
tion, because only the states |ĩ – not the |i – can be experimentally detected,
especially when working with qubits inside cavities (cf. Section 7.4.2).
people’s opinion regarding the future of the field, suggesting that it was possible
to engineer complex quantum states and quantum registers (Makhlin et al., 2001).
Later experiments with similar schemes demonstrated two-qubit Hamiltonians and
qubit–qubit interactions (Pashkin et al., 2003; Yamamoto et al., 2003), a crucial
ingredient for scalable computations.
However, Nakamura’s charge qubit implementation was inadequate for scalable
computations. The superconducting islands decohered very quickly T1 ∼ 2 ns,
because of the surrounding antennas and probes that controlled and measured the
qubit. To compensate for this, the qubits were engineered with huge energy gaps,
EJ 84 μeV ∼ 20 GHz, and operated very rapidly, and even then only allowed
for a few oscillations before losing all quantum properties.
In 2004, the Yale group had the idea of fabricating a charge qubit inside
a microwave resonator. As we have seen in Chapter 5, microwave cavitives
only absorb excitations at discrete frequencies, ωn = n × ω0 , filtering out other
electromagnetic fields. By placing the qubit inside the cavity and ensuring that
the qubit’s energy gap E was very different from the allowed photons, they
could effectively isolate and protect the qubit from most of the electromagnetic
fluctuations. The idea of Purcell filtering the environment was put forward in the
theory work by Blais et al. (2004) and immediately demonstrated by Wallraff et al.
(2004) in an experiment that showed times of 200 ns (see also Schuster et al.,
2005). That was orders of magnitude better than the bare charge qubit demonstrated
only a few years before.
The conditions of relative isolation provided by the cavity, together with oper-
ating the qubit close to the symmetry point, made it possible that the coherence of
the superconducting qubit was no longer limited by relaxation but by environment-
induced dephasing. In Yale’s charge qubits, dephasing rates of γφ ∼ 2π × 750 kHz
or T2 ∼ 1μs, allowed testing many quantum optics ideas, as well as some primitive
quantum information protocols. These were wonderful beginnings with fast-paced
developments in the study of the Rabi model, strong coupling, Wigner function
tomography, and large entangled states. However, developing scalable quantum
computers required improving the quality of the qubits even further. The avenue
that has proved most fruitful was to further suppress dephasing, reducing the
curvature of the qubit hyperbola even further. This idea led to the transmon
qubit in 2007, a significant redesign of the charge qubit’s parameter, which
we now discuss.
5 Shunting means providing an alternative (parallel) path for the charges to flow. The transmon results
from capacitively shunting a Josephson junction, with one or two capacitors that are connected in
parallel to it.
6 A maximal certainty of the quasimomentum n corresponds to the maximum uncertainty of the position ϕ.
Figure 6.4 (a) Rescaled energy levels of a transmon qubit as a function of the
offset charge ng , for ratios EJ /EC = 0.02 (dashed), 2 (dot-dashed), and 16
(solid). The values plotted lack an irrelevant constant and are scaled to the qubit
gap with no offset charge εn = (En (ng ) − E0 (0))/h̄ω10 . (b) For a transmon with
EJ /EC = 16, we plot the energy levels relative to the ground state εn (0), together
with the nonlinear inductive potential EJ cos(ϕ)/h̄ω01 . On top of each level we
plot the eigenfunctions, shifted and scaled arbitrarily. (c) Relative anharmonicity
of the transmon qubit αr = ω12 /ω01 − 1. (d) Energy levels of a transmon when we
can tune the Josephson energy with a dc-SQUID, as EJ () = EJ cos(/ϕ0 ), for
EJ (0)/EC = 50.
n + ng . The value of n is not the charge, but the index of the band, and ng
represents the quasimomentum of the Bloch wavefunction (cf. Figure 6.4a,
dot-dash line).
(3) When the confinement is deep, one or more bands fall within the periodic poten-
tial EJ 4EC n2 . The splitting between neighboring bands is so large that
they become flat. This limit corresponds to a particle hopping among the min-
ima of the inductive potential cos(φ). The width of the band |En (ng = 1/2) −
En (ng = 0)| is the tunneling amplitude between wells, which decreases expo-
nentially with the barrier height V0 ∼ EJ . In other words, when EJ /EC is very
large, the sensitivity of En (ng ) to the value of ng is exponentially small com-
pared to the original charge qubit (cf. Figure 6.4a, solid line). In this limit, we
can ignore the bias or set it to ng = 0.
The transmon qubit implements the third limit, in which EJ /EC 50 and three
or more bands fall inside the Josephson inductive potential (cf. Figure 6.4b). This
ratio decreases exponentially the sensitivity to electrostatic fluctuations, which is
extremely good for the qubit lifetime, but this advantage comes at a cost: It results
in a qubit with a reduced anharmonicity.
The charge operator q̂ plays the role of the dipole moment of the transmon,
connecting neighboring states that differ in parity – i.e., it allows transitions from
0 to 1, 1 to 2, . . . . This operator is explicitely coupled in (6.34) to a field ε(t) that
contains the gate voltage ng or qg . The source of the field ε can be a classical,
coherent pulse, which we treat as a complex, time-dependent pulse with the tools
from Section 6.1.3. However, the field may also have a quantum origin and be
generated by a resonator, a cavity, or a waveguide, giving rise to the cavity-QED
setups in Chapter 7.
It may seem counterintuitive that, since the transmon bands are flat and indepen-
dent of ng , an electric field may have any kind of effect on the system. However,
notice that the linear term can be approximately absorbed with a unitary redefinition
of the Fock operators, â → â +iε/ω01 . This reveals that, up to first order, the energy
levels really do not depend on ng , but this external field may induce a unitary change
in the transmon eigenstates ψn (ϕ). In other words, starting from our model
If we project these corrections back onto the basis {|ψ0 (ng ) , |ψ1 (ng )}, we recover
the term proportional to q̂ or i(â † − â).
d2
u(z;q) + [a − 2q cos(2z)]u(z;q) = 0. (6.40)
dz2
Given a value ν, the Mathieu equation admits a complete set of orthonormal solu-
tions labeled by nonnegative integer k such that
π
uk (z;q)uk (−z;q)dz = δkk (6.41)
0
uk (z + π ) = ei(ν+2k)π uk (z), n ∈ {0,1,2, . . .}. (6.42)
These solutions are the Floquet functions uk (z) = meν+2k (z;q), where ν + 2k is the
Floquet characteristic exponent of the function uk (z) after a period π . Note that these
functions have the required boundary condition if we identify 2zϕ. This suggests
that we rescale and shift the phase variable ϕ = 2z + π , with which 2∂ϕ = ∂z ,
so that
ε EJ
∂z2 + − cos(2z) ξ (ϕ) = 0. (6.43)
EC EC
From here it follows the identification of physical quantities to Mathieu equation
parameters
ε EJ
a= ,q= , and ν = −2ng (6.44)
EC 2EC
and the eigensolutions and eigenenergies
eing ϕ ϕ + π EJ
ψn (ϕ) = √ me−2ng +k(n,ng ) ; , (6.45)
2 2 2EC
En = MA (r,q), (6.46)
where MA (r,q) is the Mathieu characteristic function for even functions. It is cus-
tomary to define the index k as a function of n and ng that produces ordered energies:
En ≥ Em for all n > m. The function for implementing this ordering is rather con-
voluted. An expression that works for all values except 2ng ∈ Z is7
1
k(n,ng ) = −round ng + (−1)floor(ng ) −1 + (−1)n (1 + 2n) . (6.47)
4
Here round(x) is the nearest integer to the real number x, and floor(x) is the integer
immediately below x. The eigenenergies are given by the Mathieu characteristic
function MA (r,q) with the preceding identifications.
These formulas allow us to compute the eigenenergies of the transmon qubit
for different ratios of EJ /EC , as shown in Figure 6.4a. Problem 6.11 shows that
we also have access to the wavefunctions, a fact that is useful for practical studies
of transition matrix elements and coupling strengths. Figure 6.4b illustrates some
eigenstates for a ratio EJ /EC = 16, where the two lowest bands are already deep
in the noninductive potential. Notice how ω12 , the excitation energy from the first
excited level, becomes very similar to ω01 , the qubit frequency. In this case, the
anharmonicity is α = 8.2%.
with a carrier frequency ω01 and a smooth envelope w(t) that switches the pulse on
and off over a time T . The Fourier components of this pulse spread approximately
over a frequency band ω01 ± 2π/T . Thus, we expect that for long enough pulses,
2π/T α, our external drive will only be able to couple the states |0 and |1,
suppressing leakage outside the qubit space {|0 , |1} to unwanted states |2,|3, etc.
In the limit of finite bandwidth controls and no leakage, we can project the nonlin-
ear oscillator model onto the qubit manifold spanned by the ground and first excited
state. Up to constants
1
H = h̄ω01 σ z + h̄εσ y . (6.49)
2
This is essentially our qubit model (6.2) up to an irrelevant change of basis σ x → σ y
(see Problem 6.1). Following Section 6.1.4, the transmon is a fully controllable
qubit, in the sense that oscillating electromagnetic fields ε(t) with tunable ampli-
tude, frequency, and phase can be used to implement any qubit operation – even
without tuning the gap ω01 .
As in the charge qubit, frequency tunable transmons rely on a dc-SQUID with
some magnetic flux to control the effective value of the Josephson junction
EJ () = EJ (0) cos(/ϕ0 ) and adjust the qubit gap (cf. Figure 6.2d). In particular,
the external flux can only decrease the plasma frequency from a fixed, maximum
√
value approximately as ω01 () = ω01 (0) cos(/ϕ0 ).
Frequency tuning finds its use in bringing the transmon qubit on-resonance to
other circuits, such as resonators and other qubits, to implement quantum operations
and gates (e.g., Section 8.3.3). There are two limitations in this type of control.
The first one is that we cannot decrease EJ () too much, because then the ratio
EJ /EC becomes so small that decoherence kicks in and spoils the setup. The second
problem is that the Fock operators and the basis states implicitely depend on the
value of the Josephson energy – see (6.33). Nonadiabatic changes of the Josephson
energy must be taken into account by considering the explicit time dependence of
the Hamiltonian, via ∂t â,∂t â † and induces corrections into the qubit model.
Finally, regarding the detection of the qubit, there are no direct measurement
schemes for the transmon. Instead, transmon qubits will be typically connected to
one or more resonators, which we will use to indirectly probe the state of the qubit
in a nondemolition fashion. Those methods, which are general to all types of qubit,
are discussed in Section 7.4.3.
The transmon, in any of its incantations, is the most used qubit in all kinds
of superconducting experiments, from those working with photons and qubits to
reproduce quantum optics, to quantum information experiments with tens of qubits
by companies such as IBM, Google, or Rigetti Computing. It is a very popular type
of qubit for obvious experimental reasons. First of all, it is a large qubit, with a
simple design and easy to fabricate – experimentalists can basically cut and paste
masks from one lab to another. Second, the qubit has a very reproducible design,
because its properties are not extremely sensitive to the underlying parameters. For
√
instance, the plasma frequency goes as 8EJ EC , so that small deviations in the
Josephson energy – due to fluctuations in the junction’s thickness or size – appear
as small corrections of 5% in the final qubit. Finally, transmon qubits have very
competitive decoherence times, ranging from T2 ∼ 20 μs up to 0.3 ms – and maybe
better, by the time you read this book.
Interestingly, for a transmon bigger is better. We can increase the coherence
time, simplify fabrication and reproducibility, and enhance the Josephson energy
by making the samples larger. This realization led to the development of the three-
dimensional transmon by Paik et al. (2011), which is nothing but a very large trans-
mon, with a size of millimeters, fabricated not on a chip, but on a portable sapphire
substrate. This device is a tiny object that can be moved around and inserted into dif-
ferent experimental setups, including 3D cavities. Because the transmon frequency
can be verified outside the experiment through simple resistive measurements,8 and
because transmons are mobile and can be manually interchanged, the practical need
for tunability may disappear from experiments, opening the door to setups with 3D
transmon qubits that are extremely long lived, with T2∗ 0.092 ms in state-of-the-
art setups (Rigetti et al., 2012).
On the negative side, the operation speed of the transmon is ultimately limited
by its anharmonicity. For a qubit gap ω01 6 GHz, anharmonicity α 300 MHz
and T2 = 40 μs, a rough estimate would set the limit on 103 single-qubit rotations
within the qubit’s lifetime. That is reasonably large for to-date algorithms, but it may
fall short for large-scale computations. The situation gets even worse if we consider
8 The resistance of the junction at room temperature gives an idea of E , while E can be well estimated from
J C
the original design due to geometric considerations.
the interaction speed of the transmon with other circuits, such as cavities and other
qubits. In this limit, we can find operation times of 4−40 MHz, which further lower
the size accurate quantum computations. In the following chapter, we will meet the
flux qubits, a design that solves both problems, enlarging their anharmonicity and
allowing for stronger couplings to other devices.
Figure 6.5 Two flux qubit designs. (a) Equivalent circuit for an rf-SQUID qubit,
such as the one in Bennett et al. (2007). (b) Three-junction flux qubit circuit (left)
and photograph of the same qubit inmersed in a dc-SQUID for readout. Picture
courtesy of Clarke and Wilhelm (2008), adapted with permissions.
Figure 6.5 shows different incantations of a flux qubit. The device is essentially a
superconducting loop with one or more Josephson junctions. The loop is threaded
by a magnetic field, which modifies the inductive energy, causing the appearance of
two degenerate ground states with distinguishable current distributions. These will
be our qubit states.
The doubly degenerate ground state is explained by the fluxoid quantization.
The magnetic flux that passes through the qubit causes a discontinuity in the flux
variables along the superconducting circuit δφloop , which satisfies
∇φ · dl = δφloop = 0 × Z + ext . (6.50)
C
the capacitance to act, we also need to balance the Josephson energies, lowering the
energy barriers just enough to activate quantum tunneling.
1 2 1 2 1
H q + φ − EJ cos (φ + ) . (6.51)
2CJ 2L ϕ0
Figure 6.6 Energy levels of the rf-SQUID (a) exactly at the maximal frustration
point = 0 /2 and (b) 10% away from this value. Note how the shifted cosine
potential (long dashed lines) conspires with the inductor potential (parabola with
short dashed lines) to create two local minima, L and R in the frustrated case
Note also how a small change in the flux implements a relative energy shift of
the two current states |L and |R. We have exaggerated the plots using a ratio of
inductances L/LJ = 5 of the loop versus the junction.
where LJ = ϕ02 /EJ is the effective inductance associated to the Josephson junction
(see Section 3.7). Apart from the trivial maximum ϕ = 0, this equation reveals
two absolute minima whenever β = LJ /L < 1. As an example, Figure 6.6a shows
two absolute minima labeled L and R, both close to the ideal frustrated solution
φL,R ± 12 0 .
Let us complete our model, adding some “quantumness” to the classical picture
of well-defined flux states. We first consider each of the local minima separately. If
they are deep enough, each of them may be approximated by a harmonic potential.
When combined with the capacitive term, the inductive potential gives rise to two
effective LC resonator ground states,
1
HL,R 4EC n̂2 + EJ (ϕ − ϕL,R )2, (6.53)
2
each placed around a different minimum φL and φR . The approximate wavefunc-
tions of these two solutions are Gaussians:
) *
1
ψL,R (ϕ) = ϕ̂|L,R ∝ exp − 2 ϕ − ϕL,R . (6.54)
2σϕ
The wavefunctions of the |L and |R states are centered on the left and right poten-
tial minima, and correspond to the left- and right-moving currents from Figure 6.5a.
√
The width σϕ ∼ 8EC /EJ is both the uncertainty of the phase variable and the
spreading of the wavefunction. Note how it depends on the curvature of the inductive
potential, which is dominated by the contribution from the Josephson energy.
10 As explained in Chapter 9, we talk about frustration when we have a functional that is a sum of terms, but we
cannot find the minimum of that functional by looking at each term separately.
At low energies, we assume that the qubit remains close to the ground states |L
and |R. We can project the complete Hamiltonian, without approximations, onto
the subspace spanned by these two states. Note that |L and |R are not orthogonal.
The wavefunctions ψL,R (ϕ) spread outside their respective potential minima, over-
lapping on the potential barrier. Consequently, the projected Hamiltonian includes
terms that enable the tunneling of states in one well |R to the other one |L, and vice
versa. We quantify this probability amplitude with the tunneling matrix element:
h̄t = L|H |R = ψL (ϕ)∗ H ψR (ϕ)dϕ ∝ EJ e−(ϕR −ϕL ) /σϕ = 0.
2 2
(6.55)
Since states |L and |R are not strictly orthogonal, they do not form a convenient
basis. It is much better to work with new qubit states that are superpositions of the
original ones:
1 1
|0 = √ (|L − |R), and |1 = √ (|L + |R). (6.57)
2 2
These states produce a qubit Hamiltonian in the usual form:
The model contains a fixed, minimum energy gap = 2t that depends on the
tunneling, and an adjustable magnetic field bias h̄ε = μδ that is used to implement
single-qubit unitary operations (see Section 6.1.4).
The rf-SQUID has a long history that predates its uses for quantum information
processing (Clarke and Braginski, 2004). The circuit itself was conceived as an
ultrasensitive magnetometer (Silver and Zimmerman, 1967). It was only much later
identified as a qubit candidate due to its two-level subspace and relatively large
anharmonicity.
The rf-SQUID is still used in various devices, including the D-Wave quantum
annealer or quantum optimizer. Its large size can be an advantage. By stretching the
superconducting loops (Harris et al., 2007), a single qubit may physically overlap
or come into contact with many other qubits. This is used to create large lattices of
thousands of qubits with large connectivity over long distances. The resulting setups
can simulate complicated spin Hamiltonians and implement sophisticated quantum
optimizations, as explained later in Chapter 9.
The Hamiltonian (6.58) reveals the two main problems of the flux qubit. First, the
qubit gap ∝ 2t is exponentially sensitive to the Josephson energy. Small changes
in EJ , due to differences in the junction’s area or thickness, or due to trapping of
charges or impurities in the oxidized layer, can cause large fluctuations in the qubit
properties, hampering reproducibility – especially when compared to the transmon.
The second problem is decoherence. The bias field of the qubit is generated by
external magnetic fluxes δ, that are transported close to the qubit by different types
of antennas. Unfortunately, as described before when working with cavities, these
cables not only transport the static current we need to create the magnetic flux.
They also act as bath that couples to the qubit, introducing quantum fluctuations
that must be modeled as dephasing and decoherence (see Sections 6.1.5 and 6.1.6).
As in the charge qubit (Section 6.2.3), the optimal operation point of the qubit to
minimize this environment-induced dephasing lays at the symmetry point δ 0
of maximum frustration.
However, even close to the symmetry point, the decoherence of the rf-SQUID
may be inadequate for quantum information purposes. These qubits are large by
design – to have a large coupling strength μ, to overlap with faraway qubits (see
Chapter 9) and to have a large linear inductance – and this size comes together with
a greater sensitivity to stray fields and decoherence. An obvious solution is to work
with smaller qubits, where the frustrated potential is created using only Josephson
junctions, as explained in the following section.
two current states are energetically stabilized in the multiwell inductive landscape
generated just by the Josephson junctions, without any linear inductors.
The resulting qubit is smaller than an rf-SQUID – i.e., a few micrometers versus
150 μm structure in Figure 6.5a. This makes it possible to pack more qubits in the
same setup, and to embed them in exotic structures, such as a transmission line
nanoconstriction, to achieve extreme regimes of interactions with the microwave
field.11 There have also been significant improvements in flux qubit fabrication
and experiments by Yan et al. (2015) have demonstrated coherence times that are
comparable to those of the best 2D transmons T2 40 μs. All this makes the three-
junction qubit the second most used qubit in the field, and usually the one that people
refer to when discussing “flux qubits.”
The circuit of the persistent current flux qubit was derived in Section 4.8 and it
is sketched again in Figure 6.5b. It consists of three junctions. Two junctions are
identical and have the same Josephson energy and capacitance EJ and CJ . The
third junction’s area is smaller by a factor α < 1, leading to a reduced Josephson
energy α × EJ and a smaller capacitance α × CJ . The effective Hamiltonian has
the form
1 1 2
H = q+2 + q + EJ V (φ−,φ+ ) (6.59)
2(1/2 + α)CJ 2CJ −
with a nonlinear potential
− φ+ φ+ φ−
V (φ−,φ+ ) = αEJ cos − 2EJ cos cos . (6.60)
ϕ0 2ϕ0 2ϕ0
This Hamiltonian depends on the total flux jump across the two bigger junctions
φ+ = φ1 + φ2 . Note how we have eliminated the flux jump across the smaller junc-
tion, using the fluxoid quantization, φα = − φ+ .
Once more, the flux qubit Hamiltonian joins two competing potentials. A global
and fixed inductive potential ∝ EJ cos(φ+ /2ϕ0 ) favors a global minimum at
φ+ = φ− = 0. On top of this, we superpose a cosine potential that can be displaced
along the φ+ axis using the external flux . When = 12 0 , this potential favors
two minima with opposite signs of the flux.
Figure 6.7 shows two landscapes of the inductive energy: (a) without frustration
and (b) right at the maximum frustration point = 12 0 . The local minima of
energy lay always along the line φ− = 0, where the bigger junctions share the same
current φ1 = φ2 . However, in the frustrated case we find two minima along the φ+
axis. These minima correspond to two directions of a similar current, i.e., positive
11 See, for instance, the experiments by Forn-Díaz et al. (2010) and Niemczyk et al. (2010) based on the ideas by
Bourassa et al. (2009), or more recent experiments by Forn-Díaz et al. (2016) demonstrating ultrastrong
coupling between flux qubits and propagating photons.
Figure 6.7 Inductive energy and eigenstates of the three-junction flux qubit.
Nonlinear inductive potential (6.60) with (a) α = 0.9 and no external flux and
with (b) α = 0.9 and /ϕ0 = π, together with the respective cuts along φ1 = φ2
(c,d). (e) Energy levels of the three-junction flux qubit with anisotropy α = 0.7. (f)
Energy gap between the “qubit” states E1 − E0 and between the first and second
excited states, E2 − E1 , as a function of the anisotropy, for = 0 /2.
flux jumps and right-moving currents for φα,1,2.+ > 0, and negative jumps or left-
moving currents for φα,1,2,+ < 0.
The persistent current qubit is rather complicated and does not bend well to
analytical estimates. As explained in Problem 6.13, a numerical method that
works well is to diagonalize the qubit Hamiltonian in a basis of number states,
defined on the space of the ϕ± = φ± /ϕ0 variables. Following the techniques from
Section 4.9, and with a moderate number of states, we can get very good estimates
of the eigenenergies and eigenstates of the three-junction qubit for all values of the
frustration parameter .
Figure 6.7e shows the spectrum for a typical junction anisotropy α = 0.7, and
the usual ratio between inductive and capacitive energies EJ /EC = 1/50. Right on
the symmetry point, the spectrum develops two states that are very close and form a
very good candidate for a qubit space, |0 , |1. A proper inspection of those states
reveals that they are superpositions of two current states with opposite directions,
associated to opposite values of φ+ , as in (6.57). These persistent currents are the
ones giving the name to the qubit, whose original motivation was to show a quantum
superposition of two macroscopically distinguishable states (Mooij et al., 1999), in
the spirit of Schrödinger cat–type experiments.
Figure 6.7e shows the splitting between these states h̄ω01 ∼ E1 −E0 , and the gap
separating them from higher excitations h̄ω12 ∼ E2 − E1 . These values may seem
small, but Josephson energies for a multiple-junction qubit are pretty large, around
100 GHz, and a gap in the range 0.05–0.1EJ – or 5–10 GHz – is acceptable for most
experiments. One conclusion from this plot is that the relative anharmonicity of our
qubit is very large. For realistic parameters of the qubit, it is easy to get factors
of ω12 /ω01 ∼ 1000% or larger, orders of magnitude better than a transmon. The
same plots show the exponential decay of qubit energy splitting – i.e., the tunneling
amplitude h̄t – with the barrier between potential wells ∝ αEJ .
Figure 6.7e reveals that we have not only good qubit states, but also a full qubit
Hamiltonian as in (6.2). Note how small changes in the total flux introduce shifts in
the energy levels, recreating the qubit hyperbola curves from Section 6.2.3. A more
careful analysis reveals that the flux qubit behaves similarly to the rf-SQUID and
follows the qubit model:
1
= (E1 − E0 )= 1 0 , and ε μ − 0 . (6.61)
2 2
The effective magnetic dipole μ can be estimated from the numerical simulations,
fitting the energy levels to hyperbola, or semiclassically, studying the energy shifts
of the minima as a function of an external magnetic field (see Problem 6.12).
Figure 6.8 Qubit–qubit dipolar interactions. (a) Flux qubits and (b) transmon
qubits have an associated magnetic or electric dipole moment. These moments
interact with each other through the equations of electromagnetism, ∝ d1 d2 and
y y
∝ μ1 · μ2 , giving rise to effective qubit couplings J σ1x σ2x or J σ1 σ2 . Sometimes,
as in the magnetic case (c), an auxiliary circuit can be used to mediate and tune the
interaction.
14 Note that the average value I˜ is not exactly the supercurrent of the junctions themselves. For instance, in the
c
rf-SQUID we find I˜c = |φL,R |/L, in terms of the equilibrium value of the flux for either potential well (see
Problem 6.15).
15 For two dipoles μ , their interaction energy goes as E = −(μ /4π |r|3 )(3(μ · r)(μ · r) − μ · μ ) with
1,2 0 1 2 1 2
their separation r. Since both our dipoles sit on the same plane, we expect E = +μ0 μ1 μ2 /4π |r|3 .
16 For instance, in the three-junction flux qubit I
1,2 ∝ αIc sin(φ+ /ϕ0 ).
that is very similar to the one from an electric dipole di ∝ q̂i . From the classi-
cal theory of electrostatic interactions, we expect a contribution to the energy that
grows with the product of those dipoles and decreases with their separation. Unlike
the magnetic interaction, there is great variability on the sign and strength on the
interactions,17 based on the relative orientations of the qubits and their separation,
but we may write something like
( x x
1 J σ1 σ2 , for charge qubit,
Hint = ± q̂1 q̂2 ∼ y y
(6.64)
Ceff J σ1 σ2 , for transmon.
The sign and the effective mutual capacitance Ceff can be determined by the micro-
scopic model. The coupling operator may also change depending on our convention
for the charge basis: In the superconducting island, q̂ ∝ σ x , while for the transmon
with the notation from (6.33), q̂ ∝ i(|10| − |01|) ∼ σ y . Sometimes these dif-
ferences are irrelevant, and may be suppressed with a local unitary transformation,
but other times they are not, leading to unequivalent models – e.g., imagine two
transmons with SQUIDs interacting also inductively through their loops!
We may be a bit more rigorous in our treatment of the capacitive interactions by
considering the effective circuit in Figure 6.8c. The complete model includes both
the self-capacitance of the transmon or charge qubit CJ , and a mutual capacitance
between both qubits Cg . The whole circuit may be described using the flux degrees
of freedom in the two superconducting islands, with a capacitance matrix:
CJ + Cg −Cg
C= , (6.65)
−Cg CJ + Cg
where the term 12 Cg (φ̇1 − φ̇2 )2 causes Cg to appear both in the diagonal, renormaliz-
ing the effective capacitance of each qubit, and in the off-diagonal, with something
that resembles an interaction. This structure is preserved in the effective Hamilto-
nian, which depends on the inverse capacitance matrix:
⎛ 1+Cg /CJ ⎞
Cg
2
2
CJ CJ Cg
C −1 = ⎝ ⎠+O . (6.66)
Cg 1+Cg /CJ CJ
CJ2 CJ
We have a dipolar interaction, but the mutual capacitance Ceff = CJ × (CJ /Cg )
differs from Cg and we can only ignore the renormalization of the qubits when in
the limit of weak interactions Cg /CJ 1.
17 For two electric dipoles, the electrostatic energy is (d · d − 3(d · r)(d · r)/r 2 /(4π r 3 ) as a function of
1 2 1 2 0
their separation r.
18 This approximation or way of thinking is quite common in physics. It describes how electrons mediate an
interaction between atoms in a molecule, or how phonons bind electrons to form Cooper pairs.
19 Higher powers of Pauli matrices are either the identity, or the matrices themselves. Therefore, the function
ESQUID can only produce terms that are proportional to σ1,2 x or to the product σ x σ x , as seen by expanding
1 2
the function to arbitrary powers in a Taylor series.
20 Ideally, ∂E
SQUID /∂ 0. If not, this represents a shift of the qubit gaps, which can be included in the theory.
monitoring the coherence times of the qubits, as well their natural frequencies,
quality of measurements, fidelity of operations, etc.
An ab initio determination of how one or more qubits evolve can be a rather
cumbersome problem. Chapter 8 will introduce methods for a self-consistent tomo-
graphic reconstruction of how a density matrix evolves under a general control.
Fortunately, in absence of external controls, qubits tend to follow a simple decoher-
ence model that mostly contains relaxation and dephasing (6.19). This model can
be fully determined by calibrating the values of the relaxation time T1 and of the
dephasing time, T2 or Tφ , in different experiments.
To calibrate the relaxation time, we study the evolution of the excited state
|1. We reset the qubit to |0, flip the state of the qubit with a π pulse, wait
for a time t, and measure the excited population P1 (t) = 12 1 + σ z . As shown
experimentally (Steffen et al., 2010), even if we make small errors in the preparation
and measurement, the excited population can be fit to an exponential curve
P1 (t) = P1 (0) exp(−t/T1 ), which gives the value of T1 .
To obtain the value of the decoherence time, we must study the evolution of a
quantum superposition, using a protocol known as Ramsey interferometry. We start
with the |0 state, but now implement an incomplete rotation around σ y with the π/2
angle, preparing the state √12 (|0 − |1). We wait for a time t, repeat the rotation,
and measure σ z . The excited state population now follows a damped oscillation
∗
P1 (t) = 12 (1 + cos(t)e−t/T2 ).
The coherence time is always limited by relaxation, T2∗ = 2T1 . In practice, this
time is shorter because of dephasing and other sources of decoherence Tφ , as shown
in (6.20). In those cases, we can further separate and partially cancel the contribu-
tion from dephasing, using spin-echo to obtain a slightly larger effective coherence
lifetime T2echo (see Section 6.1.5).
Figure 6.9 illustrates the experimental coherence times T2∗ (or T2echo , depending
on the work) for a large variety of qubits. There is an overall positive trend, where
coherence times almost double every year,21 starting from the charge qubit in
Nakamura et al. (1999), until recent experiments with 2D transmon qubits (Place
et al., 2021) and photon-encoded qubits (Rosenblum et al., 2018). We see significant
jumps due to the placement of qubits in microwave resonators (Wallraff et al.,
2004), the introduction of large three-dimensional transmons (Paik et al., 2011), the
development of capacitively-shunted flux qubits (Yan et al., 2016), the encoding of
qubits in photonic degrees of freedom (Wang et al., 2016; Rosenblum et al., 2018),
21 This has been deemed Schoelkopf’s law, by analogy with Moore’s law for Silicon-based integrated chips.
γ
γ
10−3 E
E
T F t
10−4 T J
T t t t T
t t
J
T t t t t t t
T2∗ (seconds)
t G
t
10−5 F G
t t t
C t t t
10−6 t G
Q C t
f J
10−7 γ P
J J
10−8
C C J
2000 2005 2010 2015 2020
Year
Figure 6.9 Qubit coherence times T2∗ (or T2 measured after spin-echo) across the
years. Letters signify: charge qubit (C), flux qubit (J), fluxonium (f for 2D, F for
3D), phase qubit (P), transmon (t for 2D, T for 3D), photon encoded qubit (γ),
error-corrected bosonic qubit (E), and gatemon (G).
and recent work to improve the surface properties of 2D transmons (Place et al.,
2021). An important question is whether this trend will saturate, or whether we will
discover new and better ways to design, fabricate, and integrate qubits in quantum
circuits.
Exercises
6.1 Show that a qubit Hamiltonian of the form H = 2 σ z + 2ε σ x can be mapped
to other forms, such as H = 2 σ z + 2ε σ y and H = 2 σ x + 2ε σ z , by a choosing
a new basis of qubit states |0̃ = U |0 and |1̃ = U |1. Compute the unitary
transformation U in those two particular cases.
6.2 A single-mode harmonic oscillator with Gaussian Hamiltonian and linear
operations cannot implement a qubit. Assume a general model of the form
H = ωa † a + (f (t)a + f (t)∗ a † ) with a general driving f (t) ∈ C. Prove that
starting from state |0 it is never possible to create a perfect Fock state
|1. Discuss how this changes when we add a nonlinearity U a † a † aa to the
Hamiltonian, with U > 0.
6.3 Let us study the limit of quasistatic dephasing. We will use a diagonal
perturbation to the dynamics H ∼ h̄2 ( + δω)σ z , but the random perturbation
δω has no dynamics: It is constant through each experimental realiza-
tion, but fluctuates from experiment
√ to experiment with Gaussian distribution
P (δω) = exp(−δω /2σ )/ 2πσ .
2 2 2
(1) Compute completely positive map εt (ρ) describing the evolution of the
qubit. Compare the resulting map with (6.13). Show that coherences now
decay as ρ01 (t) ∝ ρ01 (0) exp(−t 2 σ 2 /2).
(2) Show that the evolution of the density matrix ρ(t) can be written in
Lindblad form (6.15), but now the equation is not Markovian, and the
dephasing rate depends on time γφ = tσ 2 .
(3) Assume that we apply a spin-echo refocusing protocol (6.16), which
consists of instantaneous flip operations UBB = σ x , repeated with period
TBB . Show that the density matrix after two pulses ρ(2TBB ) is the one we
would have obtained by evolving with H = 12 σ z .
6.4 Let us solve the dynamics for a qubit with Hamiltonian H = 12 σ z , in contact
with a bath at finite temperature (6.17).
(1) Using the master equation to estimate ∂t ρ, prove that the evolution of an
observable for such a qubit is
d γ
O = −i [O,H ] + σ + [O,σ − ] + [σ +,O]σ − . (6.73)
dt 2
(2) Next, derive the Bloch equations equations for σ z and σ ± . Solve
those equations in the zero temperature limit n = 0.
(3) Use those expected values to reconstruct the Bloch vector (2.25). Compute
the norm of the vector, and show that it is contracting and the state remains
inside the sphere.
(4) Using the Bloch vector S(t), prove that the density matrix evolves as
1 − P1 + (e−t/T1 − 1) [P1 − ρ11 (0)] ρ01 (0)eit−t/2T1
ρ(t) = ,
ρ10 (0)e−it−t/2T1 P1 + e−t/T2 [ρ11 (0) − P1 ]
where P1 = n/(1 + 2n) = limt→+∞ ρ11 (t) is the asymptotic excitation
probability of the qubit and T1 = (1 − 2P1 )/γ gives the timescale |T1 |
at which ρ(t) converges to its asymptotic solution. Relate this solution to
the zero temperature limit of pure relaxation.
6.5 Write down a master equation that combines a qubit gap h̄, dephasing with
a rate γφ and relaxation with a rate γ . Solve the evolution for an initial density
matrix ρ0 . Study the evolution of ρ01 (t) = 0|ρ(t)|1 and prove (6.20).
6.6 Charge qubit truncation. Let us consider the effect of small potential
perturbations ng = 12 + x onto a charge qubit with EJ /EC, x 1. We
will rely on the degenerate perturbation theory from Section A.3.2, identifying
our target subspace as the qubit space, P0,α=0 = |00| + |11|. Show that,
up to second order in EJ /EC , it only the coupling to the degenerate subspace
{|−1 , |2}. Using the series expansion (A.36), you should obtain a diagonal
effective Hamiltonian:
2 EJ2
Heff = −EJ σ x + 4EC 1
+x − |00|
2
8EC (1 + 2x)
2 EJ2
+ 4EC 1
−x − |11| .
2
8EC (1 − 2x)
Does this model differ from (6.25)? What happens when x = 0? And when x
is so small that the diagonal terms can be linearized? Can we distinguish the
effective theory from the truncated one?
6.7 Following the experiment by Nakamura et al. (1999), we prepare a charge
qubit at a point ng 0, cooled down to charge neutrality |0. We then abruptly
move the experiment to the symmetry point ng = 1/2 for a time T , after
which we return to ng = 0 and measure the charge in the island. What is
the Hamiltonian of the qubit at ng = 1/2 in the charge basis? How does
the quantum state evolve, and what is the final state of the qubit at time T ?
What is the probability that we find one excess Cooper pair at the end of
the experiment?
6.8 Let us take the charge qubit by Nakamura et al. (1999), where the energy
gap 80 μeV. We are going to analyze its sensitivity to the environment
assuming that there can be trapped charges around the qubit at a distance of
d ∼ 4 μm. As a toy model, we assume that that all of the qubit on average
experiences the potential induced by this electron V = e/4πε0 d. What is the
extra energy acquired by the |1 charge state relative to the |0? How do you
write this correction into the Hamiltonian? How much does ng change by
this external influence? What about the energy levels E? How much do
they change?
6.9 Diagonalize the oscillator part of the transmon Hamiltonian in (6.32), disre-
garding the contribution of the nonlinear terms, so as to estimate the transmon
properties.
(1) Estimate transition frequency h̄ω01 .
(2) Express the charge operator and flux operator in the oscillator basis. What
are the matrix elements of those operators between the ground and excited
state? Compare with Section 6.3.4.
(3) Using nondegenerate perturbation theory, estimate the negative anhar-
monicity due to the quartic term ∝ EJ φ 4 .
(4) Connecting (6.33) to the one for an LC resonator, what is the correspond-
ing impedance of the transmon?
6.10 Let us write down the transmon Hamiltonian for the lowest three levels with
an external, time-dependent perturbation:
H = (ω12 + ω01 ) |22| + ω01 |11| + (ε12 (t) |21| + ε01 (t) |10| + H.c.) .
22 As a curious detail, our definition of z and of q makes q positive, and later versions of Mathematica seem to
behave more stably when q ≥ 0.
Plot its behavior as a function of the EJ /EC ratio both for ng = 0 and
ng = 1/2. Hint: Mathematica also has the derivatives ceν (z) and seν (z).
6.12 Let us look at the three-junction persistent current qubit, from
Section 6.4.3. In the seminal experiment by van der Wal et al. (2000), the
two biggest junctions had critical currents and capacitances Ip 570 nA and
C 2.6 fF; the smallest junction was so by a factor α = 0.82, and the total
qubit had an area of about 10 μm2 . Using the circuit model from Section 4.8
and the preceding theory, answer these questions:
(1) How large is the energy barrier between the two potential wells along
φ− = 0? How does it compare to the total depth of the potential?
(2) Estimate the gauge-invariant phase ϕ across the big junctions and relate
it to the current around the loop.
(3) Write down the inductive energy Eind of the three Josephson junctions
operating at the flux qubit point, = 0.5ϕ0 + B , where B is the flux
induced by a uniform magnetic field perpendicular to the loop. Estimate
the qubit’s effective dipole as the change in energy with respect to B,
μ ∝ ∂E/∂B.
(4) How does the effective magnetic moment (in Bohr magnetons) compare
to the permanent magnetic moment of atoms such as chromium? Discuss
the reason for the difference.
6.13 We can diagonalize the flux qubit Hamiltonian (6.59) using the number-phase
representation from Section 4.9. Note that we have to use two integer numbers
|n+,n− , one for each flux degree of freedom.
(1) Adimensionalize Hamiltonian H expressing all terms as a function of
ratio EJ /EC .
(2) Write down the Hamiltonian H in the two-dimensional number basis
|n+,n− . Your result should produce diagonal expressions for q± and
introduce charge increments or decrements associated to cos(φ± /2ϕ0 )
and similar terms.
(3) Translate the previous expression into a small Mathematica or Matlab
program that works with a truncated space, n± ∈ {−N, . . . ,N}. Typically,
N 10 will suffice.
6.15 Flux qubit current operator. Solve approximately the equations for the
minima of the inductive potential φL,R , in the limit of large L/LJ ratio. Show
that in this limit, the phase ϕ can be written as ±(π −ε), with a small correction
ε ∝ L/LJ . Use the derivation in Section 4.7.1 to write down the current
operator for the rf-SQUID in terms of the flux jump along the inductor φ̂.
Estimate the matrix elements of the current operator on the |L and |R states.
Show that it is approximately diagonal in these states (6.63). Using the change
of basis (6.57), prove that the current operator has approximately the form
I˜c σ x , finding the expression for I˜c = |φL,R |/L.
6.16 The rotating wave approximation also applies to the dynamics of qubits.
Let us consider the interacting qubit model (6.67), in the limit of resonant
qubits 1 = 2 .
(1) Use the decomposition σ x = σ + + σ − to find what would be the RWA
Hamiltonian HRWA associated to this model.
(2) Show that H can be decomposed into two 2 × 2 matrices acting in the
even and odd subspaces, {|00 , |11}, and odd {|01 , |10}. Use this to
compute all eigenvectors and eigenenergies.
(3) Analyze the eigenvectors and show that in the limit |g| ||, this matrix
has approximately the same eigenstates as the Hamiltonian HRWA that
results from neglecting the counterrotating terms. Estimate the correction
to the states introduced by not neglecting the counterrotating terms and
show that the associated excitation probability scales as |J /|2 .
6.17 Return to the interacting qubit model (6.67), but now assume that the qubits
are far apart in frequency space, δ = 1 − 2 |J |.
(1) Show that the Hamiltonian H still can be decomposed into even and odd
subspaces, as in Problem 6.16.
(2) Show that, in the limit 1 = 2 the spectrum is formed by four different
eigenenergies. Analyze the same problem using second-order nonde-
generate perturbation theory (see Section A.3.1). Find out the effective
model for this Hamiltonian in the strongly off-resonant limit, considering
corrections up to O(J 2 /δ). Based on your exact diagionalizations, at
which limit of the interaction does perturbation theory fail?
(3) Solve analytically the Schrödinger equation for two qubits that start
in the state |01 and evolve with the full Hamiltonian (6.67) and with
an effective Hamiltonian where you neglect all interactions. Call those
solutions |ψ01eff
(t) and |ψ01
exact
(t).
(4) Estimate the fidelity F = | ψ eff |ψ exact |2 between the states evolved with
the idealized model, where effective interactions have been disconnected,
and the real one. What reasonable values can J, 1, 2 and t take to
achieve an infidelity 1 − F ≤ 10−1,10−2 and 10−4 , respectively?
(5) Repeat the previous calculation using the qubit parameters from the work
by Barends et al. (2014).
It can be argued that the work by Wallraff et al. (2004) included the two impor-
tant breakthroughs that superconducting circuits needed to become a mainstream
quantum technology. The first breakthrough was the jump in a qubit’s coherence
time that appeared by placing those qubits in the gap of a microwave resonator. The
improvement was so impressive that superconducting qubits regained momentum –
and strongly surpassed other solid state platforms – in the quest for scalable quantum
computers.
The second breakthrough in this work was to show how qubits could interact
coherently with the photons trapped in the resonator that protected the qubit. The
combined qubit–resonator sample evidenced an exchange of photons that was
compatible with the canonical Jaynes–Cummings or Rabi models from quantum
optics. Thus emerged the field of quantum optics with superconducting circuits,
circuit quantum electrodynamics or simply circuit-QED. This field brings to the
lab old ideas from quantum optics, such as quantum superpositions of photons
(Hofheinz et al., 2009), Schrödinger cat states (Vlastakis et al., 2013), and collective
effects in superradiance (Mlynek et al., 2014; Nataf and Ciuti, 2010). A large
body of fundamental results was accompanied with crucial developments in the
preparation, control, and measurement of superconducting circuis. Microwave
resonators became the tool to control and measure superconducting qubits (Schuster
et al., 2005), to create entanglement (Majer et al., 2007) and to assist quantum gates
between qubits (Rigetti and Devoret, 2010). Understanding how this happened is
the goal of this chapter.
We begin this important segment of the book introducing a model for light–matter
interaction between superconducting circuits and microwave photons. Roughly, we
quantize a large circuit with one or more qubits in a microwave transmission line, we
then identify linear and nonlinear excitations – photons and qubits – and establish
parallelisms between the resulting Hamiltonian and atoms in optical environments.
We first apply this model to study the dynamics of a qubit in an open environment,
156
explaining how an atom can emit photons, one at a time, and interact with a
propagating microwave field. We then make a drastic change, placing our qubits
inside microwave resonators. We show that photons in a resonator can interact more
strongly and for longer periods of time with our artificial atom. This leads to simpler,
radically more useful models, such as the Jaynes–Cummings Hamiltonian. We
explore applications of these circuit-QED setups to the engineering and tomography
of qubit and photon states, introducing the dispersive coupling, single- and two-tone
spectroscopy and quantum nondemolition (QND) measurements of qubits.
Figure 7.1 Transmon and flux qubit sitting inside a one-dimensional microwave
transmission line – central conductor at the bottom, ground plate on the top. The
qubits interact preferably with either the voltage or the current transported by the
waveguide, through their electric and magnetic dipoles.
1 If we follow the same convention as before, the transmon would couple using the σ y operator. However, for all
applications discussed in this chapter, we can use a local unitary transformation to transform σ y → σ x .
in gμ · B(x). Another way to reach the same conclusion is to see the currents in the
line as generators of a magnetic field that modifies the flux inside the qubit . As
in Section 6.4, we find this perturbation induces opposite energy shifts of the |L
and |R current states. When we define the qubit states at the symmetry point, the
shift adopts the same dipolar coupling form as for the charge qubit:
The coupling constant ε now accounts for the qubit–line mutual inductance M and
the qubit’s supercurrent intensity I˜c .
k
2cωk l d d
The Hamiltonian is the same for capacitively and for inductively coupled qubits,
with minor changes in the coupling constants:
cap cap h̄ωk ind k h̄
gk = −iG , and gk = iG
ind
. (7.6)
2cd l 2cωk d
The prefactor G encapsulates the microscopic details of the qubit–line coupling. For
the transmon, it is proportional to the qubit–line capacitance Gcap = −2eCg /C .
For a flux qubit, it combines the qubit–line mutual inductance M with the qubit’s
critical supercurrent Gind = M I˜c .
2 The other one, the Caldeira–Leggett model, was used in Chapter 5 to describe losses in a resonator.
The first subtlety is the distinction between continuous and discrete spectral func-
tions. The former are associated to extremely long waveguides,3 supporting all
frequencies and wavelengths. Discrete functions, on the other hand, appear when
we place the a superconducting qubit inside a λ/2 or λ/4 resonator, and we can
resolve the individual eigenfrequencies – with the qubit typically interacting with
just one or few of those modes.
Even in the continuous case, we can also find gapped spectral functions, which
are zero over an extended range of frequencies. A common example is a 3D
microwave guide made from a long tube of aluminum or copper. The waveguide
only supports modes propagating with a frequency that lays above a minimum
energy ωk ≥ ωcutoff , which is needed to excite the transverse electromagnetic profile
of the waveguide’s modes. Below this frequency, the waveguide has no excitations
that can absorb energy from the qubit. When the qubit is placed inside the bandgap,
∈ [0,ωcutoff ], the waveguide acts as a filter that blocks all excitations, protecting
the qubit from decoherence.4
Conventional waveguides on 2D chips are approximately gapless, offering exci-
tations from very low frequencies up to some intrinsic cutoff.5 The spectral function
of such waveguides grows linearly with the frequency of the excitations, describing
an Ohmic spin-boson model:
J QO (ω) = παω1 . (7.8)
This regime is to be contrasted with other polynomial spectral functions J (ω) ωs
with smaller and bigger exponents, called the sub-Ohmic s < 1 and super-Ohmic
s > 1 regimes.
We can verify (7.8), rewriting the coupling constants from (7.7) as
√
|gk | = h̄ παvωk /2d, and inserting them into the definition of the spectral func-
tion (5.63):
2πv παωk
J (ω) = 2×
QO
δ(ω −ωk ) dωk παωk δ(ω −ωk ) ∼ παω1 . (7.9)
k>0
d 2
This assumed a linear dispersion relation, a uniform density of states dωk ∼ vdk
2πv/d, and two directions of propagation for each frequency.
The Ohmic spin-boson model exhibits a rich phenomenology, depending on
the dimensionless ratio α, describing the qubit–photon interaction strength and
the competition between incoherent processes (decay) and coherent processes
3 Experiments don’t allow “infinite” lengths, but a waveguide shorted by an impedance-matched resistor
behaves as an infinitely long medium, too (Devoret, 1995).
4 The qubit can still radiate to free space and into its substrate. These are other environments that are rarely
modeled, but grouped into intrinsic or nonradiative losses.
5 The superconducting gap is at least one possible cutoff.
(qubit rotation with the Hamiltonian). As we summarize in Table 7.1, there are four
distinct phases and behaviors depending on α.
The first region α ≤ 1/2 is where coherent dynamics overcomes decoherence at
short times: The qubit evolves between the excited and ground states, with oscil-
lations that dampen at a rate γ . This dynamics is very well approximated by a
memoryless or Markovian master equation (6.17), with a decay rate that, for small
values α 1/2, is provided by the spectral function γ = J QO () πα. Note
how γ / = πα so that α can be regarded as the relative speed of spontaneous
emission, as compared to the qubit’s intrinsic dynamics .
As α approaches 1/2, the dynamics of the qubit follows a slightly corrected
Markovian dynamics, due to the entanglement between the qubit and its electro-
magnetic environment. We regard the qubit as dressed by a cloud of photons that
slows it down, shifting its frequency to lower values, and changing the decay
rate γ , which is no longer proportional to the spectral function. These corrections
were estimated analytically by Shi et al. (2018) and observed experimentally by
Forn-Díaz et al. (2016).
When the coupling strength increases above α = 1/2, correlations between
the qubit and its surrounding photon cloud prevent any Markovian description.
The dynamics of this strongly correlated system is faster than exponential and
requires very different (mostly numerical) descriptions.6 The theory around this
point describes a qubit decay that is accompanied by the emission of odd numbers
of photons – one, three, five, etc. – with varying combinations of frequencies.
6 Path integral calculations, the Kondo renormalization group, or matrix product states (MPS) (Peropadre et al.,
2013; Shi et al., 2018) provide insight into this region.
In the RWA limit, every excitation of the qubit is accompanied by the subtraction
of a photon from the environment. Such dynamics conserves the total number of
excitations given by
† σz + 1
N̂ = âk âk + . (7.11)
k
2
This is still a relatively simple Hamiltonian, with a ground state that consists of the
vacuum in the bosonic environment and an unexcited transmon, |0 ⊗ |vac. Inter-
estingly, if we allow for cyclic transitions, such as g 01,g 12,g 02 = 0, the number of
excitations will no longer be conserved. This prevents using the theoretical methods
in this chapter, but enables interesting side-effects and processes such as splitting
a photon into two other frequencies (Sánchez-Burillo et al., 2016) or facilitating
interactions between photons at different frequencies (Hoi et al., 2013).
7.2 Waveguide-QED
Waveguide-QED is a relatively new subfield of quantum optics that studies the
interaction between propagating photons confined in 1D systems and few-level sys-
tems. The term includes broad families of experiments. The photonic medium could
be nanofibers, plasmons, or photonic crystals that interact with ultracold atoms,
quantum dots, or color centers. Of course, waveguide-QED also describes the super-
conducting setups from Figure 7.1.
In this section, we introduce various tools that are developed in the context of
waveguide-QED, focusing on the RWA limit with excitations and states |ψ ∈ H1
Figure 7.2 (a) An excited qubit relaxes by emitting one photon in a superposition
of directions in an empty line. (b) Conversely, if we send a photon against a qubit,
it will be partially reflected and partially transmitted, with amplitudes R and T .
that satisfy N̂ |ψ = 1 × |ψ. As sketched in Figure 7.2 in this regime we expect
only two relevant physical processes: the spontaneous emission of a photon by an
excited qubit; and the scattering of a propagating photon by a relaxed qubit. Both
processes can be modeled using a quantum superposition of qubit and photon excita-
tions, known as the Wigner–Weisskopf ansatz. Despite its simplicity, the wavefunc-
tion model captures important concepts, including spectroscopy, Markovianity, and
decoherence.
This ansatz represents the most generic state within the single-excitation sector
of the RWA Hamiltonian (7.10). We therefore make no mistake when we project
the Schrödinger equation for this Hamiltonian onto this subspace, and obtain a
differential equation for the amplitudes of the qubit’s c1 (t) and one for the boson’s
excitations ψk (t) (see Problem 7.1):
i∂t c1 = c1 + gk ψk, and i∂t ψk = ωk ψk + gk∗ c1 . (7.14)
k
As in our study of the lossy cavity, we integrate the equation for the propagating
photon ψk (t), beginning from a time far away in the past t0 → −∞. The math-
ematical procedure is identical7 to the one used in Section B.3.2, replacing the
7 The solutions are formally identical because we are dealing with linear differential equations!
operator a(t) with the complex amplitude c1 (t). This method produces an exact
integro-differential equation for the qubit’s amplitude:
t
∂t c1 (t) = −ic1 (t) − iξ (0,t,t0 ) − K(t − τ )c1 (τ )dτ . (7.15)
t0
The equation contains a source term that describes the incoming photonic field at
the qubit’s position, derived from the initial condition of the waveguide ξ (x,t,t0 ) =
ikx−iωk t
k gk e ψk (t0 ): It also contains an integral with the memory function K(t)
of the bath (5.62). In the Markovian limit, where the bath loses all memory of the
qubit’s past, it is replaced with a complex number K(t − τ )
(iδLamb − 12 γ )δ(t − τ ):
1
i∂t c1 = − i γ c1 + ξ (x,t,t0 ). (7.16)
2
As in the cavity, the Lamb shift δLamb = − describes the slowdown of the qubit
due to the surrounding photons, while the spontaneous emission or qubit decay rate
γ J QO () models the relaxation of the qubit to the ground state.
This expressions combines a term ± (x,t,t0 ) that depends on the initial condition
ψk (t0 ), with a photon emitted by the qubit ∝ c1 (t). We distinguish two solutions
depending on whether our boundary condition is set back in the past – an input field
in (x,t) = ± (x,t, − ∞) – or in the far future – the output field out (x,t) =
± (x,t, + ∞). To move further, we use our recurrent assumptions:
(1) Without loss of generality, we place the qubit at x = 0.
(2) We assume a uniform spacing of momenta, ∼2π/d.
ground state P1 = |c1 |2 = exp(−γ t), emitting a photon into the waveguide. This
phenomenon is called spontaneous emission, and the constant γ is the spontaneous
emission rate of the qubit.
The input–output relations (7.20) allow us to compute the wavepacket of the emit-
ted photon in terms of the qubit’s dynamics. This photon is a truncated exponential,8
with a maximum amplitude at the wavefront and an exponentially decaying tail:
γ (−i −γ /2)(t∓x/v)
± (x,t) = −i (t)
out
e . (7.21)
2v
Once the photon has been fully emitted, the Fourier transform of the wavepacket
reveals a normalized Lorentzian spectrum,9 centered on the qubit’s frequency :
1 γ /2
|ψk |2 = . (7.22)
π γ /4 + ( − ωk )2
2
The full-half-width (FHW) of this profile – i.e., the bandwidth of the photon or
linewidth of the two-level system – is just the emission rate γ .
The spontaneous emission of a photon has been experimentally observed in fre-
quency and in time by Sharafiev et al. (2021). The spectrum of the emitted photon
changes in time, as shown in Figure 7.3. The Lorentzian prediction is only strictly
true at long times t 1/γ , once the photon has been fully emitted. At very short
times 1/t γ , the Markovian approximation still does not apply and the
qubit deposits energy in all frequency modes, with a rather homogeneous spectrum
8 With the Heaviside function (t) = 0 for t < 0, (t) = 1 for t > 0.
9 Compare this with the cavity transmission spectrum from Section 5.5.3.
|ψk |2 ∝ |gk |2 . At intermediate times, the photon looks like a truncated version of
the exponential wavepacket, shortened to a length t from its wavefront. During
those times, the width of the photon goes as ∼1/t γ , fulfilling the so-called
time-energy Heisenberg uncertainty relation – i.e., the shorter time we measure, the
longer the uncertainty in the energy we obtain.
During spontaneous emission, the qubit and the waveguide entangle with each
other. For this reason, if we trace out the waveguide we obtain a quantum state with
less information – i.e., a state that has decohered. Consider a qubit that is initially
prepared in the quantum superposition, in an empty waveguide:
|ψ(0) = (α |0 + β |1) ⊗ |vac . (7.23)
This state will decay in time, producing a photon only when the qubit is in the |1
state. We can derive the dynamics of the qubit-waveguide system using the Wigner–
Weisskopf ansatz, including the state that does change |0,vac. When we trace out
the photonic environment, the resulting density matrix ρqb = trphotons (ρ) is exactly
2
|α| + (1 − |β|2 e−γ t ) αβ ∗ e−it−γ t/2
ρqb = . (7.24)
α ∗ βe+it−γ t/2 |β|2 e−γ t
The qubit begins in a pure state10 in a quantum superposition with coherences ρ01 ∝
α ∗ β. The superposition deteriorates, as evidenced by the decay of the coherences
and of the excited state’s probability. Asymptotically, as t grows to infinity, the qubit
returns to a pure ground state. At this point, all the quantum information from the
qubit has been transferred to a quantum superposition of a propagating photon and
a vacuum state.
The time evolution of the qubit’s density matrix is analogous to the one predicted
for a qubit in a zero-temperature bath (6.19). It exhibits a decay time T1 = 1/γ
and a decoherence time T2 = 2/γ , given by the qubit’s spontaneous emission rate.
As shown in Problem 7.4, this analysis can be generalized, verifying that both
the Wigner–Weisskopf ansatz and the master equation (6.17) provide equivalent
descriptions of the emission process.
The way we read this equation is that an input field coming from the left (a+ = 1) or
from the right (a− = 1) is reemitted by the qubit in all directions, producing nonzero
amplitudes b+,b− .
The input amplitudes and output amplitudes, a± and b± , are related by a unitary
transformation, called the scattering matrix, which contains the transmission and
reflection coefficients, Tω and Rω :
(
b+ Tω Rω a+ Tω = 1 + Rω,
= , with (7.29)
ω = − 2 i( −ω)+γ /2 .
γ 1
b− Rω T ω a− R
Figure 7.4 illustrates the transmission and reflection probabilities, |Tω |2 and |Rω |2 ,
and the frequency-dependent phase shifts experienced by the photon. Note how
11 The term e0+ t is a normalized exponential 1 et/2 with > 0, so that the wave vanishes in the distant past
t → −∞. At the end of the calculations, we take the limit in which approaches 0 from above to recover
monocromatic fields.
(a) (b)
1.0 1.0
|T 2| R/
0.8
0.5
0.6
0.0 T/
0.4
– 0.5
0.2
|R2|
0.0 – 1.0
–4 –2 0 2 4 –4 –2 0 2 4
/ /
Figure 7.4 (a) Tranmitted and reflected probabilities for an incoming photon, as a
function of the photon–qubit detuning δ = ω − . (b) Phase slip (in units of π )
of the transmitted and reflected photon, ϕT and ϕR , respectively.
we have seen other works using the strong-coupling regime to control photons
(Hoi et al., 2012), to engineer strong photon–photon interactions (Hoi et al., 2013),
and to explore all regimes in the Ohmic spin-boson model (Forn-Díaz et al., 2016).
7.3 Cavity-QED
An artificial atom interacts coherently with the photons in a transmission line, but
these photons quickly fly away, leaving us with a decohered qubit. We can increase
the interaction time between the qubit and the photon by cutting the waveguide to
a shorter length, as shown in Figure 7.5. In this setup, the waveguide becomes a
microwave resonator or a cavity, and we are exploring the field of cavity quantum
electrodynamics, or cavity-QED.
500 mm
As explained in Chapter 5, the cavity is a quantum filter that only allows certain
microwave photons in. If the resonator is short, those frequencies are broadly spaced
(see Section 5.2.2), opening large gaps in the spectral function. If we park the
frequency of the qubit to lay deep in those gaps, the qubit is prevented from emitting
photons, and we preserve its quantum state for long times, only limited by the qubit’s
other nonradiative decoherence mechanisms. If instead we tune the qubit close to a
cavity’s resonance, the qubit will be able to emit photons that will bounce back and
forth among the resonator’s walls. The photons now interact with the qubit for longer
periods of time, only limited by the cavity’s decay rate. The Purcell enhancement
of the qubit–photon interaction allows the qubit to endow the resonator with highly
nonlinear properties, creating multiphoton states and engineering sophisticated non-
Gaussian states of light. Similarly, when the resonator is connected to the outer
world through a transmission line, we can use the cavity to efficiently control and
measure the state of the qubit.
Figure 7.5 shows a state-of-the-art circuit-QED experimental setup with the most
common ingredients. You should recognize (a) a transmission line resonator and
(b) a transmon qubit. The λ/2-resonator has an approximate length d = λ/2 ∼ 1
cm and supports an infinite number of harmonic resonances, with regularly spaced
frequencies ωn = nω1 , for n = 1,2, . . . (cf. Section 5.2.2 and Figure 5.3). The res-
onator interacts capacitively with a semi-infinite transmission line, used to read the
state of the cavity and to inject energy into it.
The transmon qubit (b) has a size of about 300 μm and sits inside the resonator,
between the core of the transmission line and one of the ground planes. The position
of the qubit is not arbitrary: It is close to the end of the cavity, because that is where
the electric potential of the resonator modes is maximal (cf. Figure 5.3), but it is
not too close so that it remains protected by the surrounding cavity ground planes.
The qubit also has some control and readout lines. Note the (c) line carrying some
current that sinks into the ground plane: The magnetic field generated by this current
affects the transmon’s SQUID inductance change and thereby changes the qubit’s
gap. Another line (d) applies a voltage into the qubit and is used to rotate the qubit
basis, as well as for readout.
Summing up, already this simple-looking circuit exhibits three control knobs and
two measurement setups. We can tune the two parameters of the qubit and ε
as requested in Section 6.1.1, we can drive and displace the cavity field using the
ideas from Section 5.5.1, and we can perform spectroscopy of the qubit and of the
cavity by sending photons and studing the reflected signals. As we explain now,
these are extremely powerful controls that can be used as much for quantum optics
experiments as to engineer scalable quantum computers.
12 See Zueco and García-Ripoll (2019) for a complete derivation of the cavity-QED model for all regimes of
losses in the cavity and the qubit.
13 Interestingly, models (7.5), (7.32), and (7.33) were all posed by physicists Jaynes and Cummings to describe
an idealized atom in a quantum electromagnetic field. However, the Jaynes–Cummings or JC Hamiltonian
nowadays refers to RWA version (7.33).
Figure 7.6 Energy-level structure of a qubit and a cavity without coupling (solid),
as described by the Jaynes–Cummings model (7.33). The interaction couples neigh-
boring states |g,n and |e,n − 1, creating new eigenstates that are superpositions
of the original ones, and whose energies split even further (dashed).
now the cavity–qubit couplings were around 2π × 220 MHz, two orders of magni-
tude larger than the cavity losses, κ = 2π ×2.4 MHz, and γ ∼ 0.8−1.2×2π MHz.
Hn≥1 = h̄ω n − 1
2
+ h̄n vn · σ̃, (7.36)
En,± = h̄ω n − 1
2
± h̄n, (7.38)
which split symmetrically around h̄ω n − 12 . We call these the lower and upper-
polariton branches, En− and En+ , because the eigenstates are hybridized states
(polaritons) of excitations living on the cavity |g,n and on the qubit |e,n − 1. The
qubit–photon coupling introduces a level repulsion between the polariton branches,
which separate by an amount proportional to the Rabi frequency 2n .
When the cavity and the qubit are straight on resonance = ω, the splitting is
√
minimal 2 n|g| and the two polariton states become symmetric superpositions:
1
|ψn± = √ (|g,n ± |e,n − 1) , n ≥ 1. (7.39)
2
√
If, on the other hand, the detuning is very large | − ω| n|g|, the coupling
strength will be insufficient to hybridize the qubit and the cavity. The two eigenstates
will be then approximately close to the bare states |g,n and |e,n − 1, and will have
close noninteracting eigenenergies En± ∼ {h̄ωn − 12 h̄, h̄ω(n − 1) + 12 h̄ω}, with
small dispersive corrections to be discussed in Section 7.3.7.
Figure 7.7 (a) Transmission through the cavity as a function of the qubit bias, ε,
for = 0.7ω and g = 0.05ω. With√the dashed line we plot the qubit energy levels
for g = 0. The linewidths assume γ κ = 0.005ω. (b) Transmission spectroscopy
for a flux qubit interacting with the third mode of a microwave resonator, in the
ultrstrastrong coupling regime g/ω = 16%. (c) Two different fits with the RWA
and Jaynes–Cummings and the full Rabi model.
of the qubit in the cavity. Fortunately, the evolution under the full JC model can be
solved analytically for any state |ψ(0) ∈ Hn , using the trick (2.22)
e,n − 1|ψ(t) 1
ψ n (t) = = e−iω(n− 2 )t cos(n t) − i sin(n t)vn · σ̃ ψ n (0).
g,n|ψ(t)
Take for instance the initial state |ψ(0) = |g,n. Our analytical solution predicts
a (partial) transfer of population from the ground to the excited qubit. The period of
the Rabi oscillations is determined by the Rabi frequency n , while the efficiency
of the transfer depends on the relative strength of the detuning | − ω|/n . Indeed,
the transfer is only perfect in the resonant case = ω,
|ψ(t) = cos(n t) |e,n − 1 − i sin(n t) |g,n , n ≥ 1, (7.40)
√
where it happens with a frequency n = g n that is half the Rabi splitting.
Actual Rabi oscillations are always damped due to decoherence in the cavity
or the qubit. To analyze this, let us project our density matrix onto the different
sectors with a fixed number of excitations ρN (t) = PN ρ(t)PN . Following (7.34),
these matrices satisfy the master equation:
i
∂t ρN = − Heff ρN − ρN Heff †
+ κ âρN+1 â † + γ σ − ρN+1 σ +, with (7.41)
h̄
h̄γ h̄κ
Heff = HJC − i σ + σ − − i â † â.
2 2
The matrix ρN is fed by excitations from higher sections ρN+1 and itself evolves
with a non-Hermitian “Hamiltonian” Heff , which is excitation number conserving,
and where the frequencies of the qubit and the cavity now contain imaginary correc-
tions. If our initial condition has N or less excitations, ρN+1 (t) = 0, we can solve
analytically the dynamics of ρN as a Schrödinger-like equation:
†
ρN (t) = e−iHeff t/h̄ ρN (0)e+iHeff /h̄ . (7.42)
We then find that the role of γ and κ in Heff is to exponentially reduce in time the
probability PN = tr(ρN ) of having N excitations (see Problem 7.9), as one would
expect from their interpretation as qubit and cavity loss rates.
Let us apply this idea to study the Rabi oscillations of an excited qubit in an
empty cavity ρ1 (0) = |e,0e,0|. The single excitation component of the density
matrix is a projector ρ1 (t) = |ψ(t)ψ(t)|, onto an unnormalized state |ψ(t) =
ce (t) |e,0 + cg (t) |g,1, which follows the equation
1 1
¯
c(t) = e−i 2 (+ω)t− 4 (κ+γ )t × cos(t)1 ¯
− i sin(t)v̄ · σ c(0), (7.43)
−ω κ −γ 2 g −ω κ −γ
¯
= g + 2 +i , v̄ = ,0, +i .
2 4 ¯
¯
2 4¯
1.0
(a) (b)
0.8
0.6
0.4
0.2
0.1 0.2
0.0 0.0
0 5 10 15 20 25 0 5 10 15 20 25
Figure 7.8 Dynamics in the weak and strong coupling regimes. (a) Excited state
population for the cavity–qubit system starting in the |e,0 state, as a function of
time and coupling strength g. (b) Horizontal cuts for g = 0.05γ ,0.1γ and 0.2γ .
Plots use ω = = 1, γ = κ = 0.1.
The norm of |ψ decreases exponentially, as does the probability of having one
excitation in the system tr(ρ1 ) = !ψ!2 ∼ exp(−(κ + γ )t/4). By virtue of (7.41),
this probability is transferred to the ground state component of the density matrix
ρ0 (t), through processes of cavity decay and the qubit’s nonradiative losses.
Our analytic solution predicts Rabi-type oscillations between the |e,0 and
|g,1 states, modulated by an exponential envelope. Figure 7.8b explores these
√
oscillations for a broad range of the cooperativity factor g/ γ κ, with select
cuts plotted in Figure 7.8a. In the weak-coupling regime, decoherence is stronger
than qubit–photon interactions g γ ,κ, and we cannot resolve the oscillations.
This is exemplified by the exponentially decaying solid line in Figure 7.8b. The
opposite regime, called strong coupling, appears when g γ ,κ. Thanks to a longer
coherence time, we can detect one or several Rabi oscillations in the dashed and
dotted lines from Figure 7.8b.
The strong-coupling regime was considered the Holy Grail of quantum optics for
a long time. One of the difficulties in reaching strong coupling was the need for a
cavity with a large quality factor14 Q = ω/κ. The first AMO experiments in the
strong-coupling regime used superconducting microwave cavities that could trap
photons for a long time. These cavities were made to interact with flying photons
in Rydberg states in a series of experiments that culminated with the measurement
of Rabi oscillations and the creation of Schrödinger cats (Haroche, 2013). In retro-
spect, circuit-QED is the natural evolution of the experiments by Nobel Prize winner
14 Q loosely represents the number of times a photon can oscillate in a cavity before it decays.
Serge Haroche. We keep the good microwave cavities, but replace the atom with an
even more strongly interacting object – the superconducting qubit.
interesting quantum phase transitions (Kurcz et al., 2014). However, we will not
have space available to discuss these applications in this book.
which can also be upgraded to a full master equation (7.34). In this model, the qubits
talk “collectively” with the resonator. The cavity exchanges excitations with all of
them and mediates interactions among the qubits.
Spectroscopically, the features of this model are very similar to the JC Hamilto-
nian. For simplicity, we will discuss the case of identical emitters, with i = and
gi = g. In this limit, the TC becomes an RWA version of the Dicke model:
1
H = Ŝ z + h̄ωâ † â + g Ŝ + â + g ∗ Ŝ − â †, (7.45)
2
N z
with the collective operators Ŝ ± = N ±
i=1 σ and Ŝ =
z
i=1 σ .
The cavity is coupled to a symmetric superposition of excitations in √ the qubit
space. This collective excitation acts as a new quasiparticle b̂† ∼ S + / N with
bosonic statistics, which interacts more strongly than the single-qubit case. In the
limit of many qubits,
√
H h̄b̂† b̂ + h̄ωâ † â + N |g|(b̂† â + H.c.). (7.46)
This physics is analogous to many other collective phenomena, from lasing to Bose–
Einstein condensation, where a bunch of identical quantum systems act coherently
in a synchronized way, and this way they become “better” at doing something – in
this case, at producing a photon.
The ground state of this model is the trivial vacuum state |0̃ := |g,g . . . ,g,0.
The single-excitation space contains N + 1 states with one excitation on any qubit
{σi+ |0̃} or on the cavity â † |0̃. These states form one upper and one lower polariton
branch, with eigenfrequencies
1
E1± = h̄ω ± ( − ω)2 + g 2 N . (7.47)
2
Note the increased √
splitting between polaritons caused by the bosonic enhancement
of the interaction Ng. In addition to the polariton states, the Hilbert space is
completed with N − 1 dark states, excitations that live in the qubits and do not
“see” the cavity. For instance, if we have N = 2 qubits, the only dark state is the
qubit singlet √12 (σ1+ − σ2+ ) |0̃.
The first term has been written in a way that evidences a shift of the resonator
frequency. The direction or sign of this shift depends on the state of the qubit σ z
and on the detuning. We will see in Section 7.4.2 how this shift can be used to
perform a nondemolition measurement of the qubit’s state. The dispersive coupling
also introduces a constant dc-Stark shift on the qubits and, when there are two or
more qubits, an exchange interaction σj+ σi− mediated by virtual transitions through
the resonator.15 These exchange terms can be used to generate entangled states,
simulate spin Hamiltonians, and implement two-qubit quantum gates, among other
applications.
Figure 7.9 Spectroscopy schemes for a qubit in a cavity. (a) Single-tone spec-
troscopy, consists in driving the cavity with a coherent state at a frequency ωd and
studying the transmitted and reflected light. (b) Two-tone spectroscopy consists in
probing the cavity transmission at a frequency ωp , while driving the qubit with a
different frequency ωd . The driving of the qubit may take place through a separate
antenna or – if enough power is available to compensate for the attenuation –
through the antenna that drives the cavity itself.
15 The mediating role of the cavity is evident when you analyze the second-order Schrieffer–Wolff correction
1 g 2 [S,V ] ∝ σ + â â † σ − which is responsible for these terms.
2 j i
sketches the two most common setups for spectroscopy of a cavity–qubit system.
The simplest method, in Figure 7.9a, drives the resonator with a monochromatic
microwave at different frequencies ωd , studying how much power is transmitted
and reflected by the cavity.
The analysis of single-tone spectroscopy mimics the spectroscopy of linear res-
onators from Section 5.5.3. We first establish a relation between the output fields
that are reflected/transmitted by the cavity, and the operator of the cavity itself. In
the transmission case, all the signal is produced by the emission from the cavity,
√
b̂out = −i κ â(t). We may compute this expectation value using our master
equation for the cavity-QED setup (7.34), with an added term (t)â † + (t)∗ â that
accounts for the probe drive (t) = 0 exp(−iωd t + iφ).
There is a simple trick that works when the strength of the drive is weak enough
that it can only excite transitions from the unique ground state – |g,0 in this case –
to a finite set of excitations that are disconnected from each other – in this setup,
the two polaritonic branches E1,± with states |1± . In this limit, we approximate
our system as a collection of harmonic oscillators, with the vacuum being the
ground state and one oscillator mode for each excited state at frequencies dictated
†
by the allowed transitions – |vac ∼ |g,0 and b̂± |vac ∼ |1± the two polaritons,
and h̄ω± = E1± − E0 . Each of the oscillators is driven by the microwave with a
†
reduced amplitude (t)β± b̂± + H.c. and experiences also a reduced decay rate
κ± = |β± | κ, where β± ∼ 1± |â † |vac is the matrix element of the cavity operator.
2
Under these conditions, the output field becomes a linear superposition of the fields
absorbed and reemitted through the different allowed transitions:
κs 1
T (ωd )0 eiφ = βs 0 eiφ . (7.50)
s=±
2 i(ω s − ω d ) + κ s /2
This model shows that a low-power spectroscopic signal will reveal two Lorentzian
resonances associated to the lower and upper polariton branches, centered on the
frequencies ω± = (E1± − E0 )/h̄. Each of these Lorentzian profiles will be very
similar to the ones in Figure 7.4, with the roles of R and T reversed if we are looking
at a transmission spectrum. The height of the resonance is the proportional to the
matrix element β± and is therefore dependent on how much of a photon component
|g,1 is present in each of the eigenstates. Conversely, the width of the Lorentzian
depends on the lifetime of these polaritons: The faster that these states decay due
to κ or γ , the broader the resonance.
In Figure 7.7, we plotted the transmission spectrum of a setup where the gap of
the qubit is tuned using the transverse field εσ x . When the qubit is far away from
the cavity resonance, only one of the two polariton branches has sufficient overlap
with the excited cavity mode |g,1. This mode allows the transmission of photons
around the cavity frequency ωd = ω, with a linewidth given by κ. When the qubit
approaches the cavity, we begin to see the vacuum Rabi splitting in action. Straight
on resonance, 2 + ε2 = ω2 , the two polariton resonances ωd = ω ± g split with a
gap 2g. In the strong coupling regime, the decay rate of the cavity and of the qubit
are smaller than this splitting, and both resonances can be resolved.
Low-power spectroscopy was also the fundamental tool in identifying the
ultrastrong coupling regime, both in the works by Niemczyk et al. (2010) and
Forn-Díaz et al. (2010). Note that even in the USC regime, the approximation of
independent resonances works, although the computation of the matrix elements
βn is complicated by the strongly hybridized and qubit–photon entanglement in
the low-energy excitations. In Figure 7.7b,c, we reproduce a transmission spectrum
from Niemczyk et al. (2010), illustrating a qubit that comes to resonance with the
third mode of a transmission line resonator. The coupling strength between the
qubit and the resonator is very large, g/ω 12%, thanks to a clever design where
the qubit is embedded in the central conductor of the resonator. As anticipated
in Section 7.3.5, the spectrum from Figure 7.7b is only fitted properly when we
consider the full Rabi model (7.32), with both rotating and counterrotating terms.
expect all the field to be reflected, possibly affected by a small delay that changes
the phase of the reflected field:
i(ω − ωd ) − κ/2
Rωd = ∼ eiφ(ωd ) . (7.52)
i(ω − ωd ) + κ/2
When we drive on resonance ωd = ω and there is no qubit, the reflected field expe-
riences a π shift and Rωd = −1. However, if there is a qubit, the reflected field on
resonance will produce a slightly different shift:
iσ z χ − κ/2 z
Rωd = ∼ ei(π+δφσ ), (7.53)
iσ z χ + κ/2
which depends on the qubit’s state σ z = ±1:
χσ z
tan(π + δφσ z ) . (7.54)
κ
Detecting this phase is a process that takes some time, during which we
integrate the quantum field reflected by the cavity, using the techniques from
Section 5.5.7. This measurement attempts to distinguish the coherent state created
by the microwave drive inside the cavity, and that depends on the state of the qubits
|α± := |αei(π±δφ) . A proper discrimination of these states requires us to strike
several balances.
– The power of our coherent microwave probe must be low enough not to create
large numbers of photons inside the cavity – which would invalidate the dis-
persive approximation – but also large enough so that the two coherent states
are distinguishable. In other words, we want a small overlap αe+iδφ |αe−iδφ ∝
e−|2αδφ| ∼ 0, while keeping |α|2 1.
2
for stronger drives, faster measurements, and reduced errors, which as of the time
of writing this book lay around 1% (2%) in 88 ns (48 ns) for state-of-the-art setups
(Walter et al., 2017).
Figure 7.10 Two-tone spectroscopy of a charge qubit. (b) shows the phase shift
experienced by the transmitted photons as the qubit energy is moved around using
an external voltage and driven with some frequency, νd . As shown in (a), when the
qubit is appropriately excited, the average phase drops significantly to zero. Figure
reproduced from Schuster et al. (2005).
half a Rabi period. This way, the qubit and the cavity exchange excitations, evolving
from |e,0 to |g,1. If at this point we rapidly bring the qubit out of resonance again,
we will have created a single-photon state.
As discussed in Section 5.4.6, this idea was demonstrated by Eichler et al. (2011)
in a seminal experiment that not only showed the generation of the photons, but also
demonstrated the nature of the quantum state by doing Wigner function tomography
of photon’s wavefunction as it left the cavity.
we can create a state |χN−1 that has one less excitation. For this, we use two
operations. The first one, Qn , consists in bringing the qubit into resonance with
the cavity = ω and wait a time tN sufficient for all population from |g,N to
be transferred to |e,N − 1. The second operation is a rotation of the qubit CN ,
transforming |e,n ↔ |g,n − 1, which converts |e,N − 1 into |g,N − 1 and
leaves us with a |χN−1 state.
Using this technique, we can compute a sequence of gates that undoes the desired
state C0 Q0 · · · CN QN |χN = |0. If we apply the reversed sequence C0†, Q†0 . . . Q†N ,
we will be able to construct the Fock superposition state experimentally, leaving the
qubit unentangled from the cavity. This technique was demostrated by Hofheinz
et al. (2009) with up to N = 5 photons, in a very interesting work that not only
created the state but used Wigner function tomography to reconstruct the complete
density matrix of the cavity.
H = χ σ z a † a. (7.60)
Assume we prepare the cavity in a coherent state |α and simultaneously excite
the qubit into a superposition √12 (|g + |e). After a time t, the combined state will
have evolved into
1 1
√ e−i(−χ)t |g |e−iχ t α + √ e−i(+χ )t |e |e+iχ t α . (7.61)
2 2
We will choose χ t = π/2, so that both coherent states have opposite signs, and
further displace the cavity along −iα, obtaining the state
1 1
√ e−i(−χ)t |g |−2iα + √ e−i(+χ )t |e |0 . (7.62)
2 2
Interestingly, we can now use the dispersive shift of the qubit’s frequency to
implement a rotation that only is effective in the vacuum state. In particular, we drive
the qubit with a long pulse, resonant with the frequency + χ , which only excites
the transition |g,0 ↔ |e,0, leaving all |g,n > 0 and |e,n > 0 approximately
unaltered. The outcome of this rotation is a new state where the qubit is decoupled
from the cavity:
1
|g ⊗ √ (|−2iα + |0) . (7.63)
2
Using further displacements, we can bring this into a more symmetric superposition
|−iα + |iα. Note that we can also change the relative phase of the coherent states
by playing with the single-qubit rotation’s phase. Moreover, due to free evolution
with the resonator’s Hamiltonian, the states ±iα rotate in phase space so that we
effectively can prepare superpositions along different directions.
Exercises
7.1 Show how (7.13) can be obtained by projecting both sides of the Schrödinger
equation i∂t |ψ(t) = Ĥ |ψ(t) onto the basis states |1,vac and âk† |0,vac.
7.2 Markovian kernel. Let’s have a look at the qubit’s dynamics:
t
i
∂t c1 = −ic1 − J (ω)e−iω(t−τ ) c1 (τ )dω dτ, (7.64)
2π 0
assuming that the spectral function is constant J (ω) = 2πα over the whole
real line, ω ∈ R. Prove that the differential equation is exactly Markovian and
only depends on the value of c1 (t). What are the decay rate γ and the Lamb
shift?
7.3 Qubit–photon entanglement. We can use (7.24) to estimate the amount of
entanglement between the qubit and the emitted photon. The von Neumann
entropy of the reduced density matrix ρqb
S = −tr(ρqb log(ρqb )) = − λi log(λi ) (7.65)
i
7.5 A chiral interaction is one that breaks time-reversal symmetry. In our context
of light–matter interactions, a qubit interacting with a chiral medium would
couple differently to photons moving along different directions g+k = g−k .
Let us assume that our waveguide is chiral because the k < 0 modes are totally
absent. Take the following steps to understand how this changes light–matter
interactions:
(1) Assuming gk ∼ gdk 1/2 , integrate the qubit dynamics and compute the rela-
tion between g and the spontaneous emission rate, γ .
(2) Write the new input–output relations: How many fields do you have to
consider? What is the new prefactor in front of c1 (t)?
(3) Repeat the derivation of the scattering coefficients for an incoming photon.
Show that Rω = 0, |Tω | = 1 and that the photon only experiences a phase
shift.
(4) Explain when the chiral model can be used to study the scattering of
photons by a qubit that is placed at the end of a semi-infinite transmis-
sion line.
7.6 Consider a multilevel circuit with N excited levels, interacting with the trans-
mission line. We assume that each level only couples to the ground state via
the propagating photons:
N
H = i |ii| + ωk âk† âk + βi σi+ gk âk + gk∗ â † . (7.66)
i=1 k k
Here σi+ = |i 0| and the βi measure the relative strength of the different
0 ↔ i transitions.
(1) Derive the evolution equations for the single-excitation ansatz:
N
|ψ = ci σi+ + ψk âk† |0,vac . (7.67)
i=1 k
(2) Integrate the photons out and apply the Markovian approximation to obtain
N equations for the ci coefficients. How does the excited state |i evolve
in time?
(3) Generalize the input–output relations to consider the N transitions.
(4) Compute the scattering amplitudes Tk and Rk for a single incoming pho-
ton. What does the spectrum look like?
7.7 Lambda system and dark states. Let us study the following energy-level
structure, with two degenerate ground states, |a and |b, and one excited state:
What can you say about the spontaneous emission from state |1 and the inter-
action with propagating photons? Hint: Show that this problem is equivalent
to a two-level system plus one noninteracting dark state for any (complex)
values of the two coupling strengths, ga,bk .
7.8 Photodetection. Let us now devise a simple photodetector using the a few-
level system that can switch the state through the absorption of a photon.
A two-level system is not a valid photodetector because even if the photon
is absorbed, the state |1 will subsequently decay. Instead, we introduce the
following three level system:
In this scheme, an incoming photon couples to the qubit and excites the tran-
sition 0 ↔ 1. At the same time, there is an incoherent transition 1 → b that
depletes the excited state population with rate and deposits it irreversibly
onto the state |b, thus signaling the presence of a photon. We are going to
model this system with the same equations as for the two-level system. The
only change is the addition of an ad hoc incoherent channel:
i∂t c1 = ( − iγ )c1 − ic1 . (7.68)
(1) Assuming an adiabatically switched on wavepacket, compute the proba-
bility of landing onto the state |b
Pb = 1 − |c1 |2, (7.69)
by integrating the dynamics of the qubit under this driving. Show that it is
always below 50%.
(2) What are the transmission and reflection amplitudes of the photon, Tω and
Rω ? How does the photodetection process modify the probabilities |T |2
and |R|2 ?
7.9 Given (7.42), prove that the probability of having N excitations, given by
P (N,t) = tr(ρN (t)) decays exponentially. More precisely, prove the equation
d
dt
P (N) ≤ −(κ(N − 1) + γ )P (N ).
7.10 Prove the expressions for the dispersive measurement (7.57) in the ideal case
where decoherence can be ignored.
198
Around the same years, we also find celebrated physicist Richard Feynman
looking at the relation between computers and quantum mechanics from a different
angle. Coming from a theoretical physics background, and having worked on
the application of computers to solving theoretical physics problems, he chal-
lenged the capabilities of classical computers (both real and abstract, such as the
Turing machine) to simulate the evolution of quantum systems. In a well-known
talk, Feynman (1982), he illustrated the exponential complexity of simulating the
Hilbert space of even very simple quantum models – particles jumping on a lattice –
and introduced the idea of quantum simulation, engineering a quantum physical
device to reproduce the dynamics of a complex quantum model that we do not
know how to solve.
Benioff’s work and Feynman’s ideas converge in the work by Deutsch (1985),
a seminal paper that introduced the concept of a quantum Turing machine. This
is a device that not only can process classical information – and therefore repro-
duce a classical Turing machine – but also information stored in generic quantum
states, including quantum superpositions and entangled states. Deutsch’s quantum
Turing machine provided the theoretical foundation for what we now understand as
a universal quantum computer.
Despite its theoretical value, both the classical and the quantum Turing machines
are inconvenient designs that do not inform the real architecture of computers.
Instead, quantum computers are developed in the lab using more scalable and highly
parallelizable frameworks. In this chapter, we focus on the particular framework of
quantum circuits introduced also by Deutsch (1989). The quantum circuit model
assumes a quantum register formed by near-ideal qubits. The register acts both as a
memory and as a processor.1 Quantum computations are implemented by modifying
this register through a composition of qubit measurements with unitary operations
that are extracted from a universal set.
The quantum circuit model includes also a very elegant graphical representation.
The algorithm’s operations and measurements are placed on a set of horizontal lines,
similar to a musical “pentagram.” Each horizontal line addresses one qubit. A box
in this pentagram is a unitary operation acting on the qubits that it is connected to.
The complete circuit can be read as a set of gates and measurements to be executed,
either in parallel or sequentially, as we read the diagram from left to right.
As an example, Figure 8.1a shows a quantum algorithm to recreate and mea-
sure a Schrödinger cat or Greenberger–Horne–Zeilinger (GHZ) state √12 |00000 +
√1
2
|1111. In this example, we begin with a set of qubits initialized to a zero state
|00000. The first operation is a single-qubit gate, a Hadamard gate H (cf. Table 8.1)
1 In some sense, one could argue that such a design is closer to an abacus than to a Turing machine (a head
moving over an infinite, 1D stream of data) or a von Neumann computer (with CPU, control unit, memory,
external storage, and I/O).
Figure 8.1 (a) Quantum circuits are schematics of quantum algorithms and proto-
cols that combine (i) qubits, represented by circles with horizontal lines; (ii) single-
qubit operations that intersect only one such lines; (iii) multiqubit operations;
and (iv) measurements. (b) Single-qubit operations, as denoted in Table 8.1.
(c) Two-qubit conditional operations for a CZ and CNOT, respectively. Quantum
measurements for feedback operations (d) can be moved to the end of an algorithm,
using two-qubit gates (e).
that sets the first qubit into a superposition state, √12 (|0+|1) ⊗ |0000. Subsequent
two-qubit operations, called CNOT’s, copy the state of the first qubit to all other
qubits, creating a large entangled GHZ state.
The power of the quantum circuit model does not lay in this nice, prescriptive,
and visually appealing representation. The circuit model gives us a minimal set of
logical ingredients that, combined together, can reproduce any classical or quantum
computation we may think of. This includes Shor’s celebrated algorithm for fac-
toring, the phase and amplitude estimation algorithm, the Harrow–Hassidim–Lloyd
(HHL) algorithm for inverting matrices, and many modern applications to real-life
problems in chemistry, finance, logistics, etc.
DiVincenzo (2000) formalized the requirements of the quantum circuit model in a
celebrated paper. The DiVincenzo criteria are a set of five requirements for scalable
quantum computation plus two for quantum communication. Cited verbatim, these
elements are the following:
(1) A scalable physical system with well-characterized qubits
(2) The ability to initialize the state of the qubits to a simple fiducial state, such
as |000 . . .
Table 8.1. Most popular single-qubit and two-qubit unitaries, with their
names and their generators – i.e., Hermitian operator O such that
U = exp(iO + iφ) with some irrelevant global phase φ.
Hadamard 1 1 Pauli-X 0 1
π y H = √1 π x X=
2σ 2 1 −1 2σ 1 0
Pauli-Y 0 −i Pauli-Z 1 0
Y = π z Z=
πσy i 0 2σ 0 −1
Phase 1 0 “π /8” 1 0
π z S= π z T =
4σ 0 i 8σ 0 eiπ /4
⎛ ⎞ ⎛ ⎞
Control-NOT 1 0 0 0 Control-Z 1 0 0 0
⎜0 1 0 0⎟ ⎜0 1 0 0 ⎟
(CNOT)
z
⎝0 0 0 1⎠
(CZ)
z z
⎝0 0 1 0 ⎠
2 (σ1 + 1)σ2 4 (σ1 −1)(σ2 −1)
π x π
0 0 1 0 0 0 0 −1
⎛ ⎞ ⎛ ⎞
1 0 0 0 1 0 0 0
√
iSWAP ⎜0 0 i 0⎟ SWAP ⎜0 1+i 1−i
0⎟
⎝0 i ⎜ 2 2 ⎟
+ −
4 (σ1 σ2 + H.c.)
3π 0 0⎠ π
8 σ1 · σ2 ⎝
0 1−i 1+i
0
⎠
0 0 0 1 2 2
0 0 0 1
(3) Long relevant decoherence times, much longer than the gate operation time
(4) A “universal” set of quantum gates
(5) A qubit-specific measurement capability
(6) The ability to interconvert stationary and flying qubits
(7) The ability to faithfully transmit flying qubits between specified locations
These criteria provide an “engineering checklist” that applies to all platforms –
trapped ions, superconducting circuits, quantum dots, etc. If all first five criteria are
met, and if the decoherence times lay above a certain threshold, we can provably
build arbitrarily large quantum computers, capable of any quantum computation
we can think of, using error correction and fault-tolerant principles to make those
computations with arbitrary precision.
In this chapter, we will go through the elements in DiVincenzo’s checklist. We
will use the tools and knowledge gathered in Chapters 6 and 7, and complete them
with additional techniques for implementing gates, architectural decisions, char-
acterizing coherence and fidelity, etc. The roadmap is as follows. In Section 8.2,
we introduce the most common forms of quantum registers, how they are mea-
sured and reset, thus covering checklist items (1), (2) and (5). Section 8.3 dis-
cusses item (4) and how arbitrary computations can be approximated using a small
set of quantum operations, a universal gate set. We show that those gates include
8.2.1 Measurements
The election of the transmon qubit naturally imposes one type of measurement
scheme: We must adopt the cavity-mediated, QND measurement that was presented
(a)
(b) (c)
Figure 8.2 (a) Superconducting qubit quantum register built from nine X-mon
qubits, a variant of the transmon qubit with a cross shape. The qubits are close
to each other for nearest-neighbor interactions; they are connected to individual
cavities (wiggly lines on the top) for dispersive qubit readout, and they are tuned by
small flux lines coming from the bottom of the picture. Modified from Barends et al.
(2014) with permissions. (b) Alternative qubit topology, resembling the work by
Arute et al. (2019), with similar qubits, but higher connectivity, and using auxiliary
qubits to mediate interactions. (c) Three-dimensional sandwich packing of two
chips, joined by superconducting indium bonds. One chip contains resonators and
readout channels structures, while the other one packs qubits.
in Section 7.4.2. As shown in Figure 8.2a, each qubit is attached to a readout res-
onator, which is connected to the outer world. The microwave cavity acts as a
narrowband filter that isolates the qubit from the readout cables. However, when
we send a resonant microwave pulse through those cables, the cavity reflects the
light with a phase that carries information about the state of the qubit. This is a
good, nondestructive POVM measurement of the qubit, with a small uncertainty of
a few percent, ∼1–3% in current experiments.
If all the resonators that talk to the qubits have slightly different frequencies,
we can multiplex the readout phase. Instead of sending a monochromatic pulse
that is resonant with one cavity, we send a broadband signal with the frequencies
of all resonators. Analyzing the reflected signal allows us to gather information
about the state of all qubits that were addressed by the probe signal. The readout of
this broadband signal requires the use of sophisticated, broadband, and quantum-
limited amplifiers, such as a TWPA (see Section 5.5.6), that can boost the signal
of all cavities at the same time. As shown by Arute et al. (2019), such multiplexed
measurement schemes can scale up to 53 qubits with significant effort. Interestingly,
the same idea that is used to implement local addressing in the measurements can
be used to scale up and parallelize local operations on the qubits.
resonators, and control lines are fabricated onto the chip. This layout will inform the
topology of the circuit, establishing which qubits can interact with others, directly
or through mediating circuits.
Take for instance the one-dimensional layout from Figure 8.2a. Qubits are placed
on a row and are in close proximity to each other. Each qubit has two nearest
neighbors, which are capacitively coupled to it. Resonators are moved away to
a separate row and are collectively addressed, and there is a third row of direct
controls on the qubits. This arrangement is clean, provides direct access to all qubits,
and minimizes the number of qubits that see each other, which is good to lower
cross-talk.
Unfortunately, the one-dimensional design does not scale well in applications.
Quantum algorithms require interactions between arbitrary pairs of qubits. These
interactions can be decomposed into other gates, but the decomposition becomes
very costly in one-dimensional topologies. For instance, we can make a quantum
operation between qubits 1 and 9 only by inserting gates between 1 and 2, 2 and 3,
and so on. If we follow this approach, it becomes clear that we are constrained by
the taxi-driver metric2 of how far two arbitrary qubits lie in a given lattice.
There are also other considerations. For instance, quantum error correcting archi-
tectures have been designed with certain lattice topologies in mind. Figure 8.2b
illustrates an arrangement of physical qubits on a square lattice, where each qubit
has four neighbors. This topology is required for error correcting algorithms based
on the surface code, as we explain in Section 8.5.4. However, moving from a simple
planar architecture (cf. Figure 8.2a) to a dense lattice (cf. Figure 8.2b) raises the
question of where to place all other components – resonators, control lines – that do
not fit in the space between qubits.
The solution to this problem comes from an early work by O’Brien et al. (2017),
which looked at how to shield qubits and protect them from decoherence. This
work introduced a sandwich architecture, where two chips are bonded together
using tiny indium contacts. One chip contains the qubits, the other one acts as a
shield against electromagnetic fields, providing better coherence. Since Rigetti’s
work, there has been steady progress in using the same sandwich architecture, but
transferring resonators and control lines to the top chip.
The idea is sketched in Figure 8.2c, using IBM’s patent as inspiration (Brink
et al., 2019). The qubit is fabricated on one slab, together with ground planes and
control lines – which also need some 3D fabrication, because they would not fit
among qubits. Resonators and readout lines are constructed on a different chip. Both
chips are attached mechanically to each other, and contacted through a small “ball”
2 The taxi-driver metric evaluates the distance between two nodes in a lattice by counting the smallest number of
edges that make a path connecting both nodes.
of indium. This superconducting ball deforms and sticks to the aluminum, estab-
lishing an electrical bridge. In this sketch, the indium bond assists in creating a
common ground plane for both chips, but it is thick enough that keeps both layers
well separated – creating what is known as an air gap.
The two eigenstates, |0,0 and |1,1, are preserved throughout evolution, accu-
mulating the same local phases they would acquire in absence of interaction. When
1 = 2 , the one states |0,1 and |1,0 are resonant and can exchange their only
excitation:
⎛ ⎞
1 0 0 0
z ⎜0 cos(J t) −i sin(J t) 0⎟
U (t) = e−iHeff t = e−it(σ1 +σ2 ) ⎜ ⎟
z
Evolution with this Hamiltonian for a brief period of time will result in the
joint action of three commuting operations: local rotations with the phases t, an
exchange of probability between the |0,1 and |1,0, and an interaction phase in
the single-excitation subspace, generated by σ1z σ2z . This combination of operations
is particularly useful to simulate fermionic models in the quantum computer, such
as in quantum chemistry applications. A particular instance of this evolution is
J t = 3π/2, which produces a gate equivalent to the iSWAP gate, up to local
transformations:
z z
U (π/2J ) = eiπ (σ1 +σ2 )/2J UiSWAP . (8.3)
The iSWAP is a universal operation – two iSWAP gates can build a CNOT gate
(Williams, 2011) – although it is less used in practical applications.
These exchange-type gates appear in all setups that allow for off-diagonal dipolar
interactions. They can be implemented in quantum computers with directly coupled
transmons – as in Figure 8.2 – but they also appear in devices with circuit-mediated
interactions. The work by Majer et al. (2007) used the dispersive coupling between
qubits in a superconducting cavity from Section 7.3.7 to implement the whole family
of rotations enabled by the exchange term, including the iSWAP. More recently, we
find these gates in the Google Sycamore quantum processor, which relies on detuned
transmons to mediate the interaction between qubits in the quantum register.
Their setup can recreate various dipolar interactions, including the exchange term
from (8.1), which they used to implement iSWAP two-qubit gates, as a foundation
for their 53-qubit quantum supremacy experiment (Arute et al., 2019).
The CZ may be generated by evolution with an interaction of the form σ1z σ2z . Unfor-
tunately, the coupling terms that we have found in Section 6.5 are not diagonal in
the qubit basis and cannot be directly used for this purpose.
A common way to implement phase gates is to do an adiabatic passage or adia-
batic deformation of the Hamiltonian that enhances the strength of interactions for
some period of time. Our system is built with a control parameter ω, which tunes the
Hamiltonian H (ω). We assume that the control begins and ends at the same point,
ω(0) = ω(T ) = ω0 , at which H (ω0 ) approximates to two noninteracting qubits
with a well-defined basis |s1,s2 . We design the passage so that around the middle
point ω(T /2), the interaction between qubits is dominant.
If the states of the two qubits are never degenerate when we change ω and if we
modify the control ω(t) slowly (adiabatically) enough, every two-qubit eigenstate
|s1,s2 of H (ω0 ) will be mapped to the instantaneous eigenstates |nω of the Hamil-
tonian and return back to their original configuration. Those states only experience
some dynamical and geometric phases |s1,s2 eiθ(s1,s2 ) , which can be derived from
the instantaneous eigenenergies they experienced during the passage:
T
1
θ (s1,s2 ) = − dtE(ns1,s2 ,ω(t)). (8.6)
h̄ 0
This idea was first used by DiCarlo et al. (2009) to implement a CZ gate for
transmon qubits in a cavity. It was later perfected by Barends et al. (2014) for
transmons that interacted directly with each other, in the design from Figure 8.2. The
CZ adiabatic gate requires two qubits with slightly different frequencies, 2 > 1 ,
and our control is the frequency of the second qubit 2 = ω(t). The transmons have
an always-on exchange interaction (cf. Section 6.5), either direct or mediated by the
cavity. This interaction is weaker than the anharmonicities α1,2 of both transmons
along the passage.
As explained in García-Ripoll et al. (2020), we can analyze the gate in the reduced
space {|0,0 , |1,0 , |0,1 , |1,1 , |0,2} of the qubit basis plus a state |0,2 with two
excitations in the second transmon. The effective Hamiltonian in this subspace reads
⎛ ⎞
0 0 0 0 0
⎜0 h̄ ⎟
⎜ 1 J 0 0 ⎟
⎜ ⎟
Heff (t) = ⎜ 0 J h̄ 2 0 √0 ⎟. (8.7)
⎜ ⎟
⎝0 0 0 h̄1√+ h̄2 2J ⎠
0 0 0 2J 2h̄2 − h̄α2
Figure 8.3 Phase gate with two transmons with different energy gaps. (a) We tune
the frequency of the second transmon ω2 (t), from a largely detuned value down to
the resonance ω2 = ω1 + α2 between the |0,2 and |1,1 states. (b) Eigenenergies
of the tuned Hamiltonian, for levels that begin and end at the states |1,1 (solid)
and |0,2 (dashed). The |1,1 state is pushed downward from the eigenenergies of
the noninteracting model (J = 0, dotted) by an amount δEint (t).
J , and the qubit basis is almost an eigenstate of Heff . The gate proceeds by low-
ering down 2 close to the point 2 = 1 + α2 and back to its original value (cf.
Figure 8.3a). If the tuning of the frequency is adiabatic enough and |J | 2 − 1 ,
states |0,0 , |1,0 , |0,1 remain approximately invariant throughout the gate.
The qubit register |1,1, on the other hand, experiences an interaction with
the |0,2 state that cannot be neglected. Throughout the evolution, |1,1 is there-
fore mapped to an adiabatically deformed eigenstate, with a trajectory shown in
Figure 8.3b as a solid line. Halfway through the passage, it reaches the super-
position state √12 (|1,1 − |0,2) around the degeneracy point 2 = 1 + α2 . This
deformation also changes the instantaneous eigenenergy, which √ is shifted downward
h̄1 + h̄2 − δEint (t), with a maximum value δEint ∼ 2 2J right at the avoided
level crossing.
The instantaneous eigenenergies of the various adiabatic passages manifest in
different dynamical phases for each of the two-qubit states. The three lowest energy
states acquire phases that are noninteracting in nature:
T
θ(s1,s2 ) = − (1 s1 + 2 (t)s2 )dt. (8.8)
0
These are phases similar to the ones generated by the single-qubit Hamiltonian
1
2
h̄1 σ1z + 12 h̄2 (t)σ2z . The |1,1 state also acquires this local phase rotation, but
it also experiences the energy shift caused by the interaction:
i t
θ(1,1) = θ (0,1) + θ(1,0) − δEint (t)dt. (8.9)
h̄ 0
By adjusting the detuning 2 (t) (total duration, waiting time around degeneracy,
ramp profile, etc.), one may engineer a very good CZ gate that satisfies (8.5). In
addition to this, it is also possible to tune the rotation angle to produce other control-
phase gates, which may be useful in algorithms such as the quantum Fourier trans-
form (see Problem 8.4).
As shown experimentally (Barends et al., 2014) and theoretically (García-Ripoll
et al., 2020), this gate can be very fast. It does not need to be implemented adia-
batically. The only requirements for the gate to succeed are (i) minimizing leakage
and (ii) ensuring a large interaction phase. For the first condition, we need to ramp
down the qubit slow enough as to prevent transitions outside the two-excitation
subspace. Fortunately, this is limited by the anharmonicity, and happens already at
the speed of local gates (T 1/|α|). The second condition only involves tuning
the waiting time, so that the integral δEint becomes large enough. Once more, √this
also happens in a relatively short time, of the order of the energy splitting 2 2J ,
with gates as short as tens of nanoseconds.
1 1
|± = √ |1,1 − √ |0,2 . (8.10)
2 2
The state |1,1 suffers a different, but still quasi-adiabatic dynamics, that depends
on the energy splitting between these eigenstates 2 × δEint (t):
1
U (t) |1,1 = √ e−i(ω1 +ω2 )t × e−iθint (t) |1,1 − e+iθint |0,2 , (8.11)
2
t
θint (t) = δEint (τ )dτ .
0
– State preparation. These are the errors we make when initializing the quantum
register. As mentioned earlier, thermalization by itself is not fast enough, and
we need to couple our qubits to other items (cavities, controls, measurement
devices) and apply complex protocols to ensure a fast reset. State preparation
errors account for deviations from the ideal zero state |0⊗N , which, as we saw
earlier, can be around 1–2%.
– Measurement errors. The precision of measurements in quantum computers is
limited. Limited integration time, measurement induced decoherence, and ampli-
fier and detector noise are some circumstances that require us to model measure-
ments as POVMs with intrinsic uncertainties. The measurement fidelity is the
probability that, given a certain outcome of our device, the measured qubit was
in that particular state. Circuit-QEDs’ fast dispersive qubit measurements seem
to have relatively good fidelities, with around ∼ 99–99.8% fidelity or success
probability (see Barends et al., 2014; Jeffrey et al., 2014; Walter et al., 2017). This
is good, but not yet so good that we can use unlimited repeated measurements
on the same qubit.
– Qubit decoherence. We have studied the loss of coherence and quantum
superpositions due to environment-induced relaxation, heating, and dephasing.
The models from Sections 6.1.6 and 6.1.5 are good descriptions when these
process are uncorrelated – i.e., when qubits talk to separate environments in
Markovian processes that degrade the purity of the states over time. Those
models are simply characterized by the relaxation and dephasing times, T1 and
T2 , for each qubit. However, sometimes this characterization is insufficient,
because decoherence is correlated – e.g., fluctuations of a surrounding magnetic
field or control cables affect various qubits similarly – or because there is
residual cross-talk, unwanted interactions between the qubits underlying this
loss of information.
– Gate imperfections. Single-qubit and two-qubit unitaries are implemented
using microwave drives in combination with magnetic fields that tune the
frequencies and couplings of qubits. The operation of these gates is affected
by intrinsic errors as well as by our limited capacity on stabilizing and crafting
the controls. A π -pulse flip of the qubit will have errors if we fail to give the
microwave pulse the right amplitude, phase, or duration, and also if the frequency
of the qubit fluctuates slightly due to environment-induced dephasing. The
quality of our gates is measured in terms of a gate fidelity, which is the probability
that we implement the right operation or, alternatively, the mathematical overlap
between the experimental and the desired outcome.
system of N qubits, the most general map is described by a matrix with about
d 4 = 24N parameters. To see that, let us examine the Bloch sphere representation
of a density matrix, introducing a complete set of observables formed by all tensor
products of Pauli matrices4 Sα ∈ {1,σ x ,σ y ,σ z }⊗N . The 22N operators form a com-
plete orthogonal basis with respect to the matrix scalar product tr(Sα† Sβ ) = dδα,β .
This allows us to expand any density matrix:
d2
1 1 †
ρ =s ·S=
T
sα Sα, with sα = tr Sα ρ = Sα ρ . (8.13)
α=1
d d
A general completely positive map ε(ρ) is a linear transformation of the generalized
Bloch vector, represented by a matrix M ∈ R4 ×4 with about 16N independent
N N
coefficients:
1
d 2
ε(ρ) = (Ms) · S = T
Mα,β tr(Sα† ρ)Sβ . (8.14)
d α,β=1
There are other alternative representations for the completely positive maps in an
experiment. The χ matrix (Gilchrist et al., 2005; O’Brien et al., 2004) expresses the
superoperator in a complete basis Sα :
d 2
For physical maps, the matrix χ is positive semidefinite and can be diagonalized
χ = V V † , producing one of many equivalent definitions for the Kraus operators
A = Vα Sα (Neeley et al., 2008). This representation is linearly related to the
Pauli transformation matrix M:
d 2
1
Mαβ = tr(Sα Sγ Sβ Sδ )χγ δ, (8.16)
γ ,δ=1
d
4 For two qubits, for instance, this would be formed by all operators in the set of 16 combinations: {1, σ x , σ y ,
1 1
y y y y y y y
σ1z , σ2x , σ2 , σ2z , σ1x σ2x , σ1x σ2 ,σ1x σ2z , σ1 σ2x , σ1 σ2 , σ1 σ2z , σ1z σ2x , σ1z σ2 , σ1z σ2z }. Note how these operators are
both Hermitian and unitary.
5 For instance, a qubit has S ∈ {1,σ x ,σ y ,σ z }, and we could use the overcomplete set of pure states
ρk ∈ { 21 (1 ± σ z ), 12 (1 ± σ x ), 12 (1 ± σ y )} as reconstruction basis (Chow et al., 2012).
their ideal targets. The natural measure for quantum states is the fidelity, a number
between 0 and 1 determining the probability that we prepared the desired quantum
state or reference. Given any two states, described by density matrices ρ and σ , their
mutual fidelity is
2
F (ρ,σ ) = tr ρ 1/2 σρ 1/2 ∈ [0,1]. (8.19)
Note how this expression simplifies to the overlap squared for pure states
F (|ψ , |φ) = |ψ|φ|2 . It is clear from this particular case that a perfect fidelity
F = 1, read as “100% fidelity,” is only achieved for identical states, up to global
phases. The infidelity = 1 − F (ρ,σ ) measures the error in approximating σ by ρ
and is related to their separation.6
The notion of fidelity can be extended to work with quantum processes in differ-
ent ways. For instance, say that we wish to compare an ideal unitary operation U
to the actual process that takes place in the lab ε(ρ). Formally, we could apply ε(ρ)
to different pure states |ψ and estimate the average probability that we obtain the
right answer – i.e., the state fidelity between U |ψ and ε(|ψψ|). This leads to the
average gate fidelity:
Favg (ε,U ) = ψ|U † ε(|ψψ|)U |ψ dψ. (8.20)
This integral is computed by averaging over all possible pure states, sampled with
a uniform (Haar) distribution.
The average gate fidelity works with general positive maps, but as argued by
Gilchrist et al. (2005), it lacks stability under composition with other trivial opera-
tions.7 The solution to this problem is given by the process fidelity or entanglement
fidelity, a quantifier based on the Choi–Jamiolkowski isomorphism between pro-
cesses and states (8.17).
We introduce the process fidelity of two completely positive maps, using their
Choi matrices as proxies for the comparison:8
1
Fproc (ε0,ε1 ) := tr ρε†0 ρε1 = tr M0† M1 = tr χ0† χ1 . (8.21)
d2
6 For normalized pure states, !ψ − φ!2 = 2(1 − F 1/2 ) for small errors. Note that traditional definitions in
the literature chose the geometrical convenience (see Nielsen and Chuang (2011)) F 1/2 = |ψ|φ| as fidelity
for pure states. You find this convention in early works in quantum technologies (Steffen et al., 2006), but it
has since been abandoned.
7 Imagine that we put together our system and an uncoupled experiment where we do nothing. We would expect
that the auxiliary system should not modify the average of the fidelity, Favg (1 ⊗ ε0,1 ⊗ ε) = Favg (ε0,ε), but
this is not true in general!
8 Originally, the process fidelity was introduced as F †
proc (ε,U ) = |U ε(||)U |, to compare positive
maps ε with ideal unitary operations (Gilchrist et al., 2005). This formula measures the fidelity of the Choi
matrix for the positive map, with respect to the pure state associated to U .
Interestingly, the average gate fidelity and the process fidelity are related:
dFproc + 1 tr(M0† M1 ) + d
Favg = = . (8.22)
d +1 d2 + d
If an experiment is designed for full process tomography, we can recover both Fproc
and Favg . This idea has been applied to the study of a qubit under no operations (Nee-
ley et al., 2008) to characterize single-qubit operations (Chow et al., 2009) and to
study universal two-qubit gates (Bialczak et al., 2010). Unfortunately, it is difficult
to scale up this quantifier to larger numbers of qubits due to exponential growth in
experimental setups required for complete process tomography ∼ d 4 ∼ 24N . This
limitation has motivated the quest for simpler experimental protocols that exploit
the power of randomized operations to estimate the fidelities of state preparation,
measurements, and single-qubit and two-qubit gates.
,
1
ρf = gl+1 Sl+1 Ggl Sl |0 . (8.25)
i=l
The operation G may be as simple as idling or waiting for a time t after each
Clifford gate. When doing so, we are calibrating the coherence of our qubits, either
individually – when the experiment is done on each qubit separately – or as a whole –
when we consider the Clifford group on more than one qubit. O’Malley et al. (2015)
has taken this idea a step further, developing an error metrology protocol that com-
bines randomized benchmarking with Ramsey-type pulses. The result is a very
sensitive analysis toolbox to explore the decoherence and dephasing properties of
the qubits, including a careful study of the phase fluctuations that bound deviations
from the exponential decay model.
Randomized benchmarking was introduced by Knill et al. (2008) and later gen-
eralized to multiple qubits by Magesan et al. (2011) and Gaebler et al. (2012). It
exploits important connections between the performance of quantum computers in
large, arbitrary calculations and the quality of the ingredients we built into the com-
puter. In particular, it demonstrates in an intuitively appealing way the exponential
accumulation of tiny errors in the quantum processors. For instance, a failure rate
as small as 0.1% error per gate transforms into a success probability of P20 = 80%
after 20 gates, P40 = 66% after 40, and P160 = 20% after 160, and so on. Considering
that useful quantum algorithms must be much longer to compete against classical
methods, it becomes obvious that we need to find ways to cope with this rapid
deterioration of the quantum register, incorporating some type of error correction.
One may argue that randomized benchmarking provides a standard, low-cost,9
and reproducible strategy to map out the quality of a quantum computer’s compo-
nents. In particular, RBM is a very generic algorithm that applies to all quantum
computers based on the circuit model. It therefore enables a systematic comparison
of setups and platforms, such as superconducting circuits versus trapped ions –
Magesan et al. (2012) versus Gaebler et al. (2012) – which is not very different
from how we compare the quality of high-end versus low-end electronics – e.g.,
using global figures such as the failure rates of their hard disks, memories, and
processors. However, keep in mind that by doing so we might be dropping relevant
information about the nature and correlations among errors, which is crucial for the
development of error correction and fault-tolerant computation strategies.
this scenario, we can use a simple error-correction strategy, which is to copy the
information redundantly, using N physical bits to encode one logical bit, our unit
of information. Because the setup is classical, we may inspect the states of these
logical bits periodically, to see whether one or more bits have changed their state
and fix them.
In this classical scenario and encoding, a simple error-correction strategy is to
implement a majority vote, which copies to all physical bits the state that is occupied
by at least N/2 of them. This majority vote strategy works, because the probability
that n out of N bits have been corrupted decreases exponentially with n. If we sum
the probability that n > N/2 bits have been corrupted – in which case the majority
vote would be wrong! – we obtain the probability that our error-correction fails. This
probability scales with the redundancy as O(N N εN/2 ). The error – correction strat-
egy therefore works provided ε 1/N, which we may achieve if we interrogate
the physical qubits with fast enough frequency.
Unfortunately, this error-correction scheme does not immediately work for quan-
tum information. The no cloning theorem (Dieks, 1982; Wootters and Zurek, 1982)
implies that we cannot copy an arbitrary quantum state of one qubit onto other
qubits, going from |φ |0N−1 to |φN . But, more fundamentally, there is no way
that we can measure a qubit without destroying the quantum superpositions that
might be stored in them. We need other encodings and other strategies that are well
adapted to the quantum world.
The good thing about stabilizer formalisms is that they can be implemented using
the tools from the circuit model. Both the measurements Pk and the corrections U
are small quantum algorithms that can be compiled into elementary single- and two-
qubit gates and single-qubit measurements, possibly with some auxiliary qubits to
assist the syndrome measurements.
Most importantly, the types of stabilizer codes that we care about also allow the
implementation of fault-tolerant computations. These are unitary gates that act on
the logical qubit space without perturbing the stabilizer space or the syndrome.
Such gates are called “fault tolerant” because they work even if we use imperfect
components to reproduce them – i.e., single qubit gates with some residual errors,
imperfect measurements, etc. When we implement the fault-tolerant gate, errors
manifest as perturbations of the syndrome that can be corrected, producing the ideal
operation.
Let us emphasize that we can only implement fault-tolerant gates when the syn-
drome does not show too many changes in between correction steps. To put limits on
this, we set the fault-tolerance threshold, a bound in the infidelity of any operation
from our quantum toolbox – including state preparation, measurement, and gates,10
all of which can be characterized via RBM. In theory, when a platform remains
within the specified thresholds, we can create fault-tolerant computers working
with quantum registers and quantum algorithms of arbitrary size. In practice, this is
largely dependent on assumptions such as noise models, code sizes, and the specific
correction protocols.
10 Including periods in which a qubit is idle, which are treated as “identity” operations.
Figure 8.4 (a) Surface code layout corresponding to the qubit arrangement from
Figure 8.2b. We show a plaquette and a vertex stabilizer operator, and one of the
ancillary qubits. (b) Quantum circuit associated to the measurement of a σ1z σ2z σ3z σ4z
plaquette stabilizer, using the central auxiliary qubit for the measurement.
– They can encode any number of logical qubits, provided the lattice sizes are large
enough.
– The codes can be implemented in a formally simple way, using the single- and
two-qubit operations from the circuit model.
We will now work out the ingredients of the surface code formalism, using a simple
lattice as an example. For further reference, we recommend the review articles by
Fowler et al. (2009, 2012).
Stabilizer Operators
Figure 8.4a shows a toy realization of a surface code, where the qubits match the
numbered transmons in the quantum register from Figure 8.2b. The surface code
operates on a square lattice, but the qubits sit on the edges, not on the vertices of
the lattice. A whole plaquette contains four edges and four qubits; we also accept
open plaquettes formed by just three edges. The example in Figure 8.4a has four
complete plaquetes and 2 + 2 incomplete ones, distributed along the two rough
edges on the sides.
There are two sets of stabilizers on this lattice. One set is formed by products of
z
σ operators on the qubits of each plaquette; the other set is constructed by products
of four σ x operators from the qubits surrounding a vertex. Figure 8.4a illustrates two
stabilizers: Z2 = σ2z σ5z σ6z σ9z and X6 = σ7x σ10
x x x
σ11 σ14 , assuming we order qubits from
left to right, top to bottom, as in Figure 8.2b.
y
Bloch vector. These operators implement the usual pseudospin algebra [jx ,k ] =
y
2i z δj k , [j ,kz ] = 2i x δj k , and [jz,kx ] = 2i x δj k . They also commute with
all stabilizers [ x,y,z,Zn ] = [ x,y,z,Xn ] = 0, because the stabilizers operate in the
independent degrees of freedom |α.
Logical qubits in the surface code are created by the boundaries and the holes
in the lattice. Each defect decreases the number of commuting stabilizers, leaving
undetermined degrees of freedom to be used for storing quantum information. The
logical qubit operators for these degrees of freedom are constructed from products
of Pauli matrices of physical qubits across the lattice, along the edges that connect
defects (holes) and borders. One may create an arbitrary number of qubits in large
lattices, provided they accommodate enough independent rough edges and holes
(Fowler et al., 2009, 2012).
The example from Figure 8.4a hosts 18 qubits and 17 stabilizers – eight stabilizers
of type Z and nine of type X. Since the stabilizers can take two values {+1,−1} and
are linearly independent, they jointly determine that the dimension of the subspace
|α is 217 . The remaining degrees of freedom can be grouped in a logical Hilbert
space of dimension 218−17 = 2, one qubit, described by three Pauli-like operators
{ x,y,z }. The diagonal Pauli operator for the logical qubit is any product of σ z along
a string that connects both rough edges – e.g., z = σ8z σ9z σ10 z z
σ11 . The real off-
diagonal operator string operator is given by = σ2 σ9 σ16 , or any topologically
x x x x
equivalent string that crosses plaquettes from top to bottom. With these two trans-
formations, we also have y = −i z x , totaling three logical gates x,y,z , that we
can implement by manipulating the physical qubits.
X5 = σ6x σ9x σ10 σ13 , which will also change sign, σ9z † X4,5 σ9z = −X4,5 . And finally,
x x
Pauli Y errors on this qubit manifest as joint phase and spin-flip errors, which affect
all four stabilizers that overlap on the affected qubit Z2,6 and X4,5 .
We build our our error-detection phase on the capacity of the surface code to
detect all these single-qubit transformations – σ x ,σ y , and σ z – in isolated or multiple
qubits. Qualitatively, we rely on the expansion of a general completely positive
map as a linear combination of products of Pauli matrices – the Sα operators from
Section 8.4.2. Following this expansion, we interpret generic errors as statistical
superpositions of single-qubit transformation, which can be detected.
The error detection works when changes on the syndrome can be mapped,
unequivocally, to the actual Pauli transformations. For this, we need the expansion
to involve products of few operators, and that these products are localized enough
as not to change the values of one or more logical qubits.11 We achieve this safe
regime by assuming architectures with spatially uncorrelated errors and ensuring
that error monitoring takes place faster than the rate of errors. In this situation, we
can identify not just the errors, but a successful error-correction strategy to bring
back the syndrome to the desired quiescent state Xk,Zj = 0 ∀k,j .
The measurement of the stabilizers, like all other operations in the surface code,
can be implemented the quantum circuit toolbox. The stabilizer operators such as
Z1 , Z2 , or X4 are three- or four-qubit observables that measure the parity12 of
those qubits in either the σ z or σ x basis. This parity measurement requires a single
auxiliary qubit and four two-qubit operations. The algorithm shown in Figure 8.4b
maps the Z-parity of four qubits – i.e., a plaquette – onto an auxiliary qubit that
is placed at the center of it. The extra qubit begins on the |0 state and suffers as
many spin-flips as neighboring qubits are in the |1 state. The outcome is |0 ⊕ P ,
where P = 12 (1 + Z) is 1 if the parity is odd on the plaquette, or 0 otherwise.
A similar algorithm can be constructed for X-parity measurements, with auxiliary
qubits placed on the vertices.
Since the parity measurements demand auxiliary qubits, it is quite common that
these qubits be included in the architecture of the quantum register. For instance,
the surface-code lattice with 18 qubits from Figure 8.4a could be constructed using
the white qubits from Figure 8.2b, with the additional 18 qubits as targets for the
parity measurement.
11 Imagine that the transformations happen to connect the boundaries, as in x = σ x σ x σ x . This error is
2 9 16
actually flipping the logical qubit, and is no longer localized to the degrees of freedom |α that we can safely
manipulate.
12 Take a set of bits s ,s , . . . ,s ∈ {0,1}. The parity is even or odd depending on whether the number of ones is
1 2 n
even or odd. This definition extends to a qubit basis by determining how many qubits return a positive result
+1 when measuring σ z or σ x .
The error-correction phase, in which we bring the syndrome back to the original
value, is implemented as a collection of single-qubit quantum gates acting on strings
of physical qubits. These strings of Pauli matrices run along the edges that connect
vertices or plaquettes with incorrect stabilizer values, correcting all of the errors
simultaneously. The identification of the actual errors and the unitary operations that
undo them is a computationally intensive stage that is still subject to investigation
and improvement.
As an illustration of the difficulty in finding the right error-correction strategy, it
is illustrative to read the article by Kelly et al. (2015). This experiment demonstrated
the operation of a repetition code – a simpler 1D stabilizer code that works with the
parities of neighboring qubits Zi = σiz σi+1 z
– discussing strategies for the detection
of errors. More recently, Chen et al. (2021) demonstrated this code working with
up to 21 qubits and 50 layers of correction using Google’s Sycamore quantum
processor, in a work that also showed an implementation of the surface code with
four qubits, for a more limited number of operations.
Universal Gates
From the previous discussion, we could infer how to implement three logical single-
qubit gates, x , z , and y , and any logical-qubit measurement, in a fault-tolerant
way. The first step is to implement the gate or measure the collective observable, as
prescribed by the definition of the logical operator. The second step involves mea-
suring the syndrome and correct any errors – which in the case of the measurement
might involve updating the measured value to reflect the changes in the physical
qubits.
In practice, the x,y,z gates are not needed. As explained by Fowler et al. (2012),
we only need to care about building a universal set of gates with the H , S, T , and
CNOTs operating on the logical qubits. The x,y,z will be “moved” out of the algo-
rithm and converted in corrections of the measurement outcomes.13 However, this
does not really simplify the task of explaining how the new gates are implemented.
Qualitatively, a standard implementation (Fowler et al., 2012) relies on the pos-
sibility to manipulate the qubit lattice in topologically nontrivial ways. First of
all, to create the surface code lattice, we need to measure the stabilizers of a set
of qubits – the ones that form the lattice – correcting those values to the desired
outcome. This already suggests that we can engineer holes or defects, or cut a
lattice into two separate pieces, by ignoring certain qubits, not considering them
part of any stabilizer. Conversely, we can merge lattices by stabilizer measurements
13 Consider a stupid algorithm that combines X , Z , and a CN OT acting on both qubits C . We can
1 2 12
transform (X1 ⊗ Z2 )C12 = (1 ⊗ Z2 )C12 (X1 ⊗ X2 ) = C12 (X1 ⊗ (−Z2 X2 )). These ideas allow us to move
the Pauli matrices to the end, where they only affect the final measurement outcome.
along their boundaries, and include new qubits or fill holes by including those
qubits in the definition of the stabilizer.14
If we define the logical qubits as strings that connect holes in the lattice, we can
implement CNOT gates in a completely fault-tolerant operation by braiding holes
around each other. The Hadamard gates H are also implemented in a fault-tolerant
way that involves isolating the region of the lattice that contains the qubit, doing
some key measurements, and merging this patch back to the lattice. In both cases,
errors introduced during either the CNOT or the H are self-corrected by the repeated
measurement and corrections of the stabilizer group.
Unfortunately, the S and T gates are more complicated. They must be imple-
mented using quantum algorithms that rely on H , CNOT, and auxiliary logical
qubits prepared in a given state. The preparation of these auxiliary states is a delicate
operation not covered by the stabilizer formalism, and which may introduce errors.
To make these errors as small as possible, we rely on quantum information recipes
that fall outside the code itself. In particular, we use state distillation, which com-
bines multiple imperfect copies of a state to produce one with higher fidelity. All
this makes the S and T operations the most costly of the universal and fault-tolerant
gate set – and a valuable resource that is thoroughly analyzed when designing a
quantum algorithm.
14 In the process, a stabilizer that contained three qubits may pass to contain four, as the lattice has one more
edge.
101
1 − F (%)
100
tgate (ns)
102
101
2008 2010 2012 2014 2016 2018 2020
Year
Figure 8.5 (a) Two-qubit gate errors and (b) two-qubit gate speeds for transmon
qubits. The type of marker denotes the type of gate: CZ (square), CNOT (dot),
cross-resonance (circle) (Chow et al., 2011), iSWAP (star).
15 Note that qubit coherence already follows a Moore’s type law – with an 80% increase in coherence time every
year (cf. Section 6.6) – this naturally leads to a decrease in gate errors. This said, gate operations are more
complicated than just “waiting,” as they are also sensitive to the quality of the controls, distortions in the
cables that bring the control signals, and many other technical sources of errors that need to be overcome.
For instance, we already have first evidence of quantum devices and experiments
that can beat what can be simulated with classical computers (Arute et al., 2019)
while ensuring full control of the experimentalist. In this rewarding time, we are see-
ing new developments in the field of quantum algorithms, adopting new paradigms
where the quantum computer works hand in hand with classical algorithms and
postprocessing techniques. This approach has illuminated new solutions to canoni-
cal (toy) problems in quantum chemistry (Kandala et al., 2019; Arute et al., 2020),
even in the presence of significant errors and decoherence! We expect that the same
platforms will continue to support scientists in studying such interesting concepts in
theoretical physics and quantum many-body physics as quantum transport, many-
body localization, and time crystals (Mi et al., 2022; Neill et al., 2021).
16 https://quantum-computing.ibm.com/.
delegating to the Hilbert space of the quantum computer other taks that might be too
large or too slow on an ordinary computer – for instance, creating and manipulating
the wavefunction of a complex material, or the complex multivariate probability
from a sophisticated risk model.
8.7 Outlook
In this chapter, we have seen how quantum computers are built and operated. The
hardware is conceptually simple. It builds on the tools developed in Chapters 6
17 By generic, we mean any rotation in SU (4) covering all of the Hilbert space for two qubits. We also mean that
those gates can take place among any pair of qubits. In practice, the arbitrary separation means that quantum
gates must be decomposed – transpiled – into a much larger number of assembly gates (Cross et al., 2019).
This is really what makes the quantum volume a challenging metric.
and 7, with additional techniques for two-qubit quantum gates and for characterizing
the computer’s performance.
It is clear we need further developments in error correction and fault-tolerant
computation models. Existing paradigms impose significant overheads – number of
measurements, postprocessing of data, and rapid feedback controls – and demand
orders of magnitude more qubits than what we are likely to supply in the following
decade. We also need a better understanding of quantum circuits, algorithms, and
compilation strategies.
However, there are also many architectural, design, and theoretical challenges
open. We can highlight the following ones:
– We need better qubits, with longer coherence times. There are hints that this
can improve with new materials and the application of state-of-the-art surface
treatment and fabrication (Place et al., 2021). But maybe we also need different
types of qubits, with intrinsic topological protection!
– Complex algorithms and error correction require us to pack more qubits into the
chip. This is an important challenge, because the use of microwave technology
sets a lower bound on the size of qubits and control devices. Maybe the route
to success lays in going beyond existing sandwich architectures to fully three-
dimensional integrated devices (Brecht et al., 2017, 2016).
– If we increase the number of qubits means, we need to increase commensurately
the number of readout and control lines. This requires important developments in
the control and microwave electronics – which already has shifted to customized
FPGA circuitry (Arute et al., 2019) – and clever strategies to optimize the limited
number of microwave lines that can be inserted into the fridge. If at some point
even this becomes an unavoidable limitation, we might need to transition to
distributed quantum computers connected by quantum links, as explained in
Section 7.2.5.
– We need better controls and higher-quality gates. This challenge may be
addressed at the hardware level, looking for new interconnects between qubits,
producing different, more scalable or stronger couplings. Or it may be addressed
at a quantum control level, improving on the design of quantum gates, making
them more robust against external perturbations, or even against the obvious
distortions suffered by microwaves when they are injected into a chip.
These are very important and very difficult challenges. I hope that the information
in this book, combined with existing literature and your ingenuity, will stimulate you
to address them, joining many of the excellent teams working in the field worldwide,
or working to create your own!
Exercises
8.1 Prove that a Hadamard gate acting on each qubit creates a quantum superpo-
sition of all possible binary numbers in a quantum register:
N −1
2
⊗N 1
H |00 . . . 0 = |b . (8.27)
2N/2 b=0
(5) Prove that it is possible to recreate any control-phase gate, with any angle
exp(iθ σ1z ⊗ σ2z ), assuming arbitrarily long adiabatic passages with no
decoherence. Hint: Show that for a large enough passage time T , there
is a nonzero interaction phase θ. Then prove that the value of θ can
be enlarged arbitrarily by deforming the adiabatic passage ω2 (t × T /T ),
with T > T .
(6) Find out the algorithm for the quantum Fourier transform. What phases
θ are required in that algorithm?
8.5 We want to simulate time-evolution with the Ising model, implementing the
unitary operation
U (t) = exp −itσ1z σ2z , (8.30)
but we only have a toolbox with control-NOTs and arbitrary single-qubit oper-
ations. Devise a protocol to approximate U (t) using such gates. Hint: Inves-
tigate what happens when you sandwich a local operation in between two
CNOTs.
8.6 Useful positive maps. We now study a set of very common positive maps.
(1) Completely depolarizing channel. Provide a physical interpretation for
the completely depolarizing channel in Table 8.2. Show that the symmetric
version is recovered from the one in that table by setting px = py = pz =
p/3.
(2) Random flip. Write down a completely positive map to describe this
physical operation: With probability p, preserve the state and elsewhere
flip the state of the qubit, exchanging |0 and |1.
(3) Dephasing. Show that the dephasing of a qubit (6.14) may be written as
a channel that with probability p preserves the state, and elsewhere keeps
only the diagonal elements:
1−p
ε(ρ) = pρ + tr(ρ)1 + (1 − p)tr(σ z ρ)σ z .
2
(4) In all previous maps, determine the Kraus operators and show that the map
preserves the trace of a density matrix: trε(ρ) = trρ.
8.7 Prove that the set of operators formed by all tensor products of Pauli matrices
Sα ∈ {1,σ x ,σ y ,σ z }⊗N (cf. Section 8.4.2) forms a basis of linearly indepen-
dent and orthogonal elements in which all matrices can be expanded. Hint:
Reinterpret matrices as column vectors, show that (A,B) = tr(A† B) is a good
scalar product for the matrix space, and prove the expansion (8.13).
8.8 Derive the Choi matrices (8.17) using the M and χ representations. Hint: Show
that |α = √1d di=1 |i ⊗ Sα |i forms an orthonormal basis.
8.9 Deduce the form of the matrices M and χ for a perfectly coherent positive
map, ε(ρ) = U ρU † , in terms of the unitary operation U . Particularize it for a
single-qubit rotation U = exp(−iθ n · σ ) and a CNOT.
8.10 Compute the process matrices M and χ for the single-qubit error map from
(6.19). Compute the average fidelity and the process fidelity for this map.
How do those figures relate to T1 and T2∗ ?
8.11 Design a randomized benchmarking protocol for a single qubit. What gates
do you need to use? Simulate this protocol for a qubit that experiences relax-
ation (T1 = 1 μs) and one that experiences dephasing (Tφ = 1 μs), and has
no SPAM errors. In the simulation, assume a fixed duration τ ∼ 10 ns for
all single-qubit gates, and that the error of the gate can be fully attributed to
decoherence, that is, εU (ρ) = εdecoh (U ρU † ). Fit the success probability Pl to
the exponential curve (8.23) and extract the parameter p. Compare the error
estimate p with the estimated average and process fidelity associated to εU (ρ).
8.12 The average fidelity (8.20) results from the fidelity of a physical process over
all pure states. It has been shown by Nielsen (2002) that this definition is
equivalent to replacing the map ε(ρ) with a totally depolarizing channel (see
Table 8.2) with depolarization p. Show that the relation between p and Favg
is given by (8.24).
239
Figure 9.1 (a) Sketch of the energies and gaps throughout an adiabatic passage
in which a Hamiltonian Ĥ () is deformed from (t = 0) to time (t = T ). The
Hamiltonian maintains a gap between the excited states and the ground state energy
EGS , which is at least Emin . (b) Qubit hyperbola for a flux qubit that begin
in its ground state and is subject to a linearly growing bias field, from a very
large negative value ε |min | to a very large positive one. In this Landau–Zener
process, there is a minimum qubit gap min , depending on which the process will
be adiabatic – the qubit remains in the ground state manyfold and |1 is deformed to
|0 – or diabatic – there are transitions to the excited state, leaving some probability
in |1 at the end.
Ĥt = H (ε(t)). We start the adiabatic passage from an eigenstate of the Hamiltonian
Ĥ0 , which we assume is easy to prepare. We change the control in time ε(t),
expecting that our experiment remains as close as possible to the ground state of
the time-dependent Hamiltonian Ĥt .
Implicit in the notion of adiabatic passage is the idea that ε can change slow
enough for the system to adapt and remain in a quasi-equilibrium state without
exciting to higher-energy states. To pose some classical analogs, think of how you
can adiabatically stretch a rubber band. If you stretch it too fast, it will acquire too
much energy, exciting phonons that show as vibrations of the band. Or think about
how you can push a glass full of water across a table. If you push too fast, the surface
of the water will deform and even spill out of the glass, but if you push it gently, the
surface of the liquid will remain unperturbed.
We now address the idea of adiabatic passage from two points of view. We start
with a very specific and simple model, a qubit, where we can solve the problem
analytically and even provide a quantitative answer to the shape of the control. This
problem will give us an intuition about adiabatic and diabatic processes, which
can be applied to many different problems. This intuition then acquires a rigor-
ous mathematical framework through the adiabatic theorems, a set of sufficient
criteria to ensure an adiabatic passage for every given Hamiltonian. This theorem
lays the foundations for the adiabatic quantum computation (AQC) model that we
discuss next.
that depends on the polar angle n = (sin θ,0, cos θ ). In the limit of large bias
|ε| |, the ground state of the Hamiltonian approximates the two current
states, (0,0,+1) or |1 = |R, for very negative field and (0,0, − 1) or |0 = |L
for very large positive bias. These are two states that we can prepare, either by
waiting for the qubit to relax to the ground state at those extreme points, or using
some type of measure-and-reset protocol — e.g., we could dispersively measure
the qubit’s state and flip it using a microwave if the qubit was in the excited
current state.
We now wonder what happens if we start from such a well-prepared ground state
and change the control parameter ε, as in Figure 9.1b. Does the qubit remain in the
instantaneous ground state or does it get excited? In the first case, which we call the
adiabatic evolution, the qubit’s wavefunction is always the one in (9.3) up to global
phases. In the second case, which is called the diabatic evolution, the qubit will not
adapt, and it will either get locked in the original current state or, more generally,
become a quantum superposition of both qubit states.
1 Note that, due to our unitary transformation, |0 and |1 are no longer current states, but an instantaneous
ground state and excited state.
2 The theorems work for one state, or for a collective of states that are very close in energy, degenerate, or
almost degenerate. We do not care much what happens within this collective of states, but we want to preserve
all probability within this subspace.
3 This said, we encourage the reader to have a look at the cited works, which are long, but relatively easy to
follow.
5 Here “gapped” means that there is an energy gap between the ground state manyfold that encodes the
computation and all other states.
The first connection was formalized by different groups, starting from Kitaev et al.
(2002) until the more recent work by Biamonte and Love (2008), which creates
apparently simple models for arbitrary quantum circuits. We will explore this
mapping in the context of Section 9.3, where we worry about which Hamiltonians
represent useful and hard computations, and which are physically implementable.
But first, let us explore how realistic it is to satisfy the second requirement, which
is building the ground state of Ĥp .
1 d min
(εt ) . (9.10)
min dt h̄
(3) Measure and characterize the final state |(T ) |ψp .
The adiabatic algorithm works by smoothly distorting a state that is easy to
prepare, typically an uncorrelated product state, through an adiabatic passage
that is guaranteed to succeed by the adiabatic theorem. Unlike other quantum
computing paradigms, such as the quantum circuits from Chapter 8, the construction
of |ψp is a continuous process, not a combination of simple, well-controlled
operations.
One important question in the AQC algorithm is the choice of initial Hamiltonian
H0 . Ideally, we want a model whose ground state |ψ0 has some overlap with the
final solution, because if we do not achieve this, the adiabatic passage might have a
vanishingly small gap. A common choice is to prepare a product state that contains
a quantum superposition of all states in the computational basis:
⊗N
⊗N 1 1 1
|(0) = |+ = √ |0 + √ |1 = N/2 |ss| . (9.11)
2 2 2 s
This state can be prepared as the ground state of a noninteracting model, a collection
of transverse fields:
N
H0 = − σmx . (9.12)
m=1
The adiabatic protocol works by interpolating between both limits, starting with
large external fields and small interactions, and gradually increasing the strength of
the couplings in Hp .
implement H (ε) – and the duration of the adiabatic passage. In AQC, the size
of the simulator grows gently with the problem size. Time, on the other hand, is
usually quantified as the time to solution, the duration of an adiabatic passage that
guarantees we reach the ground state with a minimum probability p.
As discussed earlier and in the context of the Landau–Zener processes, the time
to solution depends on the minimum energy gap of the problem Hamiltonian. This
gap typically shrinks with growing problem size, with different scalings depending
on the problem’s difficulty – exponential T ∼ e−αN for difficult NP problems,
algebraically for “manageable” problems T ∼ N −α , as discussed in Section 9.3.4.
The relation between time and gap shows how this resource can be distorted by
our choice of Hamiltonian. Imagine we can guarantee an adiabatic passage in time T
for a problem Hamiltonian Ĥ with gap min . If we can implement a simulator for
the rescaled Hamiltonian Ĥ = 2Ĥ , 3Ĥ , or even larger, we will arbitrarily increase
the gap and arbitrarily contract the time to solution.
In practice, the strength of interactions available in the lab is finite and the total
available energy Emax = maxε !H (ε)! = maxε {max σ (H (ε))}, is bounded. There-
fore, when comparing different adiabatic quantum algorithms it is often useful to
work with a rescaled dimensionless time τ := T Emax /h̄, which is free from this
uncertainty and scales more adequately with the problem size N . The different
accounting in time and size will pop up when we compare AQC’s with universal
gate-based quantum computers, when we discuss the difficulty of adiabatic compu-
tations, and in Section 9.4.3, where we discuss this paradigm critically.
This suggests that we can explore the relation between problem Hamiltonians and
families of computational problems F (s) with different complexity. We will con-
firm this idea, deriving Hamiltonians for difficult families of classical and quantum
problems, called NP-hard and QMA-complete. But first, we need to know these
classes.
6 Indeed, how do you know if the route is optimal? Note, however, that answering the question, “is there a route
of length L?” is NP-complete.
well-known examples are the quantum Fourier transform and Shor’s factoring algo-
rithm. From the definitions, using the fact that a quantum computer can reproduce
any classical computation, we can infer the following relations:
P ⊂ NP ⊂ QMA, and P ⊂ BQP ⊂ QMA. (9.14)
We believe that BQP contains problems that are not in P. There is evidence of this
in the difficulty of simulating the evolution of quantum many-body systems, or even
relatively “small” random quantum circuits, where entanglement grows quickly. It
is also widely believed in the academic world7 that NP is not contained in BQP. This
means that quantum computers cannot exponentially speed up the solution of those
problems, even though one may not rule out other less dramatic accelerations.
In the following two subsections, we will show that both NP and QMA problems
can be restated as the task of preparing the ground state of a Hamiltonian. We will
identify specific representatives of those Hamiltonians that are universal, in the
sense that they are NP- or QMA-complete and can be used to solve any problem
in those classes. Interestingly, both will turn out to be models of interacting qubits
that are good candidates to be implemented in the laboratory.
Barahona (1982) showed that computing the ground state of this model is at least
an NP-complete problem, provided the spins are arranged on a planar graph,8 inter-
actions are antiferromagnetic along the edges that connect qubits Jij = J > 0, and
we can control the sign of the magnetic fields, which have comparable strength to
the interactions |hj | ∼ J .
The spin glass in (9.15) is a diagonal Hamiltonian (9.13). Finding the ground state
of this problem is thus equivalent to finding the minimum of a classical function. If
we move from sign variables si ∈ {−1, + 1}, to boolean variables xi = 12 (si + 1) ∈
{0,1}, the spin glass become an instance of a quadratic binary optimization (QUBO)
problem:
F (x) = xi Qij xj , xi ∈ {0,1}. (9.16)
ij
The local controls and couplings in ĤpQMA can be derived constructively from
a quantum circuit. Ideally, we could simulate the XXZZ Hamiltonian in our setup,
prepare the ground state using the AQC algorithm, measure its energy – this amounts
to measuring Pauli operators, as shown in Problem 9.5 – and verify that energy lies
below a prescribed bound, to determine the success of the computation. Moreover,
using the same setup, we could also address other NP-complete problems.9
In practice, this introduces two technical difficulties. The first one is to ensure a
topology of qubits that maps to the problems we wish to solve, or that at least can
provide suitable embeddings. The second problem is how to simulate the XXZZ
interactions, when the qubits that we have studied only seemed to have one type of
dipolar moment.
One solution to this problem is to use flux qubits, and engineer them to enable
both inductive and capacitive interactions. If we use current superposition as basis
9 Note how (9.17) contains the NP-hard spin-glass model (9.15) in the limit = t = 0.
However, if we make the Ising interaction random,10 then suddenly the phase√tran-
sition from paramagnet to spin glass will scale much worse, min ∝ exp(−g N ).
Here, you recognize the spin-glass model (9.15) in combination with the initial
bias field H0 ∝ (t) from which we start the adiabatic passage (9.12). D-Wave’s
machines can be programmed to change in time all parameters from the effective
Hamiltonian (9.18), allowing for a good dynamical range and determination of the
sign in all variables. This enables the users to implement an adiabatic passage in
which (t) ∝ (1 − ε(t)) is decreased, while the problem Hamiltonian is enhanced
J (t),h(t) ∝ ε(t).
The quantum simulation, using superconducting devices, of an NP-hard Hamilto-
nian with time-dependent quantum fluctuations (9.18) is a great scientific achieve-
ment and a great technological challenge worth studying. We will discuss the
devices’ architecture, how the annealer works in practice, its performance in
comparison to classical algorithms, and how “quantum” the D-Wave machine
really seems to be.
(a) (c)
(b) (d)
3
1.5
–1.5
–3
-0.4 0 0.4
Figure 9.2 (a) Tunable coupling circuit for two flux qubits with an interac-
tion mediated by an rf-SQUID. (b) Experimentally measured mutual inductance
between the two flux qubits for this particular circuit. (c) Detail of a flux qubit
design, including elements to tune the qubit frequency (LT), tune the qubit gap
(CCJJ), induce an external field (IPC) and read-out the qubit state (RO). (d) Archi-
tectural design of a cell with eight qubits in the D-Wave chimera graph. All qubits
have local controls, tunable interactions and readout, and talk to each other through
the ICO couplers. Different cells talk to each other via the XCO couplers. Figure
adapted from Harris et al. (2009) and Harris et al. (2010b) with permissions.
To implement all these features and simulate the quantum annealing Hamiltonian,
D-Wave has introduced changes in the qubit design (Harris et al., 2010a,b). The
changes, illustrated in Figure 9.2b, work as follows:
(1) The tunable Josephson junction of the rf-SQUID is implemented by a com-
plicated circuit, called a compound Josephson junction (CCJJ). Introduced in
Harris et al. (2010a), this circuit works like a dc-SQUID, creating a nonlin-
ear inductive potential EJ (xccjj ) cos(φ/ϕ0 ) for the qubit’s flux variable φ. The
Josephson energy EJ depends on the external flux that passes through the CCJJ.
Changing this flux modifies Josephson barrier EJ between potential wells. In
other words, the field xccjj may increase or decrease the tunneling amplitude
(xccjj ) in Hamiltonian (9.18).
(2) Changes in the qubit’s Josephson energy affect the critical current in the loop,
as well as the inductive coupling to the external fields and to other qubits.
The qubit’s current is inversely proportional to the inductance of the loop
IL,R ∼ ±φ/L. We can adjust and compensate for changes in this value by
inserting a dc-SQUID, labeled LT in Figure 9.2c. The dc-SQUID operates in
the linear regime as a tunable inductor L() (see Section 4.7.2), whose value
can be decreased to increase the current, and vice versa.
(3) An adjustable inductor – labeled IPC in Figure 9.2c – mediates the interaction
of the qubit with a current source Ig . By adjusting this inductor, we can control
how much flux enters the qubit,11 thereby controlling the effective magnetic
field 12 h̄hi (t)σ z in (9.18).
(4) Another tunable inductor controls the coupling between the qubit and a readout
(RO) device. The RO is a current-biased CCJJ that operates very similarly to a
current-biased dc-SQUID (cf. Section 4.7.2). During a measurement, the tun-
able inductor is activated, so as to inject a small amount of flux into the CCJJ
δφ = δφ0 + ησ z . If the current bias and η are both large enough, the CCJJ
will switch to a voltage state when the qubit is in one of the states σ z = +1,
remaining inert in the other case. As explained in Berkley et al. (2010), this
implements the almost projective measurement described in Section 6.4.4. Nat-
urally, to minimize decoherence, at all other times the mediator is deactivated,
eliminating any residual coupling to the RO device (η = 0), and minimizing
decoherence.
We see in this design an abundant use of tunable inductors as control elements
for qubit gaps, magnetic fields, qubit potentials, measurement, and decoherence. A
simple counting of elements in Figure 9.2c reveals a total of six hardware controls
per qubit – one flux control for the inductor, one for the qubit gap, two for the
measurement, and two for the bias. Some of these controls are stabilized with the
use of the CCJJ, a device that allows better tolerance against fabrication errors – for
instance, small variations in the Josephson energies in the coupler circuit – and a
perfect cancellation of inductance.
11 See Section 6.4.2 and around (6.56), with the convention that now |L and |R are the qubit basis states.
12 Section 6.5 defined the dipolar moment operator in the qubit basis at the degeneracy point. However, in this
chapter we have changed basis, and the dipolar moment operator is defined in term of current states
σ z = |LL| − |RR|.
in Figure 9.2. It is, however, static in nature, because the mutual inductance only
depends on the topology and geometry of the qubits’ intersection.
In order to achieve tunability of the coupling J matrix, D-Wave’s researchers
added tunable inductors that mediate additional couplings. As sketched in
Figure 9.2a and d, D-Wave’s mediator is an rf-SQUID that contains a CCJJ
circuit instead of a junction. The SQUID implements the interaction mechanism
from Section 6.5.4, but now with two controls: the SQUID’s internal flux and the
amplitude of the Josephson energy.
Figure 9.2b illustrates the dynamical range of the qubit–qubit mutual inductance:
One may control not only the amplitude, but also the sign of every qubit–qubit
interaction, reinforcing, canceling, or even reversing the geometric inductance pre-
viously discussed. As for the orders of magnitude, the mutual inductance shown
before ranges around ±1 pH, a very decent value. Considering that the qubit persis-
tent currents are |Ip | 1.0 μA for each of the two possible orientations, this gives a
safe estimate for the interaction energy E = M|Ip |2 10−24 J, which corresponds
to a frequency ω E/h̄ 2π× GHz. This value is consistent with the 4 GHz
antiferromagnetic splitting measured spectroscopically for qubits within a single
cell of the D-Wave architecture (Berkley et al., 2013; Lanting et al., 2014).
then interacts with every neighboring cell via four of its qubits. For instance, in
Figure 9.2d, qubits 4, 5, 6, and 7 connect to neighboring cells on the same column,
while qubits 0, 1, 2, and 3 propagate interactions along a row. This sparse, nonplanar
topology is called the Chimera graph.
Our first goal is to find an embedding of the problem Hamiltonian Hp as an
equivalent model Hp implemented in the Chimera graph. Usually, the embedding
is done by directly identifying the logical qubits of Ĥp as a subset of the physical
qubits in the destination graph. Other qubits within the graph are used to mediate
interactions between the logical bits. The embedding must be such that Ĥp and
Ĥp exhibit similar low-energy eigenenergies, and from every configuration of the
embedded model |S we can deduce the logical bits |s of the desired model.
In some lucky cases, the graph that we want to simulate can be found as a sub-
graph – a minor – of the Chimera graph. In this case, we need no auxiliary qubits
and the interactions can be mapped directly to couplings in the Chimera lattice. For
instance, the four-qubit model H = J σaz (σbz + σcz + σdz ) can be embedded into the
unit cell from Figure 9.2d, identifying (a,b,c,d) → (q4,q0,q1,q2 ) in the unit cell,
and setting J40 = J41 = J42 = J43 = J in the annealer’s controls.
Other times we need to copy a qubit, duplicating its value, because there is not
an exact one-to-one embedding. For example, think of the triangular lattice cell
H = J (σaz σbz + σbz σcz + σcz σaz ). We can implement this model with the logical qubit
assignment (a,b,c) → (q4,q0,q1 ), and copy the value of q0 to q5 with a strong
ferromagnetic interaction. This means we will have a coupling matrix J40 = J41 =
J51 = J , and J50 = −Jextra , with a large constant |Jextra | |J |.
More generally, we need to engineer arbitrary interactions. The technique of
perturbative gadgets by Kempe et al. (2006) allows the engineering of rather general
effective interactions starting from an experimental platform. As before, we identify
logical qubits with physical qubits, and add auxiliary qubits and interactions to
reproduce the target model. However, this framework is more systematic and
powerful, enabling the simulation of arbitrary two-, three- and other many-body
interactions between the logical qubits, which are engineered from the simple-
looking Hamiltonian in (9.18).
Embedding and gadgets extend the family of models accessible to a given archi-
tecture, but they have associated costs. First, the use of auxiliary qubits increases the
size of the computation. Venturelli et al. (2015) studied the worst-case scenario of a
fully connected network and found that the Chimera graph requires O(N 2 ) physical
qubits to represent N logical ones. Second, the search for a good embedding is
itself an expensive classical overhead – minor embedding is an NP-hard problem.
Therefore, in many cases we must resort to a suboptimal, but good-enough heuristic
(Cai et al., 2014) for the simulation. Third, some gadgets will produce interactions
that can be perturbatively small, leading to a decrease in the available energies Emax
and an increase of time. Finally, some embeddings require many auxiliary qubits and
long chains of interactions. The transverse magnetic field combined with these long
chains produces a contraction of the gap13 min and an increase of passage time.
13 Qualitatively, this contraction is caused by the difficulty of transverse fluctuations σ z in effecting a change
i i
over a string of qubits that want to be on the same state (see Problem 9.6).
14 See figure 5 in Harris et al. (2010b) and how a change of x by an amount 0.02 causes an increase of I
ccjj 0 J
from 0 to 1.6 μA, and an exponential decrease in from 109 to 105 Hz.
The programming and measurement times are constant overheads that do not
depend so much on the problem at hand. These can be very large, with respective
values around 7 ms and 125 μs reported for a D-Wave 2000Q by Albash and Lidar
(2018). The annealing time is problem dependent, but it is usually the shortest value
in the whole experiment. Due to strong filtering of controls to limit the influence
of high-frequency noise, the shortest annealing time starts at 20 μs in the D-Wave
Two device and 5 μs in D-Wave One, Two X, and 2000Q. This minimal passage
time can obscure or perturb the interpretation of resource scaling, as explained by
Mandrá et al. (2016).
15 In terms of qubits, it may be argued that ultracold atoms in optical lattices have larger systems. However, as of
this writing, they have not exhibited a similar control of both local fields and separate interactions between
neighboring quantum systems.
that interacts with other qubits that implement the desired Hamiltonian (9.18). The
setup is described by the effective model
0 x ε0 z 1
Ĥ = σ + σ0 + J σ0z σkz + εcomp σ0z + Ĥp . (9.19)
2 0 2 2
We assume that the probe qubit σ0z interacts with only one other qubit σkz from the
many-body simulator Ĥp . We control the probe’s bias ε0 and tunneling , as well as
a compensating field acting on the contacted qubit. When this field is εcomp = 2J ,
the eigenstates of H split into two manifolds, with different values of σ0z = ±1,
which are coupled by a weak tunneling 0 :
1 1
Ĥ = Ĥp + |10 1| ⊗ ε0 + 2J σk + 0 1 ⊗ σkx .
z
(9.20)
2 2
Decoherence
Despite the evidence of coherent macroscopic quantum tunneling events (Lanting
et al., 2014), and simulation of quantum many-body phase transitions (Harris et al.,
2018; King et al., 2018), D-Wave quantum processors up until 2019 have too large
decoherence rates to be considered an adiabatic quantum processor.
The role of decoherence can be studied at the single-qubit level. Harris et al.
(2010b) analyzed the tunneling between current states in a qubit at = 0, to reveal
fluctuations in the qubit’s flux and environmentally induced dephasing. Later,
Berkley et al. (2013) used qubit tunneling spectroscopy to analyze the spectra
and coherence of one qubit. Both studies suggested a large decoherence rate, with
qubit linewidths 0.7 GHz, only an order of magnitude away from the qubit–
qubit interactions. This sets the decoherence time T2∗ ∼ 1 ns, three to six orders of
magnitude worse than state-of-the-art superconducting qubits (cf. Section 6.6), and
much shorter than the shortest passage times ∼ 5 μs.
Harris et al. (2010b) further studied the behavior of a small eight-qubit quantum
simulator when subject to a control designed for an adiabatic passage. They showed
that the dynamics of the qubits could be modeled using a master equation that
combines the interpolating Hamiltonian with a finite heating rate – given by the
effective temperature of the chip – and a strong dephasing rate – essentially the
T2∗ of a single qubit. Even if the low temperatures were unable to bridge the gap
of this relatively small system, and the experiment could prepare ground states
with high fidelity, the study evidences that the superconducting processor does not
implement an adiabatic quantum computation in the sense of Section 9.2. Instead,
we can describe this dynamics as an implementation of quantum annealing.
Computational Advantages
One original motivation for constructing the flux-qubit quantum simulator of the
spin-glass Hamiltonian (9.18) was the possibility of solving NP-complete problems,
and whether quantum effects – fluctuations, entanglement – can improve the effi-
ciency in finding those solutions. It is for this reason that many works have analyzed
the scaling of the physical quantum annealer, comparing it to other classical meth-
ods, with various results.
A first step in these analyses is figuring out the right metric. It turns out that
this is the time to solution, which is the shortest execution time, including rep-
etitions, that guarantees a solution with a high success probability, such as 99%
(see Problem 9.7). A second step is to choose the right classical algorithms for
the comparison. Here the focus has been on generic algorithms that work for all
problems, namely spin dynamics, simulated annealing, spin vector Monte Carlo,
and Monte Carlo simulation of quantum annealing passages. This ignores to some
extent the fact that some problems have specific tailored methods that are very
efficient.17
Denchev et al. (2016) is one of the first works to prove a constant speedup in the
physical quantum annealing (QA) as compared to the classical method, with up to
eight orders of magnitude. They found a similar scaling between SQA and QA with
the problem size, but the Monte Carlo method was about 108 times slower in real
time. Later works by Mandrá et al. (2016) and Mandrá and Katzgraber (2018) have
extended the comparison to other sequential and parallelized methods, considering
also some tailored algorithms. These works find also a similar scaling between
SQA and QA, with a constant speedup factor in time to solution, although they
also realize that this speedup could be compromised by minimum passage time
and hint at the need to find the optimal time over all possible passages. Albash
and Lidar (2018) performed the optimal time-to-solution analysis for the first time,
comparing hard problems with a planted solution. This study confirms the speedup
16 This is a method introduced by Shin et al. (2014) that represents the quantum register as a product state of
uncorrelated spins, updated in an easy-to-simulate classical method that accounts for some of the quantum
fluctuations.
17 For instance, there is a very clever and very efficient algorithm that is specific for solving QUBO problems in
the Chimera graph topology, due to Selby (2014).
There are some obvious paths to move ahead, both in the field of quantum anneal-
ing and in quantum simulation (Hauke et al., 2020):
– Study new annealing schemes, which may involve the use of different schedules
for different qubits, random perturbations, or different interaction schemes. Some
of these techniques may eliminate first-order phase transitions, or allow those to
happen in smaller regions that get connected and produce better solutions.
– Achieve better coherence times in qubits that are compatible with quantum
annealer designs. Some examples are recent designs of capacitively shunted flux
qubits, which feature slower dephasing rates (Steffen et al., 2010; Yan et al.,
2016).
– Engineer new interaction schemes between superconducting qubits, to create
nonstoquastic couplings of σix σjx simultaneous with σiz σjz . This is not a simple
task, but there are evidences for such couplings with flux qubits by Ozfidan
et al. (2020), and there are even designs to enable three- and four-body couplings
(Chancellor et al., 2017).
However, if we have learned something from the discussion and research with
D-Wave’s processors, it is that sometimes we can make something very good for the
wrong reasons. Superconducting quantum annealers were born out of commercial
interest around computational applications, which failed at demonstrating a quan-
tum advantage. However, the result is still a unique demonstration of the potential
for scalable quantum simulations with superconducting quantum circuits. Maybe
one possible future of this technology – microwave circuits and qubits – lies pre-
cisely in avenues that explore this potential, researching fundamental and practical
problems at the interface between quantum optics, condensed-matter physics, and
other fields with complex quantum systems. I hope you, as a reader, will be able to
bring this new insight into the field, and use the tools from this book to construct
those new applications.
Exercises
9.1 Let us analyze the Landau–Zener model (9.2).
(1) Show that Ĥ (ε) = 12 (ε)U σ z U † where U = exp(−iθ(ε)σ y /2).
(2) Rewrite the wavefunction of the two-level system in a rotating frame with
|ψ(t) = U (t) |ξ (t) and show that the rotated state |ξ (t) evolves with the
Hamiltonian:
1 min ε̇ y !
Ĥeff = (ε)σ z − σ ξ (t) . (9.21)
2 22
(3) Implement a change of variables t → ε, using ∂t = ε̇∂ε , to recover the
effective equation (9.5).
(4) Provide an upper bound on the rate of change of the prefactor in front of
σ y and compare it with the σ z term.
(5) Approximately diagonalize the effective Hamiltonian, showing that the
2 × ≤
|0 eigenstate of σ z gets a correction of order min 2ε̇ 2ε̇
min
.
9.2 Let’s assume that we have a qubit with frequency , interacting with a
microwave cavity of frequency ω through a coupling strength g, as described
Naturally, the precision of the simulation increases when we decrease the time
step t = T /L and increase the total passage time T .
(1) Show that evolution with Ĥ0 decomposes into local rotations.
(2) How many quantum gates do we need to simulate evolution with the spin-
glass model (9.15) if the qubits are organized in the 2D and 3D lattices
studied by Barahona (1982)?
9.5 Take the QMA-complete Hamiltonian from (9.17). Show that the energy Ĥ
of any quantum state can be estimated by repeating M times the measure-
ment18 of either all σiz or all σix Pauli observables√on all qubits. Show that
the uncertainty in these measurements scales as 1/ M. Assume we want to
keep the experimental precision below some bound ε. Prove that the number
of measurements scales at most as O(N 4 /ε), assuming the worst-case scenario
of N 2 terms in (9.17).
9.6 Assume we want to establish an interaction J between two logical qubits,
assigned to positions 1 and N + 1, which are connected by an intermediate
chain of N physical qubits, as in
1
Ĥ = h̄ σ1z + σN+1 z
− Jf σ1x σ2x + · · · + σN−1
x
σNx + J σNx σN+1
x
. (9.23)
2
Compute the minimum gap of this model in the limit of large ferromagnetic
coupling |Jf |. If we make an adiabatic passage with this Hamiltonian, using
18 This is called a two-setup experiment, because experimentalists only have to do two kinds of experiments that
only differ in the final measurement at the end.
Ĥ0 (9.12), what is the probability that the state of qubit 1 is copied to the N
intermediate qubits?
9.7 The usual metric in annealing experiments is the time to solution, that is,
the time tTTS required to find the solution with a fixed probability, typically
pTTS = 99%. Assume that we perform passages with a time tramp and find the
true ground state with a probability pramp < 99%. Compute the number of
times M that we have to repeat the annealing experiment to obtain a success
probability p ≥ 99.99%. Discuss what happens to tTTS when the ramp time is
lower-bounded due to the restrictions of the device.
1 2π n 2π
m = √ exp(ikn m),
u(n) kn = ∈ × {0,1,2, . . . N − 1}, (A.3)
N N N
that naturally satisfy the boundary conditions, um+N = um exp(ikN ) = um exp(i2π n). The
eigenvalues associated form a band λn = a + 2b cos(kn ).
It is quite often more convenient to define the quasimomenta kn so that they explore the
interval (−π,π ]. For even sizes N = 2M,
π
kn ∈ × {−M + 1, − M + 2, . . . − 1,0,1, . . . M}, (A.4)
2M
while for odd one N = 2M + 1
π
kn ∈ × {−M, − M + 1, . . . − 1,0,1, . . . M}. (A.5)
2M + 1
270
We can also use real wavefunctions if we instead diagonalize B with a real orthogonal
transformation, B = O T O, O ∈ RN×N , O T O = 1. The eigenfunctions are then linear
sinusoidal functions:
2
u(even,n)
m = cos(kn m), kn ∈ (0,π ) (A.6)
N
2
(odd,n)
um = sin(kn m),
N
plus the two extra solutions:
1
u(even,0)
m = , k0 = 0, (A.7)
N
1
u(even,n)
m = (−1)m, kn = π . (A.8)
N
These equations are equivalent to a periodic boundary condition problem (A.2) of size
2(N + 1), where one imposes uN+1 = u2(N+1) = 0. This solution is constructed from
two degenerate plane waves with opposite momenta and the same eigenenergy:
Unlike the periodic case, we can no longer shift the linear momenta, which are constrained
to the interval [0,π ).
1 Alternatively, you may use more sophisticated derivations such as the one by Bravyi et al. (2011), on which
this appendix is based.
This allows us to assume Vnm := n|V̂od |m = 0 when n = m. We can then expand our
generator g Ŝ = g Ŝ0 + O(g 2 ), with an operator Ŝ0 that cancels all first-order terms in the
expansion of Ĥeff , producing [g Ŝ, Ĥ0 ] = −g V̂od .
In the nondegenerate perturbation theory limit, this equation can be solved formally,
projecting onto the original eigenstate basis:
which is solved by
Vnm
Ŝ0 = × |nm| . (A.29)
En − Em
n=m
We will use these expansions in several qubit designs, such as when justifying the truncation
to a two-level subspace in the charge qubit (Problem 6.6), or when analyzing the residual
interactions in two highly detuned qubits (see Section 6.5.3 and Problems 6.16 and 6.17).
This structure changes the equation we have to solve for Ŝ0 and V̂ . We need to split
V̂ = V̂d + V̂od into contributions that “live” within the same energy subspace and that will
be incorporated into Ĥ0 and contributions V̂od that connect different subspaces:
P0,β [Ŝ, Ĥ0 ]P0,α = P0,β V̂ P0,α , α = β. (A.33)
These equations can be solved, producing the anti-Hermitian operator:
1
Ŝ = |n,αn,α|V̂ |n,βn,β| . (A.34)
E − Em,β
n,m n,αα=β
We can continue further with the perturbation theory argument, substituting this operator
into the series for Ĥeff and inspecting the box within one particular subspace. The result is
a generalization of the nondegenerate perturbation series:
1
α
Ĥeff = P0α Ĥ0 + g V̂d P0α + g 2 P0α Ŝ, V̂od P0α + O(g 3 ). (A.35)
2
Explicitely in terms of eigenstates,
α
Ĥeff En,α |n,αn,α| + n,α|V̂ |m,α |n,αn,α| (A.36)
n n,m
g2 1 1
+ +
2 En,α − Em,β Er,α − Em,β
β=α n,r=m
A.3.3 Considerations
Equations (A.30) and (A.36) are mathematical recipes for an effective Hamiltonian that is
expressed in the basis of eigenstates of the unperturbed model Ĥ0 . From this point of view,
the appeal of this method to the design of quantum simulators and quantum engineering is
clear: This theory allows us to design effective interactions, simply by studying how various
perturbations to a system or circuit V̂ transform into qubit–qubit or qubit–photon couplings.
In essence, this is what we do in Sections 6.5.1 and 6.5.2 when introducing the dipolar
magnetic couplings. These interactions were treated as diagonal perturbations V̂d that took
place in the qubit space. However, this method allows us to consider the effect of the
interaction via the connection of the qubit eigenspace to higher excited states, or to take
the already developed interaction and understand how it can be approximated by other,
much simpler models – as we did in Section 6.5.3 and (6.67). Finally, perturbation the-
ory also provides us with effective interactions in limits, such as the dispersive qubit–
cavity coupling from Section 7.3.7, where the interactions are only mediated by excited
states and V̂d = 0.
However, this is a “fake” frame of reference. If at some point we need to make predictions
about the observables and properties of the state |, we will need to undo the Schrieffer–
Wolff transformation:
|lab = e−g Ŝ | n (1 − g Ŝ0 ) |n,α + O(g 2 ). (A.40)
In doing so, we see that the application or Ŝ or Ŝ0 deforms the basis in which we expanded
our state, from a superposition of the original bare states of the qubit or cavity, to contain
some hybridization with excitations |m,β injected by the Ŝ0 operator. The actual proba-
bility of those excitations grows as O(g 2 ), but it may be something to consider, especially
when doing high-precision experiments, such as quantum gates and quantum computations.
1 The term completely positive is a technicality meaning that when we compose our Hilbert space with another
one, the extended operator remains positive. In other words, ε ⊗ 1B is positive for any dimension of the
identity 1B .
277
The matrices {K̂i , K̂i† } are called Kraus operators. The number of Kraus operators in a
decomposition can grow quadratically with the Hilbert space dimension. However, many
transformations require very few terms. These are then efficient representations of noise,
errors, and prototypical decohererence models – such as the ones in Problem 8.6 – that can
be experimentally calibrated and used to engineer error-correcting methods.
In the study of superconducting circuits, we are not so much concerned with specific
instances of a positive map. Our interest is more focused on the non-unitary evolution of
density matrices, given by differentiable maps ρ(t) = ε(ρ(0),t). Here the Kraus decompo-
sition is not so useful, as having a differential equation for ρ(t). Lindblad (1976) studied
such master equations and found that they are linear ordinary differential equations with a
common structure:
−1
d 2
i 1
∂t ρ = Lρ = − [Ĥ,ρ] + lˆi ρ lˆi† − lˆi† lˆi ρ − ρ lˆi† lˆi . (B.2)
h̄ 2
i=1 i
The generator of the evolution L is the Lindblad superoperator. It separates into a Hamil-
tonian term −i[Ĥ,ρ]/h̄, and up to d 2 − 1 decoherence terms, where d is the dimension of
our system. Both the Hermitian operator Ĥ and the dissipators {lˆi , lˆi† } can depend on time,
but we will focus on the Markovian dynamics with constant dissipators. This is a relevant
situation that appears when the environment is very large and “forgets” very quickly the
history of the system with which it interacts. We will see in Section B.2 how in such cases
we can derive master equations, with very few Lindblad terms, that are both simple and intu-
√
itively appealing. Examples include Pauli operators lˆi ∼ γ σ − modeling the spontaneous
√
relaxation of a qubit (see Section 6.1.6), Fock operators lˆi ∼ κ â that model the cooling
of a microwave resonator in a zero-temperature environment (see Section 5.5.4), etc.
2 The only circumstance where we would accept loss of trace would be when we want to model experiments
with postselection, where some fraction of outcomes are discarded. Then we could accept a decrease
in the norm, interpreting the trace of the density matrix as the total success probability of postselection.
differential equation that becomes local in time under a Markovian approximation – that the
environment forgets the history of our subsystem. With a slight massaging, we will end up
with a simple Lindblad equation that applies to microwave resonators and superconducting
qubits. But we will also offer prescriptions to study more complicated problems, such as
composite, driven, or highly nonlinear quantum systems.
i h̄Ẇ (t) = ei Ĥ0 t/h̄ Ĥ1 e−i Ĥ0 t/h̄ W (t) =: Ĥ1 (t)W (t). (B.4)
We can apply this interaction picture also to density matrices, splitting ρ(t) =
exp(−i Ĥ0 t/h̄)ρI (t) exp(i Ĥ0 t/h̄). The interaction picture ρ̃I satisfies a master equation
generated by the perturbations dtd ρI (t) = −i[Ĥ1 (t)/h̄,ρI (t)]. Usually, the dynamics gen-
erated by Ĥ1 (t) is much slower than the unitary evolution induced by exp(−i Ĥ0 t/h̄). This
justifies a Dyson series for the solution of the master equation, which we truncate up to
second order:
i t+δt
ρI (t + δt) ρ(t) − Ĥ1 (t ),ρ(t) dt (B.5)
h̄ t
t+δt t
1
− 2 Ĥ1 (t ), Ĥ1 (t ),ρ(t) dt dt .
h̄ t t
The next approximation is to neglect the feedback of the system onto the environment.
We impose that expectation values over the bath and system are uncorrelated, and that the
former are approximately constant.3 For an off-diagonal coupling4 TrB {Ĥ1 (t )ρ̃(t)} = 0,
t+δt t
1
ρ̃S (t + δt) ρ̃S (t) − 2 TrB Ĥ1 (t ), Ĥ1 (t ), ρ̃S (t) ⊗ ρ̃B dt dt .
h̄ t t
To continue further, we need to impose some structure in Ĥ1 . The following section
introduces a coupling that is typical of linear systems.
3 Traditionally, it has been said that the system and the environment adopt a product state structure where the
bath does not change, ρ̃(t) = ρ̃S (t) ⊗ ρ̃B , but this is not strictly needed.
4 We will see that this is not really a restriction, as typically Ĥ ∼ O ⊗ O , where the system and bath
1 S B
operators OS and OB have zero diagonal elements.
When we insert this expression into the formula (B.6), we obtain four combinations of
{Â, † } and {b̂k , b̂k† } acting on ρ̃S :
t+δt
- .
ρ̃S = dt IB † B † L + IB † B L† + IBB † L†  + IBB L† † ρ̃S (t),
t
with LX̂Ŷ ρ̃ = 2X̂ ρ̃ Ŷ − Ŷ X̂ ρ̃ − ρ̃ Ŷ X̂. (B.8)
The prefactors are four integrals of expectation values over the bath:
t+δt
!
IBB = gp gp b̂p b̂q e−i(ωp −)t−i(ωq −)t = IB∗ † B † (B.9)
pq t
t+δt !
IBB † = gp gp∗ b̂p b̂q† e−i(ωp −)t+i(ωq −)t ,
pq t
t+δt !
IB † B = gp∗ gp b̂p† b̂q e+i(ωp −)t−i(ωq −)t .
pq t
For a large enough bath, these integrals and the superoperators are approximately constant.
The equation can be summarized as the action of one superoperator ρ̃S (t + δt) = ρ̃S (t) +
δt × Lρ̃S (t), which in the limit δt → 0 generates a Lindblad equation:
d
ρ̃S = Lρ̃S . (B.10)
dt
Following Walls and Milburn (1994), the Markov approximation enters by assuming that J
and n̄ are smooth functions, almost constant over a broad range of frequencies. As in
Section B.3.1, the function F (t − t ) vanishes quickly beyond a short memory time τ δt.
We can thus take the upper integration limits to ±∞, obtaining a distribution
∞
±iτ 1
dτ e = π δ() ± iPV , (B.13)
0
where τ = t − t and = ω − . Using this distribution, we conclude
γ γ
IB † B = n̄() + iδ , and IBB † = [n̄() + 1] + iδ, (B.14)
2 2
with the decay rate γ = J QO () and two different Lamb shifts δ and δ that are usually
neglected. Inserting these values in (B.8) gives the usual quantum optics textbook master
equation:
d γ
ρ̃S = [n̄() + 1] 2Âρ † − † Âρ − ρ †  (B.15)
dt 2
γ
+ n̄() 2† ρ  − † ρ − ρ † ,
2
with the cooling and heating Lindblad operators. At this point, it is interesting to remark that
this equation can be brought back to the Schrödinger image, to compute the reduced density
matrix of the system in the laboratory frame, ρS (t) = trB ρ(t) = e−i Ĥ0 t/h̄ ρ̃S (t)ei Ĥ0 t/h̄ . In a
compact form,
d i γ γ
ρS = − [Ĥ0,ρS ] + [n̄() + 1] L† ρS + n̄() L†  ρS . (B.16)
dt h̄ 2 2
This perturbation contains three uncorrelated terms Ĥ1 = Ĥ1a + Ĥ1b + Ĥ1c , such that
tr(H1x H1y ρB ) = 0 if x = y. The Dyson series for ρ̃S has three independent contributions
that originate three Lindblad operators in the master equation dtd ρ̃S = (La + Lb + Lc )ρ̃S ,
one of which, La , will be a Hamiltonian contribution. A long and tedious but technically
simple process confirms this idea. It results in a model (7.34) with the original JC Hamil-
tonian, accompanied by dissipation terms for the cavity and the qubit. Note that the decay
QO QO
rates κ = Jb (ω) and γ = Jc () are evaluated in the spectral functions of the respective
independent environments. If the cavity and the qubit share the environment, the derivation
of the master equation will reveal interesting cross terms, Lâσ + ,Lâ † σ − , but these are out of
the scope of this book and have not been measured in experiments.
where h̄nm = En − Em > 0 are the positive frequency gaps. The analysis of Ĥ1 leads to a
small explosion of terms in (B.8), but they all support the same approximations. The result
is a generalized master equation:
↓
mn ↑
d i mn
ρS = − [Ĥ0,ρS ] + Lnm ρS + Lmn ρS . (B.19)
dt h̄ m,n
2 2
The bath equations are solved implicitely, using initial conditions in the faraway past
t0 → −∞ or in the faraway future tf → +∞. The first type of solutions reads
t
b̂k (t) = e−iωk (t−t0 ) b̂k (t0 ) − igk e−iωk (t−τ ) â(τ )dτ . (B.23)
t0
It contains the field leaking from the cavity â(t), together with the field that was injected
long ago into the line {bk (t0 )} and that is now reaching the cavity. Using this solution, we
obtain an integrodifferential equation for â(t) (5.61) that depends on (i) the bare properties
of the cavity, ω; (ii) the noise operator for the incoming field,
ξ̂ (t) = gk∗ e−iωk (t−t0 ) b̂k (t0 ); (B.24)
k
and (iii) the memory function of the bath K(t − τ ). The memory function is the Fourier
transform of the spectral function J QO (5.62). This function is broad and changes slowly
around the frequencies of interest. This produces a Fourier transform K(t) that concen-
trates around t = 0 and becomes zero past the memory time of the bath.
Let us illustrate this argument using a superconducting transmission line. The line has
a spectral function that is approximately Ohmic (i.e., linear) around the cavity resonance
J QO (ω̄) ∝ ω̄. In order to avoid ultraviolet divergencies, we regularize this function with an
exponential high-energy cutoff ωc constraining the range of frequencies that pass through
the guide. Apart from mathematical convenience, this is motivated by the filters on the
line and ultimately by the superconducting gap. Using this cutoff, we compute a memory
function:
αωc2
J QO (ω̄) = π α ω̄e−ω̄/ωc ⇒ K(t) = , (B.25)
(itωc + 1)2
which decays on a time of order tmemory 1/ωc . The memory time is the shortest timescale
in the problem – shorter than 2π /ω or than the time for the cavity to emit all its energy.
We talk about the Markovian regime as a limit in which the bath immediately loses all
memory of its interaction with the system. In this limit, the Markovian approximation is to
take the system operators as approximately constant during tmemory, replacing the memory
function with a Dirac delta K(t) ∼ δ(t).
Note how the oscillator feeds only from a narrow band κ around the resonance ω . This band
is so small that we can assume that the coupling strength gk , the spectral function J QO (ω),
and the density of states are more or less constant. Applying this idea to the study of the
input operators gives
† 2 −iωk (t−t )
[ξ̂in (t), ξ̂in (t )] = |gk | e = dωJ QO (ω)e−iω(t−t ) (B.33)
k
J QO (ω ) dωe−iω(t−t ) = κδ(t − t ). (B.34)
√
We replace the input operators with new ones b̂in (t) = ξ̂in / κ that satisfy causal com-
mutation relations, producing the Langevin equation (5.64). Note that the cavity opera-
tor â(t) is dimensionless, but κ b̂in † b̂in has dimensions of photon flux [time]−1 . Hence
[b̂in ] = [time]−1/2 .
We can repeat the whole preceding study, but using initial conditions that are far away
in the future, tf → +∞. The equation is slightly different:
d √
â = (−iω + κ/2)â − i κ b̂out, (B.35)
dt
and now involves the field reflected by the cavity b̂out . The problem with this equation is that
we do not really know the state of the environment in the future, b̂k (+∞), so that we cannot
compute b̂out . Fortunately, we can guess b̂out indirectly, realizing that both boundary condi-
tions, b̂out and b̂in, must give the same value for the field experienced by the cavity (B.23):
t
∗ √ in
gk bk (t) = κ b̂ (t) − i K(t − τ )â(τ )dτ (B.36)
k −∞
+∞
√
= κ b̂out (t) + i K(t − τ )â(τ )dτ . (B.37)
0
Combining both equations and using R K(t − τ )â(τ )dτ κ â(t), we arrive at the input–
output relation:
√
b̂out (t) = b̂in (t) − i κa(t). (B.38)
Let us close this section with some remarks. The first one is to point out that the input–
output relations vary with how we write the coupling Hamiltonian. If instead of (5.60) we
use h̄i(gk abk† −gk∗ bk a † ), this amounts to a phase shift â → i â that suppresses some factors:
d √
â(t) = (−iω − κ/2)â(t) − κ b̂in (t), (B.39)
dt
√
b̂out (t) = b̂in (t) + κ â(t). (B.40)
Another important variation is when we want to consider multiple input and output chan-
nels. This happens in transmission experiments, when the LC resonator is inserted in the
middle of a line, or when it is connected to multiple lines. In that case, the input–output
equations read
√
d
â(t) = −iω − κ /2 â(t) − i
r
κ r b̂in (t), (B.41)
dt r n
√
b̂rout (t) = b̂rin (t) − i κ r â(t). (B.42)
The index r = 1,2, . . . R runs over all input–output channels, which may have very different
decay rates κ r .
B.3.4 Spectroscopy
Throughout this book, we offer many examples of spectroscopy: experiments where we
illuminate a quantum system with microwaves and study the radiation that it absorbs, emits,
and scatters. Such experiments fit naturally in the input–output framework that we just
developed, provided it is extended to treat more general setups, such as qubits, cavity-QED,
and other composite and nonlinear systems.
The generalization of input–output theory parallels Section B.2.5. We consider an arbi-
trary quantum system with discrete eigenstates, coupled to R input–output channels:
R
†
Ĥ1 = Armn gr,k |En Em | b̂r,k + H.c. (B.43)
r=1 m=n k
Applying the same treatment and approximations as before, we arrive at a series of input–
output relations:
b̂rout (t) = b̂rin (t) − i γmn
r σ̂ ,
mn (B.44)
with system ladder operators σ̂mn (t = 0) = |Em En | that feed the different lines at
different rates γmn r ∝ |Ar |2 J QO ( ).
mn nm
Studying the evolution of the system operators σ̂mn is rather complicated, and it is rarely
done in the Heisenberg picture. Most often, we are just interested in the expectation val-
ues and correlations of the output fields, b̂rout . In this case, we can get away with any
method to estimate quantities such as σ̂mn (t) . The most popular method is to calculate the
coherences using a master equation such as (B.19), adapted to reproduce the conditions of
a spectroscopy protocol.
The simulations presume that the waveguides are close to zero temperature and host some
input fields b̂rin (t) ∼ r (t). The Hamiltonian part of the master equation accommodates
the coherent driving with which we illuminate our system. The losses are extended to add up
all spectroscopy channels and the intrinsic decoherence of our sample. We typically neglect
heating. In other words,
Ĥ0 = Em |Em Em | + γmnr (t)∗ |E E | + H.c., (B.45)
r m n
m r,m,n
↓ ↑
mn = int
γmn + r
γmn and nm 0.
r
The Lorentzians are centered on the different transition frequencies mn . The weight and
the linewidth of these resonances give information on the decoherence rates and on the
transition strengths, mn ∝ |Amn |2 . This is a lot of information that can be used to con-
firm and fit our microscopic models, such as the strong coupling and ultrastrong coupling
Hamiltonians from our circuit-QED models (Section 7.4.1).
Abdo, B., Suchoi, O., Segev, E., et al. 2009. Intermodulation and parametric amplification in
a superconducting stripline resonator integrated with a dc-SQUID. EPL (Europhysics
Letters), 85(6), 68001.
Adesso, Gerardo, Ragy, Sammy, and Lee, Antony R. 2014. Continuous variable quantum
information: Gaussian states and beyond. Open Systems & Information Dynamics,
21(01–02), 1440001.
Aharonov, Dorit, van Dam, Wim, Kempe, Julia, Landau, Zeph, Lloyd, Seth, and Regev,
Oded. 2004. Adiabatic quantum computation is equivalent to standard quantum
computation. SIAM Review, 50(4), 755–787.
Albash, Tameem, and Lidar, Daniel A. 2018. Demonstration of a scaling advantage for a
quantum annealer over simulated annealing. Physical Review X 8, 031016 (2018).
Albash, Tameem, Vinci, Walter, Mishra, Anurag, Warburton, Paul A., and Lidar, Daniel A.
2015. Consistency tests of classical and quantum models for a quantum annealer.
Physical Review A 91, 042314 (2015).
Anderson, P. W., and Rowell, J. M. 1963. Probable observation of the Josephson supercon-
ducting tunneling effect. Physical Review Letters, 10(6), 230–232.
Arute, Frank, Arya, Kunal, Babbush, Ryan, et al. 2019. Quantum supremacy using a
programmable superconducting processor. Nature, 574(7779), 505–510.
Arute, Frank, Arya, Kunal, Babbush, Ryan, et al. 2020. Hartree–Fock on a superconducting
qubit quantum computer. Science, 369(6507), 1084–1089.
Astafiev, O., Zagoskin, A. M., Abdumalikov, A. A., et al. 2010. Resonance fluorescence of
a single artificial atom. Science (New York, N.Y.), 327(5967), 840–843.
Ballentine, L. E. 1970. The statistical interpretation of quantum mechanics. Reviews of
Modern Physics, 42(4), 358–381.
Ballentine, Leslie E. 1998. Quantum Mechanics. World Scientific.
Bapst, V., Foini, L., Krzakala, F., Semerjian, G., and Zamponi, F. 2013. The quantum
adiabatic algorithm applied to random optimization problems: the quantum spin glass
perspective. Physics Reports, 523(3), 127–205.
Barahona, F. 1982. On the computational complexity of Ising spin glass models. Journal of
Physics A: Mathematical and General, 15(10), 3241–3253.
Bardeen, J., Cooper, L. N., and Schrieffer, J. R. 1957a. Microscopic theory of superconduc-
tivity. Physical Review, 106(1), 162–164.
Bardeen, J., Cooper, L. N., and Schrieffer, J. R. 1957b. Theory of superconductivity.
Physical Review, 108(5), 1175–1204.
287
Barends, R., Kelly, J., Megrant, A., et al. 2014. Superconducting quantum circuits at the
surface code threshold for fault tolerance. Nature, 508(7497), 500–503.
Baumgratz, T., Cramer, M., and Plenio, M. B. 2014. Quantifying coherence. Physical
Review Letters, 113(14), 140401.
Benioff, Paul. 1980. The computer as a physical system: a microscopic quantum mechanical
Hamiltonian model of computers as represented by Turing machines. Journal of
Statistical Physics, 22(5), 563–591.
Bennett, Douglas A., Longobardi, Luigi, Patel, Vijay, Chen, Wei, and Lukens, James E.
2007. rf-SQUID qubit readout using a fast flux pulse. Superconductor Science and
Technology, 20(11), S445–S449.
Berkley, A. J., Johnson, M. W., Bunyk, P., et al. 2010. A scalable readout system for
a superconducting adiabatic quantum optimization system. Superconductor Science
and Technology, 23(10), 105014.
Berkley, A. J., Przybysz, A. J., Lanting, T., et al. 2013. Tunneling spectroscopy using a
probe qubit. Physical Review B, 87(2), 020502.
Bialczak, R. C., Ansmann, M., Hofheinz, M., et al. 2010. Quantum process tomography of a
universal entangling gate implemented with Josephson phase qubits. Nature Physics,
6(6), 409–413.
Biamonte, Jacob D., and Love, Peter J. 2008. Realizable Hamiltonians for universal
adiabatic quantum computers. Physical Review A, 78(1), 012352.
Bishop, Lev S., Bravyi, Sergey, Gambetta, Jay M., and Smolin, John. 2017. Quantum
Volume. Technical Report.
Blais, Alexandre, Huang, Ren-Shou, Wallraff, Andreas, Girvin, S., and Schoelkopf, R.
2004. Cavity quantum electrodynamics for superconducting electrical circuits: an
architecture for quantum computation. Physical Review A, 69(6), 062320.
Boixo, Sergio, Albash, Tameem, Spedalieri, Federico M., Chancellor, Nicholas, and Lidar,
Daniel A. 2013. Experimental signature of programmable quantum annealing. Nature
Communications, 4(1), 2067.
Boixo, Sergio, Rønnow, Troels F., Isa kov, Sergei V., et al. 2014. Evidence for quantum
annealing with more than one hundred qubits. Nature Physics, 10(3), 218–224.
Boixo, Sergio, Smelyanskiy, Vadim N., Shabani, Alireza, et al. 2016. Computational
multiqubit tunnelling in programmable quantum annealers. Nature Communications,
7(Jan), 10327.
Bombin, H., and Martin-Delgado, M. A. 2006. Topological Quantum Distillation. Physical
Review Letters, 97(Oct), 180501.
Bouchiat, V., Vion, D., Joyez, P., Esteve, D., and Devoret, M. H. 1998. Quantum coherence
with a single Cooper pair. Physica Scripta, T76(1), 165–170.
Bourassa, J., Gambetta, J., Abdumalikov, A., Astafiev, O., Nakamura, Y., and Blais, A. 2009.
Ultrastrong coupling regime of cavity QED with phase-biased flux qubits. Physical
Review A, 80(3), 032109.
Bravyi, S. B., and Kitaev, A. Yu. 1998. Quantum codes on a lattice with boundary. e-print
arXiv:quant-ph-9811052.
Bravyi, Sergey, DiVincenzo, David P., and Loss, Daniel. 2011. Schrieffer–Wolff transfor-
mation for quantum many-body systems. Annals of Physics, 326(10), 2793–2826.
Brecht, T., Chu, Y., Axline, C., et al. 2017. Micromachined integrated quantum circuit
containing a superconducting qubit. Physics Review Applied, 7(Apr), 044018.
Brecht, Teresa, Pfaff, Wolfgang, Wang, et al. 2016. Multilayer microwave integrated
quantum circuits for scalable quantum computing. npj Quantum Information, 2(1),
16002.
Brink, Markus, Corcoles-Gonzalez, Antonio, Gambetta, Jay M., Rosenblatt, Sami, and
Solgun, Firat. 2019 (May 28). Low loss architecture for superconducting qubit
circuits. US Patent 10,305,015.
Brown, K. R., Wilson, A. C., Colombe, Y., et al. 2011. Single-qubit-gate error below 10−4
in a trapped ion. Physics Review A, 84(Sep), 030303.
Cai, Jun, Macready, William G., and Roy, Aidan. 2014. A practical heuristic for finding
graph minors. e-print arXiv:1406.2741.
Calderbank, A. R., and Shor, Peter W. 1996. Good quantum error-correcting codes exist.
Physical Review A, 54(Aug), 1098–1105.
Castellanos-Beltran, M. A., Irwin, K. D., Vale, L. R., Hilton, G. C., and Lehnert, K. W. 2009.
Bandwidth and dynamic range of a widely tunable Josephson parametric amplifier.
IEEE Transactions on Applied Superconductivity, 19(3), 944–947.
Chancellor, N., Zohren, S., and Warburton, P. A. 2017. Circuit design for multi-body
interactions in superconducting quantum annealing systems with applications to a
scalable architecture. npj Quantum Information, 3(1), 21.
Chen, Yu, Neill, C., Roushan, P., et al. 2014. Qubit architecture with high coherence and
fast tunable coupling. Physical Review Letters, 113(Nov), 220502.
Chen, Zijun, Satzinger, Kevin J., Atalaya, Juan, et al. 2021. Exponential suppression of bit
or phase errors with cyclic error correction. Nature, 595(7867), 383–387.
Chiorescu, I., Bertet, P., Semba, K., Nakamura, Y., Harmans, C. J. P. M., and Mooij, J. E.
2004. Coherent dynamics of a flux qubit coupled to a harmonic oscillator. Nature,
431(7005), 159–162.
Chow, J. M., Gambetta, J. M., Tornberg, L., et al. 2009. Randomized benchmarking and
process tomography for gate errors in a solid-state qubit. Physical Review Letters,
102(9), 090502.
Chow, Jerry M., Córcoles, A. D., Gambetta, Jay M., et al. 2011. Simple all-microwave
entangling gate for fixed-frequency superconducting qubits. Physical Review Letters,
107(8), 080502.
Chow, Jerry M., Gambetta, Jay M., Córcoles, A. D., et al. 2012. Universal quantum gate set
approaching fault-tolerant thresholds with superconducting qubits. Physical Review
Letters, 109(6), 060501.
Cirac, J. I., and Zoller, P. 1995. Quantum computations with cold trapped ions. Physical
Review Letters, 74(20), 4091–4094.
Cirac, J. I., Ekert, A. K., Huelga, S. F., and Macchiavello, C. 1999. Distributed quantum
computation over noisy channels. Physical Review A, 59(6), 4249–4254.
Clarke, John, and Braginski, Alex I. (eds). 2004. The SQUID Handbook: Fundamentals and
Technology of SQUIDs and SQUID Systems, Volume I. Wiley-VCH.
Clarke, John, and Wilhelm, Frank K. 2008. Superconducting quantum bits. Nature,
453(7198), 1031–1042.
Cohen-Tannoudji, Claude, Diu, Bernard, and Laloë, Franck. 1977. Quantum Mechanics.
Wiley.
Consani, Gioele, and Warburton, Paul A. 2020. Effective Hamiltonians for interacting super-
conducting qubits: local basis reduction and the Schrieffer–Wolff transformation. New
Journal of Physics, 22(5), 053040.
Cooper, Leon N. 1956. Bound electron pairs in a degenerate Fermi gas. Physical Review,
104(4), 1189–1190.
Cory, D. G., Fahmy, A. F., and Havel, T. F. 1997. Ensemble quantum computing by NMR
spectroscopy. Proceedings of the National Academy of Sciences of the United States
of America, 94(5), 1634–1639.
Cross, Andrew W., Bishop, Lev S., Sheldon, Sarah, Nation, Paul D., and Gambetta, Jay M.
2019. Validating quantum computers using randomized model circuits. Physical
Review A, 100(Sep), 032328.
Cullen, A. L. 1960. Theory of the travelling-wave parametric amplifier. Proceedings of the
IEE Part B: Electronic and Communication Engineering, 107(32), 101.
Deaver, Bascom S., and Fairbank, William M. 1961. Experimental evidence for quantized
flux in superconducting cylinders. Physical Review Letters, 7(2), 43–46.
Denchev, Vasil S., Boixo, Sergio, Isakov, Sergei V., et al. 2016. What is the computational
value of finite-range tunneling? Physical Review X, 6(3), 031015.
Deutsch, David Elieser. 1985. Quantum theory, the Church–Turing principle and the
universal quantum computer. Proceedings of the Royal Society A: Mathematical,
Physical and Engineering Sciences, 400(1818), 97–117.
Deutsch, David Elieser. 1989. Quantum computational networks. Proceedings of the Royal
Society of London. A. Mathematical and Physical Sciences, 425(1868), 73–90.
Devoret, M. H. 1995. Quantum fluctuations in electrical circuits. Les Houches, Session
LXIII, 351–386.
DiCarlo, L., Chow, J. M., Gambetta, J. M., et al. 2009. Demonstration of two-qubit
algorithms with a superconducting quantum processor. Nature, 460(7252), 240–244.
DiCarlo, L., Reed, M. D., Sun, L., et al. 2010. Preparation and measurement of three-qubit
entanglement in a superconducting circuit. Nature, 467(7315), 574–578.
Dickson, N. G., Johnson, M. W., Amin, M. H., et al. 2013. Thermally assisted quantum
annealing of a 16-qubit problem. Nature Communications, 4(1), 1903.
Dieks, D. 1982. Communication by EPR devices. Physics Letters A, 92(6), 271–272.
DiVincenzo, D. P. 1995. Quantum computation. Science, 270(5234), 255–261.
DiVincenzo, David P. 2000. The physical implementation of quantum computation.
Fortschritte der Physik, 48(9-11), 771–783.
Dowling, Jonathan P., and Milburn, Gerard J. 2003. Quantum technology: the second
quantum revolution. Philosophical Transactions of the Royal Society of London.
Series A: Mathematical, Physical and Engineering Sciences, 361(1809), 1655–1674.
Egger, D. J., Werninghaus, M., Ganzhorn, M., et al. 2018. Pulsed reset protocol for
fixed-frequency superconducting qubits. Physical Review Applied, 10(Oct), 044030.
Eichler, C., Bozyigit, D., and Wallraff, A. 2012. Characterizing quantum microwave
radiation and its entanglement with superconducting qubits using linear detectors.
Physical Review A, 86(3), 032106.
Eichler, C., Bozyigit, D., Lang, C., Steffen, L., Fink, J., and Wallraff, A. 2011. Experimental
state tomography of itinerant single microwave photons. Physical Review Letters,
106(22), 220503.
Farhi, Edward, Goldstone, Jeffrey, Gutmann, Sam, and Sipser, Michael. 2000. Quantum
computation by adiabatic evolution. arXiv, quant-ph:0(jan).
Feynman, Richard P. 1982. Simulating physics with computers. International Journal of
Theoretical Physics, 21(6-7), 467–488.
Forn-Díaz, P., Lisenfeld, J., Marcos, D., et al. 2010. Observation of the Bloch–Siegert
shift in a qubit-oscillator system in the ultrastrong coupling regime. Physical Review
Letters, 105(23), 237001.
Forn-Díaz, P., García-Ripoll, J. J., Peropadre, B., et al. 2016. Ultrastrong coupling of a single
artificial atom to an electromagnetic continuum in the nonperturbative regime. Nature
Physics, 13(1), 39–43.
Forn-Díaz, P., Warren, C. W., Chang, C. W. S., Vadiraj, A. M., and Wilson, C. M. 2017.
On-demand microwave generator of shaped single photons. Physical Review Applied,
8(5), 054015.
Fowler, Austin G., Mariantoni, Matteo, Martinis, John M., and Cleland, Andrew N. 2012.
Surface codes: towards practical large-scale quantum computation. Physical Review
A, 86(Sep), 032324.
Fowler, Austin G., Stephens, Ashley M., and Groszkowski, Peter. 2009. High-threshold
universal quantum computation on the surface code. Physical Review A, 80(Nov),
052312.
Gaebler, J. P., Meier, A. M., Tan, T. R., et al. 2012. Randomized benchmarking of multiqubit
gates. Physical Review Letters, 108(26), 260503.
Galindo, Alberto, and Pascual, Pedro. 1991. Time-dependent perturbation theory. Pages
161–199 of: Quantum Mechanics II. Springer Berlin Heidelberg.
García-Ripoll, J. J., Peropadre, B., and De Liberato, S. 2015. Light–matter decoupling and
A2 term detection in superconducting circuits. Scientific Reports, 5(1), 16055.
García-Ripoll, J. J., Ruiz-Chamorro, A., and Torrontegui, E. 2020. Quantum control of
frequency-tunable transmon superconducting qubits. Physical Review Applied, 14(4),
044035.
Gardiner, C. W., and Zoller, P. 2004. Quantum Noise. 3rd ed. Berlin, Heidelberg: Springer-
Verlag Berlin Heidelberg.
Gershenfeld, Neil A., and Chuang, Isaac L. 1997. Bulk spin-resonance quantum computa-
tion. Science (New York, N.Y.), 275(5298), 350–356.
Gilchrist, Alexei, Langford, Nathan K., and Nielsen, Michael A. 2005. Distance measures
to compare real and ideal quantum processes. Physical Review A, 71(6), 062310.
Gor’kov, Lev P. 1959. Microscopic derivation of the Ginzburg–Landau equations in the
theory of superconductivity. JETP 36(9), 1364.
Gottesman, Daniel. 1998. Theory of fault-tolerant quantum computation. Physical Review
A, 57(Jan), 127–137.
Haroche, Serge. 2013. Nobel lecture: controlling photons in a box and exploring the
quantum to classical boundary. Reviews of Modern Physics, 85(3), 1083–1102.
Harris, R., Berkley, A. J., Johnson, M. W., et al. 2007. Sign- and magnitude-tunable coupler
for superconducting flux qubits. Physical Review Letters, 98(17), 177001.
Harris, R., Johansson, J., Berkley, A. J., et al. 2010a. Experimental demonstration of a robust
and scalable flux qubit. Physical Review B, 81(13), 134510.
Harris, R., Johnson, M. W., Lanting, T., et al. 2010b. Experimental investigation of an
eight-qubit unit cell in a superconducting optimization processor. Physical Review
B, 82(2), 024511.
Harris, R., Lanting, T., Berkley, A. J., et al. 2009. Compound Josephson-junction coupler
for flux qubits with minimal crosstalk. Physical Review B, 80(5), 052506.
Harris, R., Sato, Y., Berkley, A. J., et al. 2018. Phase transitions in a programmable quantum
spin glass simulator. Science, 361(6398), 162–165.
Hauke, Philipp, Katzgraber, Helmut G, Lechner, Wolfgang, Nishimori, Hidetoshi, and
Oliver, William D. 2020. Perspectives of quantum annealing: methods and implemen-
tations. Reports on Progress in Physics, 83(5), 054401.
Heisenberg, W. 1925. Über quantentheoretische Umdeutung kinematischer und mechanis-
cher Beziehungen. Zeitschrift für Physik, 33(1), 879–893.
Hime, T., Reichardt, P. A., Plourde, B. L. T., et al. 2006. Solid-state qubits with current–
controlled coupling. Science (New York, N.Y.), 314(5804), 1427–1429.
Hita-Pérez, María, Jaumá, Gabriel, Pino, Manuel, and García-Ripoll, Juan José. 2022.
Ultrastrong capacitive coupling of flux qubits. Physical Review Applied 17, 014028.
Hofheinz, Max, Wang, H., Ansmann, M., et al. 2009. Synthesizing arbitrary quantum states
in a superconducting resonator. Nature, 459(7246), 546–549.
Hoi, Io-Chun, Kockum, Anton F., Palomaki, Tauno, et al. 2013. Giant Cross–Kerr effect
for propagating microwaves induced by an artificial atom. Physical Review Letters,
111(5), 053601.
Hoi, Io-Chun, Palomaki, Tauno, Lindkvist, Joel, Johansson, Göran, Delsing, Per, and
Wilson, C. M. 2012. Generation of nonclassical microwave states using an artificial
atom in 1d open space. Physical Review Letters, 108(26), 263601.
Hoi, Io-Chun, Wilson, C. M., Johansson, Göran, Palomaki, Tauno, Peropadre, Borja, and
Delsing, Per. 2011. Demonstration of a single-photon router in the microwave regime.
Physical Review Letters, 107(7), 073601.
Jansen, Sabine, Ruskai, Mary-Beth, and Seiler, Ruedi. 2007. Bounds for the adiabatic
approximation with applications to quantum computation. Journal of Mathematical
Physics, 48(10), 102111.
Jaynes, E. T., and Cummings, F. W. 1963. Comparison of quantum and semiclassical
radiation theories with application to the beam maser. Proceedings of the IEEE, 51(1),
89–109.
Jeffrey, Evan, Sank, Daniel, Mutus, J. Y., et al. 2014. Fast accurate state measurement with
superconducting qubits. Physical Review Letters, 112(19), 190504.
Johnson, M. W., Amin, M. H. S., Gildert, S., et al. 2011. Quantum annealing with
manufactured spins. Nature, 473(7346), 194–198.
Johnson, M. W., Bunyk, P., Maibaum, F., et al. 2010. A scalable control system for a
superconducting adiabatic quantum optimization processor. Superconductor Science
and Technology, 23(6), 065004.
Josephson, B. D. 1962. Possible new effects in superconductive tunnelling. Physics Letters,
1(7), 251–253.
Kadowaki, Tadashi, and Nishimori, Hidetoshi. 1998. Quantum annealing in the transverse
Ising model. Physical Review E, 58(5), 5355–5363.
Kandala, Abhinav, Temme, Kristan, Córcoles, Antonio D., Mezzacapo, Antonio, Chow,
Jerry M., and Gambetta, Jay M. 2019. Error mitigation extends the computational
reach of a noisy quantum processor. Nature, 567(7749), 491–495.
Kato, Tosio. 1950. On the adiabatic theorem of quantum mechanics. Journal of the Physical
Society of Japan, 5(6), 435–439.
Kelly, J., Barends, R., Fowler, A. G., et al. 2015. State preservation by repetitive error
detection in a superconducting quantum circuit. Nature, 519(7541), 66–69.
Kempe, Julia, Kitaev, Alexei, and Regev, Oded. 2006. The complexity of the local Hamil-
tonian problem. SIAM Journal on Computing, 35(5), 1070–1097.
King, Andrew D., Carrasquilla, Juan, Raymond, Jack, et al. 2018. Observation of topo-
logical phenomena in a programmable lattice of 1,800 qubits. Nature, 560(7719),
456–460.
Kitaev, A., Shen, A., and Vyalyi, M. 2002. Classical and Quantum Computation. Graduate
Studies in Mathematics. American Mathematical Society.
Knill, E., Leibfried, D., Reichle, R., et al. 2008. Randomized benchmarking of quantum
gates. Physical Review A, 77(1), 012307.
Koch, Jens, Yu, Terri, Gambetta, Jay, et al. 2007. Charge-insensitive qubit design derived
from the Cooper pair box. Physical Review A, 76(4), 042319.
Krinner, S., Lazar, S., Remm, A., et al. 2020. Benchmarking coherent errors in con-
trolled-phase gates due to spectator qubits. Physical Review Applied, 14(2), 024042.
Kurcz, Andreas, Bermudez, Alejandro, and García-Ripoll, Juan José. 2014. Hybrid quantum
magnetism in circuit QED: from spin-photon waves to many-body spectroscopy.
Physical Review Letters, 112(18), 180405.
Landau, L. 1932. Zur theorie der energieubertragung II. Physik. Z. Sowjet, 2, 46–50.
Lanting, T., Przybysz, A. J., Smirnov, A. Yu., et al. 2014. Entanglement in a quantum
annealing processor. Physical Review X, 4(2), 021041.
Laurat, Julien, Keller, Gaëlle, Oliveira-Huguenin, José Augusto, et al. 2005. Entanglement
of two-mode Gaussian states: characterization and experimental production and
manipulation. Journal of Optics B: Quantum and Semiclassical Optics, 7(12), S577–
S587.
Law, C. K., and Eberly, J. H. 1996. Arbitrary control of a quantum electromagnetic field.
Physical Review Letters, 76(7), 1055–1058.
Leggett, A., Chakravarty, S., Dorsey, A., Fisher, Matthew, Garg, Anupam, and Zwerger,
W. 1987. Dynamics of the dissipative two-state system. Reviews of Modern Physics,
59(1), 1–85.
Leibfried, D., Blatt, R., Monroe, C., and Wineland, D. 2003. Quantum dynamics of single
trapped ions. Reviews of Modern Physics, 75(1), 281–324.
Lidar, Daniel A., Rezakhani, Ali T., and Hamma, Alioscia. 2009. Adiabatic approximation
with exponential accuracy for many-body systems and quantum computation. Journal
of Mathematical Physics, 50(10), 102106.
Lindblad, G. 1976. On the generators of quantum dynamical semigroups. Communications
in Mathematical Physics, 48(2), 119–130.
London, F., London, H., and Lindemann, Frederick Alexander. 1935. The electromagnetic
equations of the supraconductor. Proceedings of the Royal Society of London. Series
A - Mathematical and Physical Sciences, 149(866), 71–88.
Loss, Daniel, and DiVincenzo, David P. 1998. Quantum computation with quantum dots.
Physical Review A, 57(1), 120–126.
Macklin, C., O’Brien, K., Hover, D., et al. 2015. A near-quantum-limited Josephson
traveling-wave parametric amplifier. Science (New York, N.Y.), 350(6258), 307–310.
Magesan, Easwar, Gambetta, J. M., and Emerson, Joseph. 2011. Scalable and robust
randomized benchmarking of quantum processes. Physical Review Letters, 106(18),
180504.
Magesan, Easwar, Gambetta, Jay M., Johnson, B. R., et al. 2012. Efficient measurement
of quantum gate error by interleaved randomized benchmarking. Physical Review
Letters, 109(8), 080505.
Magnard, P., Kurpiers, P., Royer, B., et al. 2018. Fast and unconditional all-microwave reset
of a superconducting qubit. Physical Review Letters, 121(Aug), 060502.
Magnard, P., Storz, S., Kurpiers, P., et al. 2020. Microwave quantum link between super-
conducting circuits housed in spatially separated cryogenic systems. Physical Review
Letters, 125(26), 260502.
Majer, J., Chow, J. M., Gambetta, J. M., et al. 2007. Coupling superconducting qubits via a
cavity bus. Nature, 449(7161), 443–447.
Makhlin, Yuriy, Schön, Gerd, and Shnirman, Alexander. 2001. Quantum-state engineering
with Josephson-junction devices. Reviews of Modern Physics, 73(2), 357–400.
Mallet, F., Castellanos-Beltran, M. A., Ku, H. S., et al. 2011. Quantum state tomography of
an itinerant squeezed microwave field. Physical Review Letters, 106(22), 220502.
Mandrá, Salvatore, and Katzgraber, Helmut G. 2018. A deceptive step towards quantum
speedup detection. Quantum Science and Technology, 3(4), 04LT01.
Mandrá, Salvatore, Zhu, Zheng, Wang, Wenlong, Perdomo-Ortiz, Alejandro, and Katz-
graber, Helmut G. 2016. Strengths and weaknesses of weak-strong cluster problems:
a detailed overview of state-of-the-art classical heuristics versus quantum approaches.
Physical Review A, 94(2), 022337.
McEwen, M., Kafri, D., Chen, Z., et al. 2021. Removing leakage-induced correlated errors
in superconducting quantum error correction. Nature Communications, 12(1), 1761.
Menzel, E. P., Deppe, F., Mariantoni, M., et al. 2010. Dual-path state reconstruction scheme
for propagating quantum microwaves and detector noise tomography. Physical Review
Letters, 105(10), 100401.
Menzel, E. P., Di Candia, R., Deppe, F., et al. 2012. Path entanglement of continuous-
variable quantum microwaves. Physical Review Letters, 109(25), 250502.
Merkel, Seth T., Gambetta, Jay M., Smolin, John A., et al. 2013. Self-consistent quantum
process tomography. Physical Review A, 87(Jun), 062119.
Mi, Xiao, Ippoliti, Matteo, Quintana, Chris, Greene, et al. 2022. Observation of time-
crystalline eigenstate order on a quantum processor. Nature 601, 531.
Mirrahimi, Mazyar, Leghtas, Zaki, Albert, et al. 2014. Dynamically protected cat-qubits:
a new paradigm for universal quantum computation. New Journal of Physics, 16(4),
045014.
Mlynek, J. A., Abdumalikov, A. A., Eichler, C., and Wallraff, A. 2014. Observation of
Dicke superradiance for two artificial atoms in a cavity with high decay rate. Nature
Communications, 5(1), 5186.
Mooij, J. E., Orlando, T. P., Levitov, L., Tian, L., van der Wal, C. H., and Lloyd, S. 1999.
Josephson persistent-current qubit. Science, 285(5430), 1036–1039.
Nakamura, Y., Chen, C. D., and Tsai, J. S. 1997. Spectroscopy of energy-level splitting
between two macroscopic quantum states of charge coherently superposed by Joseph-
son coupling. Physical Review Letters, 79(12), 2328–2331.
Nakamura, Y., Pashkin, Yu. A., and Tsai, J. S. 1999. Coherent control of macroscopic
quantum states in a single-Cooper-pair box. Nature, 398(6730), 786–788.
Nataf, Pierre, and Ciuti, Cristiano. 2010. No-go theorem for superradiant quantum phase
transitions in cavity QED and counter-example in circuit QED. Nature Communica-
tions, 1(6), 1–6.
Neeley, Matthew, Ansmann, M., Bialczak, Radoslaw C., et al. 2008. Process tomography
of quantum memory in a Josephson-phase qubit coupled to a two-level state. Nature
Physics, 4(7), 523–526.
Neill, C., McCourt, T., Mi, X., et al. 2021. Accurately computing the electronic properties
of a quantum ring. Nature, 594(7864), 508–512.
Nielsen, M. A, and Chuang, I. L. 2011. Quantum Computation and Quantum Information.
Cambridge University Press.
Nielsen, Michael A. 2002. A simple formula for the average gate fidelity of a quantum
dynamical operation. Physics Letters A, 303(4), 249–252.
Niemczyk, T., Deppe, F., Huebl, H., et al. 2010. Circuit quantum electrodynamics in the
ultrastrong-coupling regime. Nature Physics, 6(10), 772–776.
Nigg, Simon E., Paik, Hanhee, Vlastakis, et al. 2012. Black-box superconducting circuit
quantization. Physical Review Letters, 108(24), 240502.
Niskanen, A O, Harrabi, K, Yoshihara, F, Nakamura, Y, Lloyd, S, and Tsai, J S. 2007.
Quantum coherent tunable coupling of superconducting qubits. Science (New York,
N.Y.), 316(5825), 723–726.
O’Brien, J. L., Pryde, G. J., Gilchrist, A., et al. 2004. Quantum process tomography of a
controlled-NOT gate. Physical Review Letters, 93(Aug), 080502.
O’Brien, William, Vahidpour, Mehrnoosh, Whyland, Jon Tyler, et al. 2017. Superconduct-
ing caps for quantum integrated circuits. e-print arXiv:1708.02219.
Ofek, Nissim, Petrenko, Andrei, Heeres, Reinier, et al. 2016. Extending the lifetime of a
quantum bit with error correction in superconducting circuits. Nature, 536(7617),
441–445.
Olivares, S. 2012. Quantum optics in the phase space. European Physical Journal Special
Topics, 203(1), 3–24.
Olver, F. W. J., Lozier, D. W., Boisvert, R. F., and Clark, C. W. (eds). 2010. NIST Handbook
of Mathematical Functions. Cambridge University Press.
O’Malley, P. J. J., Kelly, J., Barends, R., et al. 2015. Qubit metrology of ultralow phase noise
using randomized benchmarking. Physical Review Applied, 3(Apr), 044009.
Orlando, Terry P. 1991. Foundations of Applied Superconductivity. Addison-Wesley Pub-
lishing Company.
Ortuño, M., Somoza, A. M., Vinokur, V. M., and Baturina, T. I. 2015. Electronic transport in
two-dimensional high dielectric constant nanosystems. Scientific Reports, 5(1), 9667.
Otterbach, J. S., Manenti, R., Alidoust, N., et al. 2017. Unsupervised machine learning on
a hybrid quantum computer. e-print arXiv:1712.05771.
Ozfidan, I., Deng, C., Smirnov, A.Y., et al. 2020. Demonstration of a nonstoquastic
Hamiltonian in coupled superconducting flux qubits. Physical Review Applied, 13(3),
034037.
Paauw, F., Fedorov, A., Harmans, C., and Mooij, J. 2009. Tuning the gap of a superconduct-
ing flux qubit. Physical Review Letters, 102(9), 090501.
Paik, Hanhee, Schuster, D. I., Bishop, Lev S., et al. 2011. Observation of high coherence in
Josephson junction qubits measured in a three-dimensional circuit QED architecture.
Physical Review Letters, 107(24), 240501.
Pashkin, Yu. A., Yamamoto, T., Astafiev, O., Nakamura, Y., Averin, D. V., and Tsai, J. S.
2003. Quantum oscillations in two coupled charge qubits. Nature, 421(6925), 823–
826.
Pechal, M., Huthmacher, L., Eichler, C., et al. 2014. Microwave-controlled generation of
shaped single photons in circuit quantum electrodynamics. Physical Review X, 4(4),
041010.
Pegg, D., and Barnett, S. 1989. Phase properties of the quantized single-mode electromag-
netic field. Physical Review A, 39(4), 1665–1675.
Peropadre, B., Forn-Díaz, P., Solano, E., and García-Ripoll, J. 2010. Switchable ultrastrong
coupling in circuit QED. Physical Review Letters, 105(2), 023601.
Peropadre, B., Zueco, D., Porras, D., and García-Ripoll, J. 2013. Nonequilibrium and
nonperturbative dynamics of ultrastrong coupling in open lines. Physical Review
Letters, 111(24), 243602.
Pino, Manuel, and García-Ripoll, Juan José. 2018. Quantum annealing in spin-boson model:
from a perturbative to an ultrastrong mediated coupling. New Journal of Physics,
20(11), 113027.
Pitaevskii, Lev, and Stringari, Sandro. 2016. Bose-Einstein Condensation and Superfluidity.
Oxford University Press.
Place, Alexander P. M., Rodgers, Lila V. H., Mundada, Pranav, et al. 2021. New material
platform for superconducting transmon qubits with coherence times exceeding 0.3
milliseconds. Nature Communications, 12(1), 1779.
Preskill, John. 2018. Quantum computing in the NISQ era and beyond. Quantum, 2(Aug.),
79.
Raimond, J. M., Brune, M., and Haroche, S. 2001. Manipulating quantum entanglement
with atoms and photons in a cavity. Reviews of Modern Physics, 73(3), 565–582.
Ramos, Tomás, and García-Ripoll, Juan José. 2018. Correlated dephasing noise in sin-
gle-photon scattering. New Journal of Physics, 20(10), 105007.
Reed, M. D., DiCarlo, L., Johnson, B. R., et al. 2010. High-fidelity readout in circuit
quantum electrodynamics using the Jaynes–cummings nonlinearity, 173601. Physical
Review Letters, 105(17).
Rigetti, Chad, and Devoret, Michel. 2010. Fully microwave-tunable universal gates in super-
conducting qubits with linear couplings and fixed transition frequencies. Physical
Review B, 81(13), 134507.
Rigetti, Chad, Gambetta, Jay M., Poletto, Stefano, et al. 2012. Superconducting qubit in
a waveguide cavity with a coherence time approaching 0.1 ms. Physical Review B,
86(10), 100506.
Rol, M. A., Battistel, F., Malinowski, F. K., et al. 2019. Fast, high-fidelity conditional-phase
gate exploiting leakage interference in weakly anharmonic superconducting qubits.
Physical Review Letters, 123(Sep), 120502.
Romero, G., García-Ripoll, J. J., and Solano, E. 2009a. Microwave photon detector in circuit
QED. Physical Review Letters, 102(17), 173602.
Romero, Guillermo, García-Ripoll, Juan José, and Solano, Enrique. 2009b. Photodetection
of propagating quantum microwaves in circuit QED. Physica Scripta, T137(T137),
014004.
Rosenblum, S., Reinhold, P., Mirrahimi, M., Jiang, Liang, Frunzio, L., and Schoelkopf, R. J.
2018. Fault-tolerant detection of a quantum error. Science, 361(6399), 266–270.
Roy, Ananda, and Devoret, Michel. 2016. Introduction to parametric amplification of
quantum signals with Josephson circuits. Comptes Rendus Physique, 17(7), 740–755.
Sánchez-Burillo, E., Martín-Moreno, L., García-Ripoll, J. J., and Zueco, D. 2016. Full
two-photon down-conversion of a single photon. Physical Review A, 94(5), 053814.
Sandberg, M., Wilson, C. M., Persson, F., et al. 2008. Tuning the field in a microwave
resonator faster than the photon lifetime. Applied Physics Letters, 92(20), 203501.
Schneider, Christian Markus Florian. 2014. On-Chip Superconducting Microwave Beam
Splitter. Ph.D. thesis, TU Munich.
Schrödinger, E. 1926. Quantisierung als Eigenwertproblem. Annalen der Physik, 384(4),
361–376.
Schuster, D. I., Houck, A. A., Schreier, J. A., et al. 2007. Resolving photon number states
in a superconducting circuit. Nature, 445(7127), 515–518.
Schuster, D. I., Wallraff, A., Blais, A., et al. 2005. ac stark shift and dephasing of a
superconducting qubit strongly coupled to a cavity field. Physical Review Letters,
94(12), 123602.
Selby, Alex. 2014. Efficient subgraph-based sampling of Ising-type models with frustration.
e-print, sep, arXiv:1409.3934.
Sharafiev, Aleksei, Juan, Mathieu L., Gargiulo, Oscar, et al. 2021. Visualizing the emission
of a single photon with frequency and time resolved spectroscopy. Quantum, 5(Jun),
474.
Shi, Tao, Chang, Yue, and García-Ripoll, Juan José. 2018. Ultrastrong coupling few-photon
scattering theory. Physical Review Letters, 120(15), 153602.
Shin, Seung Woo, Smith, Graeme, Smolin, John A., and Vazirani, Umesh. 2014. How
“quantum” is the D-Wave machine? e-print arXiv:1712.05771.
Shnirman, Alexander, Schön, Gerd, and Hermon, Ziv. 1997. Quantum manipulations of
small Josephson junctions. Physical Review Letters, 79(12), 2371–2374.
Shor, Peter W. 1995. Scheme for reducing decoherence in quantum computer memory.
Physical Review A, 52(Oct), R2493–R2496.
Silver, A., and Zimmerman, J. 1967. Quantum states and transitions in weakly connected
superconducting rings. Physical Review, 157(2), 317–341.
Steane, Andrew. 1996. Multiple-particle interference and quantum error correction. Pro-
ceedings of the Royal Society of London. Series A: Mathematical, Physical and
Engineering Sciences, 452(1954), 2551–2577.
Steffen, Matthias, Ansmann, M., Bialczak, Radoslaw C., et al. 2006. Measurement of
the entanglement of two superconducting qubits via state tomography. Science,
313(5792), 1423–1425.
Steffen, Matthias, Kumar, Shwetank, DiVincenzo, David P., et al. 2010. High-coherence
hybrid superconducting qubit. Physical Review Letters, 105(Sep), 100502.
Sundaresan, Neereja M., Liu, Yanbing, Sadri, Darius, et al. 2015. Beyond strong coupling
in a multimode cavity. Physical Review X, 5(2), 021035.
Susskind, L., and Glogower, J. 1964. Quantum mechanical phase and time operator. Physics,
1, 49–61.
Teufel, Stefan. 2003. Adiabatic Perturbation Theory in Quantum Dynamics. Lecture Notes
in Mathematics, vol. 1821. Springer Berlin Heidelberg.
Townsend, Christopher, Ketterle, Wolfgang, and Stringari, Sandro. 1997. Bose–Einstein
condensation. Physics World, 10(3), 29–36.
van den Brink, Alec Maassen, Berkley, A. J., and Yalowsky, M. 2005. Mediated tunable
coupling of flux qubits. New Journal of Physics, 7(1), 230–230.
van der Ploeg, S. H. W., Izmalkov, A., van den Brink, Alec Maassen, et al. 2007.
Controllable Coupling of Superconducting Flux Qubits. Physical Review Letters,
98(5), 057004.
van der Wal, C. H., ter Haar, A. C., Wilhelm, F. K., et al. 2000. Quantum superposition of
macroscopic persistent-current states. Science, 290(5492), 773–777.
Venturelli, Davide, Mandrá, Salvatore, Knysh, Sergey, O’Gorman, Bryan, Biswas, Rupak,
and Smelyanskiy, Vadim. 2015. Quantum optimization of fully connected spin
glasses. Physical Review X, 5(3), 031040.
Vlastakis, Brian, Kirchmair, Gerhard, Leghtas, Zaki, et al. 2013. Deterministically encoding
quantum information using 100-photon Schrödinger cat states. Science (New York,
N.Y.), 342(6158), 607–610.
Wallraff, A., Schuster, D. I., Blais, A., et al. 2004. Strong coupling of a single photon to
a superconducting qubit using circuit quantum electrodynamics. Nature, 431(7005),
162–167.
Wallraff, A., Schuster, D. I., Blais, A., et al. 2005. Approaching unit visibility for control
of a superconducting qubit with dispersive readout. Physical Review Letters, 95(6),
0650501.
Walls, D. F., and Milburn, G. J. 1994. Quantum Optics. Springer Berlin Heidelberg.
Walter, T., Kurpiers, P., Gasparinetti, S., et al. 2017. Rapid high-fidelity single-shot
dispersive readout of superconducting qubits. Physical Review Applied, 7(5), 054020.
Wang, C., Gao, Y. Y., Reinhold, P., et al. 2016. A Schrodinger cat living in two boxes.
Science, 352(6289), 1087–1091.
Williams, Colin P. 2011. Quantum Gates. Springer London. Pages 51–122.
Wilson, C. M., Johansson, G., Pourkabirian, A., et al. 2011. Observation of the dynamical
Casimir effect in a superconducting circuit. Nature, 479(7373), 376–379.
Wootters, W. K., and Zurek, W. H. 1982. A single quantum cannot be cloned. Nature,
299(5886), 802–803.
Wu, Yulin, Bao, Wan-Su, Cao, Sirui, Chen, et al. 2021. Strong quantum computational
advantage using a superconducting quantum processor. Physical Review Letters 127,
180501.
Yamamoto, T., Neeley, M., Lucero, E., et al. 2010. Quantum process tomography of
two-qubit controlled-Z and controlled-NOT gates using superconducting phase qubits.
Physical Review B, 82(18), 184515.
Yamamoto, T., Pashkin, Yu. A., Astafiev, O., Nakamura, Y., and Tsai, J. S. 2003. Demon-
stration of conditional gate operation using superconducting charge qubits. Nature,
425(6961), 941–944.
Yan, F., Gustavsson, S., Kamal, A., et al. 2015. The flux qubit revisited to enhance coherence
and reproducibility. e-print, arXiv:1508.06299.
Yan, Fei, Gustavsson, Simon, Kamal, Archana, Birenbaum, et al. 2016. The flux qubit
revisited to enhance coherence and reproducibility. Nature Communications, 7(1),
12964.
Yoshihara, Fumiki, Fuse, Tomoko, Ashhab, Sahel, Kakuyanagi, Kosuke, Saito, Shiro,
and Semba, Kouichi. 2016. Superconducting qubit–oscillator circuit beyond the
ultrastrong-coupling regime. Nature Physics, 13(1), 44–47.
Yurke, Bernard, and Denker, John. 1984. Quantum network theory. Physical Review A,
29(3), 1419–1437.
Zener, C. 1932. Non-adiabatic crossing of energy levels. Proceedings of the Royal Society
A: Mathematical, Physical and Engineering Sciences, 137(833), 696–702.
Zueco, David, and García-Ripoll, Juanjo. 2019. Ultrastrongly dissipative quantum Rabi
model. Physical Review A, 99(Jan), 013807.
299