Theoretical and Computational Aspects
of New Lattice Fermion Formulations
Christian Zielinski
Division of Mathematical Sciences
School of Physical and Mathematical Sciences
Nanyang Technological University
2016
arXiv:1703.06364v1 [hep-lat] 18 Mar 2017
T  C A
 N L F F
C Z
Division of Mathematical Sciences
School of Physical and Mathematical Sciences
Nanyang Technological University
A thesis submitted to the Nanyang Technological University
in partial fulfillment of the requirement for the degree of
Doctor of Philosophy in the Mathematical Sciences
2016
A
In this work we investigate theoretical and computational aspects of novel lat-
tice fermion formulations for the simulation of lattice gauge theories. The lattice
approach to quantum gauge theories is an important tool for studying quantum
chromodynamics, where it is the only known framework for calculating physi-
cal observables from first principles. In our investigations we focus on staggered
Wilson fermions and the related staggered domain wall and staggered overlap for-
mulations. Originally proposed by Adams, these new fermion discretizations bear
the potential to reduce the computational costs of state-of-the-art Monte Carlo
simulations. Staggered Wilson fermions combine aspects of both staggered and
Wilson fermions while having a reduced number of fermion doublers compared to
usual staggered fermions. Moreover, they can be used as a kernel operator for the
domain wall fermion construction with potentially significantly improved chiral
properties and for the overlap operator with its exact chiral symmetry. This allows
the implementation of chirality on the lattice in a controlled manner at potentially
significantly reduced costs. The practical potential and limitations of these new
lattice fermions are also critically discussed.
v
A
Over the course of the four years of my Ph.D. I had the pleasure to work with
many brilliant people in my field. My special thanks go to Dr. David H. Adams
for guiding me in my first three years, for plenty of insightful discussions, for all
the constructive feedback and his confidence in my abilities. I also want to thank
our collaborator Asst. Prof. Daniel Nogradi from the Eötvös Loránd University for
patiently answering many questions by email. My thanks also go to Assoc. Prof.
Wang Li-Lian for his supervision in my final year. Both in private discussions and
in his excellent lectures I learned a lot about the numerical foundations for my
research.
At the Nanyang Technological University I want to thank my lecturers, who
all contributed in widening my horizons. Moreover, I had the pleasure working
together with many great colleagues and having plenty of interesting discussions
with friends. My special thanks go here to Dr. Reetabrata Har, Jia Yiyang, Sai
Swaroop Sunku, Andrii Petrashyk, Christian Engel, Ido David Polak, Nguyen Thi
Phuc Tan, Liu Zheng, Chen Penghua and Lim Kim Song. During my time at the
Nanyang Technological University I got to know many great people, for whom I
am very thankful.
I spent the final year of my Ph.D. in the Theoretical Particle Physics Group at
the University of Wuppertal (Bergische Universität Wuppertal). Here my special
thanks go to PD Dr. Christian Hoelbling and Prof. Zoltán Fodor, who made my stay
in Wuppertal possible. I greatly enjoyed working together with Christian Hoelbling,
as not only did I learn a lot in our discussions, but he was also always encouraging
and highly motivated. My thanks also go to PD Dr. Stephan Dürr, who made a lot
of great suggestions and gave helpful feedback on our projects and manuscripts,
and to Prof. Ting-Wai Chiu at the National Taiwan University for answering many
questions by email. I was lucky to share my office with two great colleagues, Dr.
Attila Pásztor and Dr. Thomas Rae. My stay in Wuppertal was very enjoyable and
productive and my thanks go also to the rest of the group.
I want to thank PD Dr. Christian Hoelbling, Assoc. Prof. Wang Li-Lian, Jia Yiyang,
Christian Engel and Pooi Khay Chong for providing feedback on various parts of
a draft of this thesis. I also thank the three anonymous examiners of this thesis
for their constructive criticisms and valuable comments. Furthermore, I want to
acknowledge financial support from the Singapore International Graduate Award
(SINGA) and Nanyang Technological University.
Finally, I want to express my utmost gratitude to my family and my beloved wife
for their unconditional support over all these years.
vii
L  
As part of this Ph.D. the following work was published:
1.
C. Hoelbling and C. Zielinski, “Staggered domain wall fermions, PoS
LAT-
TICE2016 (2016) 254, arXiv:1609.05114 [hep-lat]
2. C. Hoelbling and C. Zielinski, “Spectral properties and chiral symmetry vio-
lations of (staggered) domain wall fermions in the Schwinger model, Phys.
Rev. D94 no. 1, (2016) 014501, arXiv:1602.08432 [hep-lat]
3.
D. H. Adams, R. Har, Y. Jia, and C. Zielinski, “Continuum limit of the axial
anomaly and index for the staggered overlap Dirac operator: An overview,
PoS LATTICE2013 (2014) 462, arXiv:1312.7230 [hep-lat]
4.
D. H. Adams, D. Nogradi, A. Petrashyk, and C. Zielinski, “Computational
efficiency of staggered Wilson fermions: A first look, PoS
LATTICE2013
(2014) 353, arXiv:1312.3265 [hep-lat]
Note: In lattice field theory authors are traditionally ordered by name.
ix
L  
Parts of this work have been presented by the author at the following venues:
2016 Workshop of the Collaborative Research Center Transregio SFB/TR-55,
University of Wuppertal, Wuppertal, Germany
2016 Spring Meeting of the German Physical Society,
University of Hamburg, Hamburg, Germany
2015 Invited Talk at the Theoretical Particle Physics Seminar,
University of Wuppertal, Wuppertal, Germany
2015 33
rd
International Symposium on Lattice Field Theory,
Kobe International Conference Center, Kobe, Japan
2015 Conference on 60 Years of Yang-Mills Gauge Field Theories,
Institute of Advanced Studies, Nanyang Technological University, Singapore
2015 9
th
International Conference on Computational Physics,
National University of Singapore, Singapore
2014 5
th
Singapore Mathematics Symposium,
National University of Singapore, Singapore
2014 32
nd
International Symposium on Lattice Field Theory,
Columbia University, New York City, United States
xi
CONTENTS
Abstract v
Acknowledgements vii
List of publications ix
List of presentations xi
1 Introduction 1
1.1 Outline of this thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Original contributions . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2 Gauge theories in a nutshell 5
2.1 Gauge theories in the continuum . . . . . . . . . . . . . . . . . . . 5
2.1.1 Fermions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.1.2 Gauge theories . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1.3 Chiral symmetry . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.4 Path integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2 Lattice gauge theories . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2.1 Wick rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.2 Naïve lattice fermions . . . . . . . . . . . . . . . . . . . . . . 12
2.2.3 Fermion doubling . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2.4 The Nielsen-Ninomiya No-Go theorem . . . . . . . . . . . . . 15
2.2.5 Coupling to gauge fields . . . . . . . . . . . . . . . . . . . . . 15
2.2.6 Wilson fermions . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2.7 Staggered fermions . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2.8 Staggered Wilson fermions . . . . . . . . . . . . . . . . . . . . 21
2.2.9 Gauge fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3 Numerical simulations . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.1 Monte Carlo method . . . . . . . . . . . . . . . . . . . . . . . 22
3 Staggered Wilson fermions 25
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.2 Symmetries of staggered fermions . . . . . . . . . . . . . . . . . . . 26
3.3 Flavored mass terms . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.4 Adams mass term . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.4.1 Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.4.2 Interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.4.3 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
xiii
Contents
3.4.4 Aoki phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.5 Hoelbling’s mass term . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.5.1 Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.5.2 Interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.5.3 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.6 Generalized mass terms . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.6.1 The four-dimensional case . . . . . . . . . . . . . . . . . . . . 35
3.6.2 The d-dimensional case . . . . . . . . . . . . . . . . . . . . . 37
4 Computational efficiency 39
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.2 Theoretical estimates . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.2.1 Number of iterations . . . . . . . . . . . . . . . . . . . . . . . 41
4.2.2 Cost of matrix-vector multiplications . . . . . . . . . . . . . . 42
4.3 Numerical study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.3.1 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.3.2 Overall speedup . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.4 Arithmetic intensities . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.4.1 Wilson fermions . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.4.2 Staggered fermions . . . . . . . . . . . . . . . . . . . . . . . . 51
4.4.3 Staggered Wilson fermions . . . . . . . . . . . . . . . . . . . . 51
4.4.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5 Pseudoscalar meson spectrum 55
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.2 Pseudoscalar mesons . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.3 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6 Staggered overlap fermions 63
6.1 Neubergers overlap fermions . . . . . . . . . . . . . . . . . . . . . . 64
6.1.1 The Ginsparg-Wilson relation . . . . . . . . . . . . . . . . . . 64
6.1.2 The overlap operator . . . . . . . . . . . . . . . . . . . . . . . 66
6.2 Staggered overlap fermions . . . . . . . . . . . . . . . . . . . . . . . 67
6.3 The index theorem and the axial anomaly . . . . . . . . . . . . . . . 68
6.4 Continuum limit of the index and axial anomaly . . . . . . . . . . . 69
6.4.1 The index and axial anomaly . . . . . . . . . . . . . . . . . . . 70
6.4.2 Continuum limit of the index density . . . . . . . . . . . . . . 71
7 Spectral properties 75
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.2 Free-field case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.2.1 Wilson fermions . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.2.2 Staggered Wilson fermions . . . . . . . . . . . . . . . . . . . . 78
7.3 Quenched quantum chromodynamics . . . . . . . . . . . . . . . . . 79
xiv
Contents
7.3.1 Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7.3.2 Eigenvalue spectra . . . . . . . . . . . . . . . . . . . . . . . . 81
7.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
8 Staggered domain wall fermions 85
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
8.2 Kernel operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
8.3 Domain wall fermions . . . . . . . . . . . . . . . . . . . . . . . . . 87
8.3.1 Standard construction . . . . . . . . . . . . . . . . . . . . . . 88
8.3.2 Boriçi’s construction . . . . . . . . . . . . . . . . . . . . . . . 89
8.3.3 Optimal construction . . . . . . . . . . . . . . . . . . . . . . 90
8.3.4 Staggered formulations . . . . . . . . . . . . . . . . . . . . . . 92
8.4 Effective Dirac operator . . . . . . . . . . . . . . . . . . . . . . . . . 93
8.4.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
8.4.2 The N
s
limit . . . . . . . . . . . . . . . . . . . . . . . . . 95
8.4.3 Approximate sign functions . . . . . . . . . . . . . . . . . . . 96
8.5 Setting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
8.5.1 Setting the domain wall height . . . . . . . . . . . . . . . . . . 98
8.5.2 Effective mass . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
8.5.3 Normality and Ginsparg-Wilson relation . . . . . . . . . . . . 99
8.5.4 Topological charge . . . . . . . . . . . . . . . . . . . . . . . . 100
8.6 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
8.6.1 Free-field case . . . . . . . . . . . . . . . . . . . . . . . . . . 101
8.6.2 U
(
1
)
gauge field case . . . . . . . . . . . . . . . . . . . . . . . 105
8.6.3 Smearing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
8.6.4 Topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
8.6.5 Spectral flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
8.6.6 Approaching the continuum . . . . . . . . . . . . . . . . . . . 112
8.7 Quenched quantum chromodynamics . . . . . . . . . . . . . . . . . 119
8.8 Optimal weights . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
8.9 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
9 Conclusions 123
9.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
9.2 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
Bibliography 125
xv
C
I
Gauge theories are some of the most fundamental building blocks in modern
physics. In a gauge theory, physical observables are invariant under transforma-
tions of some fundamental fields. This property is referred to as gauge symmetry.
Currently all known fundamental forces in nature can be described as classical
or quantum gauge theories. General relativity is invariant under arbitrary coordi-
nate transformations, quantum electrodynamics (QED) is an Abelian gauge theory,
while quantum chromodynamics (QCD) and the electroweak interaction are of the
Yang-Mills type.
The generally accepted theory of the strong nuclear force is quantum chromody-
namics, which describes the interactions between quarks and gluons. The strong
nuclear force binds quarks and gluons to form hadrons, among them nuclear par-
ticles such as the proton and neutron. In the case of quantum chromodynamics,
we deal with a non-Abelian gauge theory with gauge group
SU
(
3
)
. The associated
force carrier is the so-called gluon, the associated charge is called color. Quan-
tum chromodynamics is the primary setting of this thesis and has some peculiar
properties, in particular confinement and asymptotic freedom.
Confinement refers to the property that we can only observe color-singlet states
and, thus, quarks and gluons cannot propagate freely. When quarks are pulled
apart, the energy of the gluonic field in between them increases until a new quark
pair is created. While confinement is phenomenologically well-established, no
rigor mathematical proof of confinement is known. In fact, confinement would
follow from proving the existence of a positive mass gap in quantum Yang-Mills
theory, which is one of the famous Millennium Problems of the Clay Mathematics
Institute.
Asymptotic freedom, on the other hand, refers to the fact that strong nuclear
interactions become weak at high energy scales. In experiments, such as heavy ion
collisions, one can then find a new state of matter, the quark-gluon plasma. Due
to the extremely high energy densities in the quark-gluon plasma, quarks become
deconfined.
1
1 Introduction
In general, the non-Abelian nature of quantum chromodynamics makes its
treatment with analytical methods difficult. While in certain regimes quantum
chromodynamics can be treated by perturbative methods much in the same way
quantum electrodynamics is treated, in the low-energy regime of bound hadrons
the coupling constant of the strong nuclear interaction is of order
O
(
1
)
and pertur-
bation theory is rendered inapplicable.
An alternative and non-perturbative approach is the framework of lattice quan-
tum chromodynamics. Originally proposed by Wilson in Refs. [
5
,
6
], one discretizes
the theory by replacing continuous space-time by a hypercubic space-time grid,
the so-called lattice. The resulting numerical formulation can, then, be simulated
on computers. As the resulting lattices usually have a very large number of sites, in
particular in the case of four or more space-time dimensions, extracting accurate
predictions from the theory is computationally extremely challenging. Enormous
amounts of computational resources are, thus, required for lattice field theoretical
simulations and the interest in finding more efficient and accurate lattice formu-
lations is high. As a consequence the subject has been and remains a very active
topic of research. The vast numbers of proposed lattice formulations suggest that
no optimal formulation is known, but that some are more suitable than others
depending on the application.
As part of these ongoing research efforts, in this thesis we study theoretical and
computational aspects of staggered Wilson, staggered overlap and staggered do-
main wall fermions as proposed by Adams. These novel lattice fermion discretiza-
tions have the potential of reducing computational costs in numerical simulations
and allow the generalization of the overlap and domain wall fermion formulation
to the case of a staggered kernel. While the construction of these lattice fermi-
ons are very interesting from a theoretical point of view, we also evaluate them
under practical aspects and try to understand in what setting their use can be
advantageous.
Although the lattice approach is traditionally concerned with quantum chromo-
dynamics, it is also used to study a large spectrum of other quantum field theories.
In this thesis, for example, we consider both four-dimensional quantum chromody-
namics and the Schwinger model, i.e. two-dimensional quantum electrodynamics.
2
1.1 Outline of this thesis
1.1 Outline of this thesis
This thesis is structured as follows. In Chapter 2, we give a short review of Dirac
fermions and gauge theories both in the continuum and on the lattice. In Chapter
3, we recapitulate the construction and properties of flavored mass terms and
staggered Wilson fermions. In Chapter 4, we discuss the computational efficiency
of staggered Wilson fermions and compare them to the one of Wilson fermions. In
Chapter 5, we discuss the feasibility of spectroscopy calculations with staggered
Wilson fermions and explicitly compute the pseudoscalar meson spectrum. In
Chapter 6, we review the formulation of usual and staggered overlap fermions on
the lattice and analytically evaluate the continuum limit of the index and the axial
anomaly. In Chapter 7, we investigate the eigenvalue spectra of both the Wilson and
staggered Wilson kernel and relate them to the computational efficiency of their
corresponding overlap operators. In Chapter 8, we discuss the formulation of usual
and staggered domain wall fermions in detail and compare spectral properties and
chiral symmetry violations in the case of the Schwinger model. Finally, in Chapter
9 we summarize the central results of this thesis and provide an outlook.
1.2 Original contributions
Apart from an extensive review of the lattice fermion formulations considered in
this thesis, there are various novel contributions as listed in the following:
The generalized flavored mass terms for arbitrary mass splittings in arbitrary
even dimensions in Sec. 3.6 as proposed in Ref. [
2
] (done in collaboration
with C. Hoelbling).
The quantitative analysis of the computational efficiency and memory band-
width requirements of the staggered Wilson kernel in Chapter 4 as discussed
in Refs. [
4
,
7
,
8
] (done in collaboration with D. H. Adams, D. Nogradi and A.
Petrashyk).
The adaption of staggered spectroscopy methods to the two-flavor case and
the computation of the pseudoscalar meson spectrum with staggered Wilson
fermions in Chapter 5 as presented in Refs. [
9
,
10
] (done in collaboration
with D. H. Adams).
3
1 Introduction
The analysis of the continuum limit of the index and axial anomaly for the
staggered overlap Dirac operator
1
in Sec. 6.4 as given in Ref. [
3
] (done in
collaboration with D. H. Adams, R. Har and Y. Jia).
The discussion of the usual Wilson and staggered Wilson eigenvalue spec-
trum and its connection to the computational efficiency of the overlap con-
struction in Chapter 7 as presented in Refs. [
4
,
7
] (done in collaboration with
D. H. Adams).
The study of spectral properties and chiral symmetry violations of various
known and new variants of usual and staggered domain wall fermions in
Chapter 8 as discussed in Refs. [
1
,
2
,
12
] (done in collaboration with C. Hoel-
bling).
Some results presented in this thesis have already been published in the literature
as listed on page ix, while others were presented at international conferences as
listed on page xi. Throughout the thesis the author uses the inclusive “we, to
include both the reader and collaborators in the discussion.
1
Note here also R. Hars thesis in Ref. [11], elaborating on this derivation in more detail.
4
C
G   
Quantum gauge theories play a central role in modern physics. Their defining prop-
erty is that the action is invariant under a continuous group of local transformations.
The standard model can be understood as a gauge theory with non-Abelian gauge
group
SU
(
3
)
× SU
(
2
)
×
U
(
1
)
. It includes quantum chromodynamics (QCD) with
gauge group
SU
(
3
)
and the electroweak sector with gauge group
SU
(
2
)
×
U
(
1
)
. The
electroweak interaction describes a unified theory of the weak and electromagnetic
interactions.
Throughout this thesis we are dealing primarily with two specific gauge theories.
The first being quantum chromodynamics, which is a Yang-Mills theory with non-
Abelian symmetry group
SU
(
3
)
. The other one is quantum electrodynamics, which
is an Abelian gauge theory with gauge group U
(
1
)
.
In this chapter, we review the formulation of gauge theories in the continuum in
Sec. 2.1 and on the lattice in Sec. 2.2. We also briefly discuss the use of Monte Carlo
methods for the simulation of lattice gauge theories in Sec. 2.3. For notational ease,
we only discuss the case of four space-time dimensions. For a detailed discussion
of the continuum case, we refer the reader to the many textbooks in the field such
as Refs. [1315]. For the lattice formulation, see e.g. Refs. [1619].
2.1 Gauge theories in the continuum
We begin by reviewing some aspects of gauge theories in continuous Minkowski
space-time. Throughout this thesis we are using natural units, where
~ = c = 1 (2.1.1)
with the reduced Planck constant
~ h/
(
2π
)
and the speed of light
c
. In this unit
system, energy is the only remaining independent dimension. For converting back
to conventional units, we note that ~c = 197.3 MeV· fm.
5
2 Gauge theories in a nutshell
In our convention the Minkowski metric tensor reads
η
µν
= η
µν
=
1
1
1
1
(2.1.2)
and we follow Einsteins sum convention by summing over repeated indices. Greek
indices run over 0, 1, 2, 3, while roman indices are restricted to the spacial compo-
nents 1, 2, 3. Indices of four-vectors can be lowered and raised using the metric
tensor. If x denotes the spatial part of a four-vector, we have
x
µ
=
x
0
, x
|
, x
µ
= η
µν
x
ν
=
x
0
, x
|
. (2.1.3)
Furthermore, we introduce the derivative operator as
µ
=
x
µ
. (2.1.4)
In the following, we discuss the basic framework for the quantum field theories of
interest in continuous Minkowski space-time. We begin by discussing fermions in
Subsec. 2.1.1 and gauge theories in Subsec. 2.1.2, followed by chiral symmetry in
Subsec. 2.1.3 and the path integral formalism in Subsec. 2.1.4. In our notation and
discussion we follow Ref. [15].
2.1.1 Fermions
In the discussion of fermions we restrict ourselves to the case of spin-
1
2
Dirac fer-
mions, such as quarks and electrons. They can be described by the Dirac equation,
a relativistic first-order partial differential equation first introduced in Ref. [
20
]. In
the free-field case it is of the form
iγ
µ
µ
m
f
Ψ
(
x
)
= 0, (2.1.5)
where
m
f
denotes the fermion mass and
Ψ
is a four-component spinor. As
Ψ
is of
fermionic nature, it is described by anti-commuting Grassmann numbers [
21
]. The
γ
µ
matrices are complex 4
×
4 matrices and satisfy the Dirac algebra, given by the
anticommutation relations
{
γ
µ
, γ
ν
}
γ
µ
γ
ν
+ γ
ν
γ
µ
= 2η
µν
1
. (2.1.6)
6
2.1 Gauge theories in the continuum
An explicit representation of the γ
µ
matrices is given by the Weyl or chiral basis
γ
0
=
1
1
, γ
k
=
σ
k
σ
k
, γ
5
=
1
1
, (2.1.7)
where
1
refers to the 2 × 2 identity matrix, the σ
k
refer to the Pauli matrices
σ
1
=
1
1
, σ
2
=
i
i
, σ
3
=
1
1
(2.1.8)
and we introduced the chirality matrix
γ
5
iγ
0
γ
1
γ
2
γ
3
. We note that
γ
5
anticom-
mutes with all the other γ
µ
matrices.
The Dirac equation implies the Klein-Gordon equation
µ
µ
+ m
2
f
Ψ = 0. (2.1.9)
This is a manifestly covariant second-order partial differential equation and ex-
presses the relativistic energy-momentum relation. Its solutions describe spinless
particles. The Dirac equation can be seen as a square root” of the Klein-Gordon
equation, in the sense if Eq.
(2.1.5)
is multiplied by its complex conjugate, one
recovers Eq. (2.1.9).
Finally, the corresponding action of free Dirac fermions is given by
S
f
=
d
4
x Ψ
(
x
)
iγ
µ
µ
m
f
Ψ
(
x
)
, (2.1.10)
where
Ψ Ψ
γ
0
denotes the Dirac adjoint of
Ψ
. In the presence of a gauge field,
the partial derivatives
µ
are replaced by the covariant derivative
D
µ
as discussed
in the following subsection.
2.1.2 Gauge theories
We continue by briefly reviewing the formulation of the gauge sector. We discuss
both quantum electrodynamics as well as Yang-Mills theory.
The action
Quantum electrodynamics is a gauge theory with Abelian symmetry group U
(
1
)
.
The action is given by
S
g
=
1
4
d
4
x
F
µν
2
, (2.1.11)
where F
µν
=
µ
A
ν
ν
A
µ
is the field-strength tensor with vector potential A
µ
.
7
2 Gauge theories in a nutshell
The general case of a compact, semi-simple Lie group is described by a Yang-
Mills theory [
22
]. Here, we restrict ourselves to the case of the special unitary groups
SU
(
N
c
)
. The action reads
S
g
=
1
4
d
4
x
F
a
µν
2
, (2.1.12)
where
F
a
µν
=
µ
A
a
ν
ν
A
a
µ
+ g f
abc
A
b
µ
A
c
ν
is the field-strength tensor, the
A
a
µ
denote
the gauge fields with group index
a =
1
, . . . , N
2
c
1 and
g
is the coupling constant.
Coupling to matter
We can couple the gauge field to matter by replacing the partial derivatives in
Eq.
(2.1.10)
by covariant derivatives. In the case of quantum electrodynamics, this
replacement takes the form
µ
D
µ
=
µ
+ ieA
µ
, (2.1.13)
where
e
denotes the electrical charge of the electron. Note that the field-strength
tensor can be represented as
D
µ
, D
ν
= ieF
µν
. The action for quantum electrody-
namics then takes the form
S
QED
=
d
4
x
L
f
+ L
g
,
L
f
= Ψ
iγ
µ
D
µ
m
f
Ψ,
L
g
=
1
4
F
µν
2
.
(2.1.14)
By construction, the action
S
QED
is invariant under local U
(
1
)
transformations, for
which the infinitesimal form reads
Ψ
(
x
)
(
1 + iα
)
Ψ
(
x
)
,
Ψ
(
x
)
Ψ
(
x
)(
1 iα
)
,
A
a
µ
(
x
)
A
a
µ
(
x
)
1
e
µ
α
(2.1.15)
with a scalar function α α
(
x
)
.
For a Yang-Mills theory this generalizes to
µ
D
µ
=
µ
ig A
a
µ
t
a
, (2.1.16)
where
D
µ
, D
ν
= ig F
a
µν
t
a
. The
t
a
denote the generators of the gauge group
SU
(
N
c
)
, which obey commutation relations of the form
t
a
, t
b
= if
abc
t
c
with
8
2.1 Gauge theories in the continuum
structure constants f
abc
. The resulting action reads
S
YM
=
d
4
x
L
f
+ L
g
,
L
f
= Ψ
iγ
µ
D
µ
m
f
Ψ,
L
g
=
1
4
F
a
µν
2
.
(2.1.17)
The infinitesimal transformation law takes the form
Ψ
(
x
)
(
1 + iα
a
t
a
)
Ψ
(
x
)
,
Ψ
(
x
)
Ψ
(
x
)(
1 iα
a
t
a
)
, (2.1.18)
A
a
µ
(
x
)
A
a
µ
(
x
)
+
1
g
µ
α
a
+ f
abc
A
b
µ
α
c
with scalar functions α
a
α
a
(
x
)
.
The case of quantum chromodynamics
An important case throughout this thesis is quantum chromodynamics and there-
fore in the following we shall introduce some relevant terminology. Quantum
chromodynamics is the gauge theory of strong interactions with gauge group
SU
(
3
)
. The fermions are referred to as quarks and come in three colors. We say that
the quarks are in the fundamental representation of
SU
(
3
)
, while the eight gluon
fields
A
a
µ
(
a =
1
, . . . ,
8) transform under the adjoint representation of the
SU
(
3
)
color group. Hadrons come as color-singlet states and both quarks and gluons
have never been observed as free particles.
2.1.3 Chiral symmetry
One of the central symmetries discussed in this thesis is chiral symmetry. For the
discussion of this symmetry we follow Ref. [16] and write the Lagrangian density
L
0
= Ψ iγ
µ
D
µ
Ψ (2.1.19)
of a massless Dirac fermion. We observe the invariance of the Lagrangian density
under global chiral rotations of the form
Ψ
(
x
)
e
iαγ
5
Ψ
(
x
)
, Ψ
(
x
)
Ψ
(
x
)
e
iαγ
5
(2.1.20)
9
2 Gauge theories in a nutshell
and note that the presence of a mass term breaks this symmetry explicitly. We can
reformulate chiral symmetry also elegantly in the form of the anticommutation
relation
{
D, γ
5
}
= 0, (2.1.21)
where D = iγ
µ
D
µ
is the massless Dirac operator.
When considering
N
f
massless fermion flavors, one finds a larger group of chiral
symmetry transformations of the form
Ψ
(
x
)
e
iαγ
5
τ
a
Ψ
(
x
)
, Ψ
(
x
)
Ψ
(
x
)
e
iαγ
5
τ
a
, (2.1.22)
Ψ
(
x
)
e
iαγ
5
1
Ψ
(
x
)
, Ψ
(
x
)
Ψ
(
x
)
e
iαγ
5
1
, (2.1.23)
where the spinors now carry an implicit flavor index and the
τ
a
are the generators
of the group SU
(
N
f
)
with a = 1, . . . , N
2
f
1.
2.1.4 Path integrals
A systematic approach to the study of quantum field theories is the path integral
formalism [
23
,
24
]. It is also the starting point for the lattice approach, giving rise
to a discrete formulation suitable for numerical simulations.
To illustrate the formalism, let us consider an action
S
[
φ
]
=
d
4
x L
[
φ
]
of some
fields
φ
i
, which we collectively refer to as
φ
{
φ
i
}
, and an observable
O
[
φ
]
. If we
ignore some subtleties regarding the rigorous definition of the path integral for the
moment (cf. Ref. [
25
]), the time-ordered vacuum expectation value can be written
as
h
O
i
=
1
Z
Dφ O
[
φ
]
e
iS
[
φ
]
, Z =
Dφ e
iS
[
φ
]
. (2.1.24)
The notation
Dφ
i
Dφ
i
represents an infinite-dimensional integration
over all possible field configurations of the
φ
i
fields. The partition function is
denoted as Z and ensures the normalization of the expectation value.
When used for the study of gauge theories, one has to integrate over all gauge
fields
A
and fermion fields
Ψ
,
Ψ
and evaluate integrations of the form
DA DΨ DΨ
.
Due to gauge invariance, expectation values of observables are ill-defined due to
over-counting of physical degrees of freedom. This can be overcome by imposing a
gauge fixing condition using the Faddeev-Popov trick [
26
]. For more details of the
path integral method, we refer the reader to the widely available textbooks such as
Refs. [1315].
10
2.2 Lattice gauge theories
2.2 Lattice gauge theories
One of the standard tools for studying quantum field theories is perturbation theory.
It has proven itself to be an extremely successful method, in particular when applied
to quantum electrodynamics. As one of the major achievement of quantum physics,
the theoretical prediction of the electrons anomalous magnetic dipole moment
agrees with experiment to more than ten significant figures [
27
]. In quantum
electrodynamics, perturbative expansions take the form of a power series in the
fine-structure constant
α
1
/
137. As
α
1, higher terms in the expansion are
typically strongly suppressed and one can derive accurate predictions by including
only the first few orders.
For quantum chromodynamics, the coupling constant
α
s
at high energies or
short distances is also small. In this setting quantum chromodynamics can be
studied by perturbative methods as well. However, in the important low-energy
regime of quantum chromodynamics where bound hadron states are present, the
coupling constant
α
s
of the strong nuclear interaction is of order
O
(
1
)
and pertur-
bation theory is rendered inapplicable. Here lattice quantum chromodynamics is
the only known framework for calculating physical observables, such as hadron
masses, from first principles. After its proposal by Wilson in Refs. [
5
,
6
], the lattice
approach has become one of the standard tools for non-perturbative studies of the
strong nuclear interaction. In order to arrive at a lattice formulation of a quantum
field theory, one replaces continuous Euclidean space-time by a discrete lattice.
The discretization of the fields, the operators and the path integral then makes the
use of numerical methods possible.
In this section, we begin by reviewing the Wick rotation in Subsec. 2.2.1, followed
by the introduction of naïve lattice fermions in Subsec. 2.2.2. We then discuss
the problem of fermion doubling in Subsec. 2.2.3 and try to get an understanding
of the implications of the Nielsen-Ninomiya No-Go theorem in Subsec. 2.2.4. In
Subsec. 2.2.5, we discuss the coupling to gauge theories before we introduce Wilson
fermions in Subsec. 2.2.6, staggered fermions in Subsec. 2.2.7 and staggered Wilson
fermions in Subsec. 2.2.8. Finally in Subsec. 2.2.9, we discuss the discretization of
the gauge action on the lattice.
11
2 Gauge theories in a nutshell
2.2.1 Wick rotation
In Minkowski space-time the complex factor
exp
(
iS
)
appears in the path integral,
forbidding a probabilistic interpretation of expectation values. In order for Monte
Carlo techniques to be applicable, we move to Euclidean space-time. To that end
we analytically continue the time-component of four-vectors to purely imaginary
values. Instead of the four-vector
x
µ
=
x
0
, x
|
, we consider
x
µ
=
(
x, x
4
)
|
with
x
4
ix
0
. After this so-called Wick rotation [
28
], we find a weight factor
exp
(
S
)
with Euclidean action S in the path integral.
In Euclidean space-time the Lorentz group reduces to the four-dimensional
rotation group. For convenience we replace the Dirac algebra as stated in Eq.
(2.1.6)
by
γ
µ
, γ
ν
= 2δ
µ, ν
1
. (2.2.1)
An explicit representation of the γ
µ
matrices in the chiral basis is given by
γ
k
=
iσ
k
iσ
k
, γ
4
=
1
1
, γ
5
=
1
1
(2.2.2)
with the Euclidean chirality matrix
γ
5
γ
0
γ
1
γ
2
γ
3
. In Euclidean space-time the
action of a gauge theory as discussed in Subsec. 2.1.2 takes the form
S =
d
4
x
L
g
+ L
f
, L
f
=
Ψ
γ
µ
D
µ
+ m
f
Ψ, (2.2.3)
where
L
g
=
1
4
F
2
with
F
being the respective field-strength tensor of the theory.
Note that in Euclidean space-time the lowering or raising of indices has no effect,
so we consequently write lowered indices.
2.2.2 Naïve lattice fermions
After having moved to Euclidean time, we can now begin with the discretization
of the path integral. We replace the continuous space-time domain by a discrete
hypercubic lattice
= (a
x
Z) × (a
y
Z) × (a
z
Z) × (a
t
Z) (2.2.4)
with lattice spacings
a
x
,
a
y
,
a
z
in the spatial directions and
a
t
in the temporal
direction. In numerical simulations one considers a hypercubic subdomain
Λ
12
2.2 Lattice gauge theories
of the form
Λ =
a
x
n
x
a
y
n
y
a
z
n
z
a
t
n
t
0 n
x
< N
x
0 n
x
< N
y
0 n
x
< N
z
0 n
x
< N
t
(2.2.5)
with
N
x
,
N
y
,
N
z
and
N
t
being the number of slices in each respective dimension.
On this finite domain one usually imposes (anti-)periodic boundary conditions for
the fields. To simplify our discussion, we restrict ourselves to the most common
special case of an isotropic lattice
a a
x
= a
y
= a
z
= a
t
, (2.2.6)
where the lattice spacing is the same along all axes.
Having introduced the lattice domain, we discretize the derivative in Eq.
(2.2.3)
,
which can be done in a straightforward manner by using the antihermitian sym-
metric difference operator
µ
Ψ
(
x
)
µ
Ψ
(
x
)
1
2a
[
Ψ
(
x + a ˆµ
)
Ψ
(
x a ˆµ
)]
, (2.2.7)
where
ˆµ
refers to a unit vector in
µ
-direction. This discretization has the correct
continuum limit
lim
a0
µ
φ
(
x
)
=
µ
φ
(
x
)
, (2.2.8)
namely it reproduces the derivative operator for a 0.
In the action, the space-time integration is replaced by a sum over the lattice
sites, i.e.
d
4
x a
4
x Λ
. The resulting action reads
S
n
= ΨD
n
Ψ a
4
x, y Λ
Ψ
(
x
)
D
n
(
x, y
)
Ψ
(
y
)
(2.2.9)
with the lattice Dirac operator
D
n
= γ
µ
µ
+ m
f
1
. (2.2.10)
The operator density
D
(
x, y
)
is here defined as
DΨ
(
x
)
=
y
D
(
x, y
)
Ψ
(
y
)
for an
operator D, which yields in our case
D
n
(
x, y
)
=
1
2a
γ
µ
δ
y,x+a ˆµ
δ
y,xa ˆµ
+ m
f
δ
x, y
. (2.2.11)
We note that the Dirac operator respects chiral symmetry in the sense that
D
0
n
, γ
5
=
13
2 Gauge theories in a nutshell
0, where D
0
n
denotes the massless naïve lattice Dirac operator.
In our discrete setting the path integral measure now takes the form
DΨ DΨ
x, µ
dΨ
µ
(
x
)
y, ν
dΨ
ν
(
y
)
(2.2.12)
and expectation values translate to
h
O
i
=
1
Z
x, µ
dΨ
µ
(
x
)
y, ν
dΨ
ν
(
y
)
O
Ψ, Ψ
e
S
f
Ψ, Ψ
, (2.2.13)
Z =
x, µ
dΨ
µ
(
x
)
y, ν
dΨ
ν
(
y
)
e
S
f
Ψ, Ψ
. (2.2.14)
We emphasize that we discretized here in the simplest possible way. As the name
suggests, naïve lattice fermions are not suitable for actual simulations as we explain
in the following subsection.
2.2.3 Fermion doubling
To understand a fundamental problem of the naïve discretization, let us compute
the free-field propagator by inverting the lattice Dirac operator in Eq. (2.2.11). For
this well-known derivation we follow Ref. [
16
]. We begin by Fourier transforming
D
n
and writing the result as
ˆ
D
N
(
p, q
)
= δ
(
p q
)
ˆ
D
(
p
)
with
ˆ
D
(
p
)
=
i
a
γ
µ
sin
p
µ
a
+ m
f
1
. (2.2.15)
Here the hat indicates a Fourier transformed expression. We can invert
ˆ
D
(
p
)
to
find
ˆ
D
1
(
p
)
=
m
f
1
iγ
µ
sin
p
µ
a
/a
m
2
f
+
ν
sin
2
(
p
ν
a
)
/a
2
. (2.2.16)
Specializing it to the massless case, Eq. (2.2.16) reduces to
ˆ
D
1
(
p
)
m
f
=0
=
iγ
µ
sin
p
µ
a
/a
ν
sin
2
(
p
ν
a
)
/a
2
. (2.2.17)
We can verify that Eq.
(2.2.17)
has the correct continuum limit for fixed values
of
p
. The resulting propagator
iγ
µ
p
µ
/p
2
of a massless fermion has a pole at
p =
(
0, 0, 0, 0
)
|
only. On the other hand, we observe that the lattice propagator
in Eq.
(2.2.16)
has a pole whenever all components take up values
p
µ
{
0, π/a
}
.
In addition to the physical pole, we find additional 15 unphysical poles in the
14
2.2 Lattice gauge theories
corners of the Brillouin zone, among them e.g.
p =
(
π/a, 0, 0, 0
)
|
and
p =
(
π/a, π/a, π/a, π/a
)
|
.
The appearance of these spurious states on the lattice is the notorious fermion
doubler problem. The presence of these doubler states is related to the use of
the symmetric lattice derivative in Eq.
(2.2.11)
. However, the use of left or right
derivatives would give rise to non-covariant contributions, which would render
the theory non-renormalizable [29].
2.2.4 The Nielsen-Ninomiya No-Go theorem
A more complete understanding of the phenomenon of fermion doublers and their
connection to chiral symmetry is given by the Nielsen-Ninomiya No-Go theorem
[
30
32
]. It states that there can be no net chirality under a set of relatively general
conditions, namely the lattice Hamiltonian of a fermion being local, quadratic in
the fields and invariant under both lattice translations as well as under a change of
the phase of the fields [
33
]. In this case the number of left-handed and right-handed
states must be equal.
This theorem explains the difficulties in constructing fermion formulations with
all the properties one would like to have. Up to this day, no theoretically sound
lattice fermion formulation is known which satisfies all of the following nontrivial
conditions:
1. Absence of fermion doublers;
2. Respecting chiral symmetry;
3. Being computationally efficient.
In a realistic setting (e.g. dynamical four-dimensional quantum chromodynamics)
one can generally satisfy at most two of the three conditions and we evaluate the
novel fermion discretizations discussed in this thesis with respect to these points.
2.2.5 Coupling to gauge fields
Up to this point we only considered naïve lattice fermions in the free-field case. If
we seek to couple fermions to a gauge field, we have to include a suitable interaction
term in the fermion action.
A peculiarity of the lattice formulation is that one introduces the gauge fields as
elements of the gauge group
G
, rather than as elements of the Lie algebra
g
like in
15
2 Gauge theories in a nutshell
the continuum. In the following, we restrict ourselves to the case of the (special)
unitary groups, i.e. for all U G the defining property U
= U
1
holds.
Let us now introduce (oriented) gauge links
U
µ
(
x
)
G
for
µ =
1
, . . . ,
4 and
x Λ
. We include interactions with the gauge field by incorporating the gauge
links in the finite difference operator via
µ
Ψ
(
x
)
1
2a
U
µ
(
x
)
Ψ
(
x + a ˆµ
)
U
µ
(
x a ˆµ
)
Ψ
(
x a ˆµ
)
, (2.2.18)
resulting in
D
n
(
x, y
)
=
1
2a
γ
µ
U
µ
(
x
)
δ
y,x+a ˆµ
U
µ
(
x a ˆµ
)
δ
y,xa ˆµ
+ m
f
δ
x, y
. (2.2.19)
Note that the fermionic degrees of freedom live on the lattice sites, while the gauge
degrees of freedoms are located on the links connecting neighboring sites. If we
now choose
(
x
)
G and do a gauge transformation of the form
Ψ
(
x
)
(
x
)
Ψ
(
x
)
,
Ψ
(
x
)
Ψ
(
x
)
(
x
)
,
U
µ
(
x
)
(
x
)
U
µ
(
x
)
(
x + a ˆµ
)
,
(2.2.20)
we find that the action
S
n
= ΨD
n
Ψ
is invariant with respect to the symmetry
group
G
. This transformation is the lattice equivalent of the continuum gauge
transformations as described in Eqs. (2.1.15) and (2.1.18).
2.2.6 Wilson fermions
Let us now return to the problem of fermion doublers. Wilsons original proposal
[5, 6] for a lattice Dirac operator can be written in the form
D
w
(
m
f
)
= γ
µ
µ
+ m
f
1
+ W
w
, W
w
=
ar
2
(2.2.21)
and is the result from adding the so-called Wilson term
W
w
to the naïve lattice
Dirac operator. The parameter
r
(
0, 1
]
denotes the Wilson parameter and
is
the covariant lattice Laplacian. The corresponding action reads
S
w
= ΨD
w
Ψ
. In
terms of the parallel transporters
T
µ
Ψ
(
x
)
= U
µ
(
x
)
Ψ
(
x + a ˆµ
)
, (2.2.22)
T
µ
Ψ
(
x
)
= U
µ
(
x a ˆµ
)
Ψ
(
x a ˆµ
)
= T
1
µ
Ψ
(
x
)
, (2.2.23)
16
2.2 Lattice gauge theories
we can write all terms concisely as
µ
=
1
2a
T
µ
T
µ
, =
2
a
2
µ
C
µ
1
, C
µ
=
1
2
T
µ
+ T
µ
. (2.2.24)
We note that
D
w
= γ
5
D
w
γ
5
and therefore
det D
w
R
. For the important case of two
degenerate fermion flavors, we can construct an action in terms of the operator
D
w
D
w
with det
D
w
D
w
0.
Due to the Wilson term, the free-field propagator in momentum space changes
from Eq. (2.2.16) to
ˆ
D
1
w
(
p
)
=
M
(
p
)
1
iγ
µ
sin
p
µ
a
/a
M
2
(
p
)
+
ν
sin
2
(
p
ν
a
)
/a
2
(2.2.25)
with
M
(
p
)
= m
f
+
2
r a
1
µ
sin
2
p
µ
a/2
, see e.g. Ref. [
19
]. In the corners of the
Brillouin zone the mass term is effectively modified to
m
f
+
2
r n/a
, where
n
is the
number of components of the lattice momentum
p
with value
π/a
. While the
physical mode is unaffected, the fermion doublers acquire a mass of
O
a
1
. In
total we find four doubler branches, where the
k
th
branch has a mass of
m
f
+
2
r k /a
and contains
4
k
modes. In the continuum limit the number of flavors is then
reduced from sixteen to a single physical flavor.
A shortcoming of Wilson fermions is that even for
m
f
=
0 chiral symmetry is bro-
ken, i.e.
{
D
w
, γ
5
}
,
0. Nevertheless, Wilson fermions and their improved versions
remain some of the most popular choices for lattice theoretical calculations due to
their conceptual simplicity.
2.2.7 Staggered fermions
An alternative approach to the fermion doubling problem are staggered fermions,
also known as Kogut-Susskind fermions [
34
37
]. Here one lifts an exact four-fold
degeneracy of naïve lattice fermions. As a result, the number of fermion species is
reduced from sixteen to four (in four space-time dimensions).
Let us briefly review the construction of staggered fermion, where we follow
Ref. [19]. Our starting point is the naïve fermion action
S
n
= a
4
x
Ψ
(
x
)
1
2a
γ
µ
[
Ψ
(
x + a ˆµ
)
Ψ
(
x a ˆµ
)]
+ m
f
Ψ
(
x
)
(2.2.26)
17
2 Gauge theories in a nutshell
in the free-field case. We can do a local change of variables of the form
Ψ
(
x
)
Γ
(
x
)
χ
(
x
)
, Ψ
(
x
)
χ
(
x
)
Γ
(
x
)
, (2.2.27)
where Γ
(
x
)
is a unitary matrix. By choosing Γ
(
x
)
so that
Γ
(
x
)
γ
µ
Γ
(
x + a ˆµ
)
= η
µ
(
x
)
1
, (2.2.28)
where
η
µ
(
x
)
is a scalar function, we can achieve a spin diagonalization. That this
is possible can be verified explicitly for the choice
Γ
(
x
)
=
µ
γ
x
µ
/a
µ
, η
µ
(
x
)
=
(
1
)
µ1
ν=0
x
ν
/a
(2.2.29)
with η
1
(
x
)
= 1. The action now takes the form
S
n
= a
4
4
ν=1
x
χ
ν
(
x
)
1
2a
η
µ
(
x
)[
χ
ν
(
x + a ˆµ
)
χ
ν
(
x a ˆµ
)]
+ m
f
χ
ν
(
x
)
,
(2.2.30)
where we wrote the spinor index
ν
explicitly. After this change of variables, we can
see that there are four identical copies in the sum over
ν
. We can lift this degeneracy
manually by omitting three of them, giving rise to the staggered fermion action
S
st
= a
4
x
χ
(
x
)
1
2a
η
µ
(
x
)[
χ
(
x + a ˆµ
)
χ
(
x a ˆµ
)]
+ m
f
χ
(
x
)
. (2.2.31)
We emphasize that
χ
(
x
)
is now a one-component spinor and all that remains
from the
γ
µ
matrices is the staggered phase factor
η
µ
(
x
)
. By gauging the derivative
operators, we can couple staggered fermions to a gauge field as discussed in Subsec.
2.2.5. In summary, the staggered action and the staggered Dirac operator read
S
st
= χD
st
χ, D
st
= η
µ
µ
+ m
f
1
, (2.2.32)
where we again make use of the matrix-vector-notation. We note that
D
st
= D
st
,
where
(
x
)
(
1
)
µ
x
µ
/a
, (2.2.33)
and det D
st
0 as all eigenvalues of D
st
come in complex conjugate pairs.
As the action in Eq.
(2.2.31)
describes four fermion species, a natural problem is
the one of flavor identification. The two common approaches are in coordinate
space [
38
,
39
] or in momentum space [
40
42
]. We note that both approaches
18
2.2 Lattice gauge theories
are identical in the continuum limit [
43
] and quickly review the schemes in the
following.
Flavor identification in coordinate space
We begin with the identification of flavors in coordinate space, following Refs. [
16
,
19
]. To this end let us relabel the fields as
χ
ρ
(
x
)
χ
(
2x + a ρ
)
, where the multi-
index takes values
ρ
{
0, 1
}
4
. We remark that the field
χ
ρ
(
x
)
effectively lives on a
lattice with spacing
a
0
=
2
a
and that
η
µ
(
2x + a ρ
)
= η
µ
(
a ρ
)
. We then define the
matrix-valued fermion fields
Ψ
(
x
)
=
1
8
ρ
Γ
(
a ρ
)
χ
ρ
(
x
)
,
Ψ
(
x
)
=
1
8
ρ
χ
ρ
(
x
)
Γ
(
a ρ
)
.
(2.2.34)
This relation can be inverted to give
χ
(
2x + a ρ
)
= 2 tr
Γ
(
a ρ
)
Ψ
(
x
)
,
χ
(
2x + a ρ
)
= 2 tr
Ψ
(
x
)
Γ
(
a ρ
)
.
(2.2.35)
Reformulating the free staggered fermion action using these new fields, we eventu-
ally find
S
st
= a
02
f ,x
Ψ
f
(
x
)
γ
µ
µ
Ψ
f
(
x
)
a
0
2
γ
5
ξ
5
ξ
µ
f g
µ
Ψ
g
(
x
)
+ m
f
Ψ
f
(
x
)
, (2.2.36)
where we introduced a flavor index
f =
1
, . . . ,
4 via
Ψ
f
µ
(
x
)
Ψ
µf
(
x
)
. Furthermore,
all finite difference operators are with respect to the lattice spacing
a
0
of the blocked
lattice and we define
ξ
µ
γ
|
µ
. While the first and third term in Eq.
(2.2.36)
is the
usual kinetic and mass term of the four fermion species
Ψ
f
, the second term has
the form of a Wilson term with an additional nontrivial spin-flavor structure.
In contrast to Wilsons construction, staggered fermions partially preserve chiral
symmetry in the massless case. For
m
f
=
0, the staggered fermion action has a
U
(
1
)
× U
(
1
)
symmetry, given by the transformations
Ψ
(
x
)
e
iα
Ψ
(
x
)
, Ψ
(
x
)
Ψ
(
x
)
e
iα
, (2.2.37)
Ψ
(
x
)
e
i βΘ
55
Ψ
(
x
)
, Ψ
(
x
)
Ψ
(
x
)
e
i βΘ
55
. (2.2.38)
19
2 Gauge theories in a nutshell
Here
Θ
55
γ
5
ξ
5
is the generator of the remnant chiral symmetry, which can be
also expressed as
D
0
st
,
=
0 with
D
0
st
being the massless staggered Dirac operator.
Flavor identification in momentum space
Another method of flavor identification is carried out in momentum space. We
follow Ref. [
19
] in the discussion and refer the reader to this reference for a more
complete discussion.
First note that we can write the staggered phase as
η
µ
(
x
)
= e
iδ
µ
·x/a
with δ
µ
ν
= π for ν < µ and δ
µ
ν
= 0 else. In momentum space we can write
S
st
=
π
π
d
4
p
(
2π
)
4
d
4
q
(
2π
)
4
χ
(
q
)
ˆ
D
(
q, p
)
χ
(
p
)
, (2.2.39)
where the lattice momenta
p
and
q
are in dimensionless lattice units,
χ
is 2
π
-
periodic, the Fourier transformed Dirac operator reads
ˆ
D
(
q, p
)
=
(
2π
)
4
i sin
p
µ
δ
p
(
p q + δ
µ
)
+ m
f
δ
p
(
p q
)
, (2.2.40)
and
δ
p
refers to the periodic Kronecker
δ
. If we define multi-indices
A
and
B
, which
each sum over all elements in
{
0, 1
}
4
, we can rewrite the staggered action as
S
st
=
A,B
π/2
π/2
d
4
p
(
2π
)
4
d
4
q
(
2π
)
4
χ
A
(
q
)
ˆ
D
A,B
(
q, p
)
χ
B
(
p
)
(2.2.41)
with
ˆ
D
A,B
(
q, p
)
=
(
2π
)
4
e
iπB
µ
i sin
p
µ
δ
p
(
p q + πB πA + δ
µ
)
+m
f
δ
p
(
p q + πB πA
)
. (2.2.42)
The fields are defined as
χ
A
(
p
)
χ
(
p + πA
)
and
χ
B
(
p
)
χ
(
p + πB
)
. Now
note the relations
δ
p
(
p q + πB πA
)
= δ
A,B
δ
(
p q
)
, (2.2.43)
δ
p
(
p q + πB πA + δ
µ
)
= ϕ
µ
A,B
δ
(
p q
)
, (2.2.44)
20
2.2 Lattice gauge theories
where
ϕ
µ
A,B
ν
1
2
e
i
(
πBπA+δ
µ
ν
)
+ 1
. (2.2.45)
If we now define
Γ
µ
A,B
exp
iπB
µ
ϕ
µ
A,B
, we can show that they satisfy the Dirac
algebra
{
Γ
µ
, Γ
ν
}
=
2
δ
µ, ν
1
and that the
Γ
µ
are unitary equivalent to
γ
µ
1
in
spin
flavor language. By using the above relations and adopting the spin
flavor
notation, Eq. (2.2.41) takes the form
S
st
=
π/2
π/2
d
4
p
(
2π
)
4
q
(
p
)
γ
µ
1
i sin
p
µ
+ m
f
(
1
1
)
q
(
p
)
, (2.2.46)
where
q
now has an implicit spin and flavor index. The action in this form is
invariant under the full U
(
4
)
U
(
4
)
chiral symmetry group, but this comes at the
price of a non-local action in position space [19].
2.2.8 Staggered Wilson fermions
The construction of staggered Wilson fermions was originally proposed by Adams
[
44
,
45
] and later extended by Hoelbling [
46
]. Starting from the staggered fermion
action a suitable flavored mass term is added, reducing the number of doublers and
making them suitable kernel operators for the overlap and domain wall fermion
construction. Due to the central role of staggered Wilson fermions in this thesis, we
dedicate Chapter 3 to the discussion of their derivation, properties and symmetries.
2.2.9 Gauge fields
In addition to the fermionic fields, we need to formulate gauge fields on the lattice.
To this end we have to discretize the continuum gauge action given in Subsec. 2.1.2.
We require the lattice gauge action to be formulated in terms of the link variables
and to respect gauge invariance. The simplest gauge invariant objects we can form
with gauge links are closed loops. This leads us to the introduction of the so-called
plaquette
U
µν
(
x
)
U
µ
(
x
)
U
ν
(
x + a ˆµ
)
U
µ
(
x + a ˆν
)
U
ν
(
x
)
, (2.2.47)
which is a minimal closed loop in the
µ
-
ν
-plane. We can then formulate a lattice
gauge action in terms of plaquettes only, ensuring gauge invariance by construc-
tion.
21
2 Gauge theories in a nutshell
In the case of an Abelian U
(
1
)
gauge theory, let us consider
S
g
= β
x
µ<ν
Re
1 U
µν
(
x
)
, (2.2.48)
where
β =
1
/e
2
is the inverse coupling. One can verify that in the limit of zero lattice
spacing and by rescaling
1
e
A
µ
A
µ
, one recovers the well-known continuum
expression for the action.
Although in the context of non-Abelian gauge theories we are primarily interested
in quantum chromodynamics, it makes sense to discuss the general case of the
symmetry group
SU
(
N
c
)
for
N
c
2. An important difference compared to the
Abelian case is that the ordering in Eq.
(2.2.47)
is now important as link variables
do not commute. The lattice action now generalizes to
S
g
=
β
N
c
x
µ<ν
Re
1
U
µν
(
x
)
, (2.2.49)
where
β =
2
N
c
/g
2
and
1
is the
N
c
×N
c
identity matrix. One can again easily verify
the correct continuum limit of the lattice discretization.
While the Wilson gauge action [
5
,
6
] in Eq.
(2.2.49)
is the simplest case of a gauge
action on the lattice, there are also improved actions in use. For examples we refer
the reader here to the Iwasaki [
47
,
48
] and DBW2 [
49
,
50
] actions, see also Ref. [
51
].
2.3 Numerical simulations
One of the major applications of the lattice approach to quantum field theory is
the simulation on a computer using Monte Carlo methods. Instead of giving a
detailed discussion of how a continuum limit is taken and physical observables
can be extracted, we give a short conceptual overview of the Monte Carlo method
in the context of lattice gauge theory and refer the reader to the many excellent
textbooks in the field, such as Refs. [
16
19
]. For a more general overview of the
Monte Carlo method, see e.g. Ref. [52].
2.3.1 Monte Carlo method
Assume we want to determine the expectation value of an observable
O
, e.g. in
order to extract a hadron mass, in the setting of a lattice gauge theory. We then
22
2.3 Numerical simulations
have to evaluate the expression
h
O
i
=
1
Z
DΨDΨDU O
Ψ, Ψ, U
e
S
Ψ, Ψ,U
, (2.3.1)
Z =
DΨDΨDU e
S
Ψ, Ψ,U
, (2.3.2)
where the combined fermion and gauge action reads
S
Ψ, Ψ, U
= S
f
Ψ, Ψ, U
+ S
g
[
U
]
(2.3.3)
and the generalization to several fermion species is straightforward. Assuming
the fermionic degrees of freedom appear quadratically, we can integrate them out
exactly. We arrive at an integrand, which only depends on the gauge field and reads
h
O
i
=
1
Z
DU O
0
[
U
]
det D
[
U
]
e
S
g
[
U
]
, (2.3.4)
Z =
DU det D
[
U
]
e
S
g
[
U
]
. (2.3.5)
We note that the integral over the spinors
Ψ
,
Ψ
is of the Grassmann / Berezin
kind [
21
] and gives rise to the fermion determinant
det D
[
U
]
with the lattice Dirac
operator
D
. The operator
O
0
[
U
]
follows from
O
Ψ, Ψ, U
after integrating out
the fermions in the path integral with the help of Wicks theorem [
53
]. We note that
in general O
0
[
U
]
contains propagators of the form D
1
[
U
]
.
To give an explicit example following Ref. [
16
], consider two Dirac fermion species
Ψ
u
, Ψ
d
and let
O = Ψ
d
(
x
)
Γ Ψ
u
(
x
)
Ψ
u
(
y
)
Γ Ψ
d
(
y
)
(2.3.6)
be an iso-triplet operator, where
Γ
is a monomial of
γ
µ
matrices. After integrating
out both the Ψ
u
and Ψ
d
fields, one finds
O
0
= tr
Γ D
1
u
(
x, y
)
Γ D
1
d
(
y, x
)
, (2.3.7)
where
D
u
and
D
d
are the lattice Dirac operators of the respective fermion species.
As the resulting integral is of extremely high dimension, one makes use of Monte
Carlo methods for the numerical evaluation. To this end one interprets the factor
1
Z
det D
[
U
]
e
S
g
[
U
]
, (2.3.8)
the so-called Gibbs-measure, as a probability measure. One first generates a
23
2 Gauge theories in a nutshell
Markov chain of gauge configurations
U
i
with respect to this probability distribu-
tion. Under certain conditions we can then replace the path integral expectation
value by
h
O
i
= lim
N →∞
1
N
N
i =1
O
[
U
i
]
, (2.3.9)
where we take an ensemble average.
The computationally most expensive part is the generation of the gauge ensem-
bles according to the probability distribution in Eq.
(2.3.8)
due to the presence of the
fermion determinant
det D
[
U
]
. Moreover, as one approaches the continuum limit
and one gets closer to the physical point, the computational costs rapidly increase,
see e.g. Refs. [
54
,
55
]. While in the early days of lattice quantum chromodynamics
numerical simulations where done in the so-called quenched approximation, i.e.
one does the replacement
det D
[
U
]
1, dynamical simulations are nowadays
state of the art.
The high computational costs of these Monte Carlo simulations explain why
research in novel lattice fermion formulations remains a very active field, see e.g.
Refs. [56, 57] for an overview.
24
C
S W 
The central topic of this thesis are staggered Wilson fermions and their related
formulations, namely staggered overlap and staggered domain wall fermions. For
this reason, we dedicate this chapter to the discussion of flavored mass terms and
the construction of staggered Wilson fermions.
The staggered Wilson fermion formulation first arose in investigations regarding
the index theorem on the lattice using staggered fermions [
44
]. The lattice Dirac
operator which emerged from these investigations [
45
] shares properties of both
staggered and Wilson fermions. Later Hoelbling expanded upon this idea and
proposed a related fermion formulation [
46
]. In Ref. [
2
], we went one step further
and introduced generalizations of these constructions for arbitrary flavor splittings
in arbitrary even dimensions.
3.1 Introduction
To understand the idea behind staggered Wilson fermions, we recall that there are
two traditional approaches to deal with the problem of fermion doublers. Wilsons
original approach is to add the Wilson term as discussed in Subsec. 2.2.6, which
is a symmetric covariant discretization of the Laplacian. The effect of this term
is that the spurious fermion states acquire a mass of
O
a
1
and, thus, decouple
in the continuum limit. The other approach is the one of staggered fermions as
introduced in Subsec. 2.2.7. Here one lifts an exact degeneracy of naïve fermions
using a spin diagonalization. As a result, not all fermion doublers are removed,
but the number of species is reduced from sixteen to four. We note that in practical
simulations one can further reduce the number of flavors to
N
f
by taking an appro-
priate root of the fermion determinant. Explicitly, one can make a replacement of
the form
det D
st
(
det D
st
)
N
f
/4
. (3.1.1)
25
3 Staggered Wilson fermions
Due to the lack of a formal proof of correctness, the method remained controversial
for a long time and one can find arguments in favor [
58
63
] and against [
64
66
]
this approach in the literature, see also Refs. [67, 68] for an overview.
Staggered Wilson fermions combine both approaches by taking the staggered
fermion action and adding a suitable staggered Wilson term to further reduce the
number of doublers. This staggered Wilson term turns out to be a combination of
the flavored mass terms as discussed in detail in the classical paper by Golterman
and Smit [
42
]. Due to the presence of this term, staggered Wilson fermions have
technical properties similar to Wilson fermions. By using them as a kernel opera-
tor, they allow the construction of staggered overlap and staggered domain wall
fermions. Usual staggered fermions, on the other hand, do not define a suitable
kernel operator. This is because these constructions rely on the property
γ
2
5
=
1
,
which does not hold for its staggered equivalent as we discuss in more detail in
Sec. 6.2.
In order to understand the construction of the staggered Wilson term and its
variants, we begin by discussing symmetries of the staggered fermion action in Sec.
3.2 and properties of flavored mass terms in Sec. 3.3. In Sec. 3.4, we then review
Adams proposal, followed by Hoelblings construction in Sec. 3.5. We end this
chapter with the discussion of our generalized mass terms in Sec. 3.6.
3.2 Symmetries of staggered fermions
In the following we discuss the symmetries of the massless staggered fermion
action
S =
1
2
x, µ
η
µ
(
x
)
¯χ
(
x
)
U
µ
(
x
)
χ
(
x + a ˆµ
)
¯χ
(
x
)
U
µ
(
x a ˆµ
)
χ
(
x a ˆµ
)
=
1
2
x, µ
η
µ
(
x
)
¯χ
(
x
)
U
µ
(
x
)
χ
(
x + a ˆµ
)
¯χ
(
x + a ˆµ
)
U
µ
(
x
)
χ
(
x
)
. (3.2.1)
The action is invariant under the transformations listed below, where we follow
the discussion and notation of Ref. [
42
] (cf. Ref. [
41
]). We note that in order for the
path integral measure to be invariant under all these transformations, we consider
an even number of lattice sites and (anti-)periodic boundary conditions.
26
3.2 Symmetries of staggered fermions
Shift invariance. The transformation is given by
χ
(
x
)
ζ
ρ
(
x
)
χ
(
x + a ˆρ
)
,
¯χ
(
x
)
ζ
ρ
(
x
)
¯χ
(
x + a ˆρ
)
,
U
µ
(
x
)
U
µ
(
x + a ˆρ
)
(3.2.2)
with
ζ
µ
(
x
)
=
(
1
)
4
ν=µ+1
x
ν
/a
(3.2.3)
and ζ
4
(
x
)
= 1.
Rotational invariance. The transformation is given by
χ
(
x
)
S
R
R
1
x
χ
R
1
x
,
¯χ
(
x
)
S
R
R
1
x
¯χ
R
1
x
,
U
(
x, y
)
U
R
1
x, R
1
y
,
(3.2.4)
where
R R
ρσ
is the rotation
x
ρ
x
σ
,
x
σ
x
ρ
,
x
µ
x
µ
with
µ , σ, ρ
. We
also introduced
S
R
R
1
x
=
1
2
1 ± η
ρ
(
x
)
η
σ
(
x
)
ζ
ρ
(
x
)
ζ
σ
(
x
)
+ η
ρ
(
x
)
η
σ
(
x
)
ζ
ρ
(
x
)
ζ
σ
(
x
)
(3.2.5)
for
ρ σ
and note that
S
R
R
1
x
η
µ
(
x
)
S
R
R
1
[
x + a ˆµ
]
= R
µν
η
ν
R
1
x
. More-
over, we make use of the compact notation
U
(
x, y
)
=
U
µ
(
x
)
, y = x + a ˆµ,
U
µ
(
y
)
, x = y + a ˆµ,
(3.2.6)
for the gauge links.
Axis reversal. The transformation is given by
χ
(
x
)
(
1
)
x
ρ
χ
I
ρ
x
,
¯χ
(
x
)
(
1
)
x
ρ
¯χ
I
ρ
x
,
U
(
x, y
)
U
I
ρ
x, I
ρ
y
,
(3.2.7)
where I
ρ
x is the reversal of the ρ-axis, i.e. x
ρ
x
ρ
, x
µ
x
µ
for µ , ρ.
27
3 Staggered Wilson fermions
U
(
1
)
symmetry. The transformation is given by
χ
(
x
)
e
iα
χ
(
x
)
,
¯χ
(
x
)
¯χ
(
x
)
e
iα
.
(3.2.8)
This symmetry is associated with the conversation of charge.
U
(
1
)
symmetry. The transformation is given by
χ
(
x
)
e
i β
(
x
)
χ
(
x
)
,
¯χ
(
x
)
¯χ
(
x
)
e
i β
(
x
)
.
(3.2.9)
with
(
x
)
=
(
1
)
µ
x
µ
/a
. (3.2.10)
This is the remnant chiral symmetry of staggered fermions.
Exchange symmetry. The transformation is given by
χ
(
x
)
¯χ
|
(
x
)
,
¯χ
(
x
)
χ
|
(
x
)
,
U
µ
(
x
)
U
µ
(
x
)
,
(3.2.11)
which interchanges fermion and anti-fermion field.
Remarks.
For the following discussion, let us also define the parity transforma-
tion as
I
1
I
2
I
3
followed by a shift in the 4-direction. We note, that in addition to
the position-space formulations of these symmetries, one can also express them
elegantly in momentum space. For brevity we do not quote the momentum repre-
sentation here explicitly, but refer the reader to Ref. [42] for details.
3.3 Flavored mass terms
Following Refs. [
41
,
42
] closely, we now discuss flavored mass terms. In the classical
continuum limit we can add a mass term
S
cont
M
=
d
4
x ΨM Ψ (3.3.1)
28
3.3 Flavored mass terms
to the action. We demand that
M
is trivial in spinor space and that the action is
rotationally invariant. The latter requires the mass term to commute with
γ
µ
γ
ν
.
Using an explicit spin
flavor-notation, a general flavored mass term now assumes
the structure
M =
1
m
1
+ m
µ
ξ
µ
+
1
2
m
µν
σ
µν
+ m
5
µ
iξ
µ
ξ
5
+ m
5
ξ
5
, (3.3.2)
where
m
µν
is antisymmetric,
σ
µν
iξ
µ
ξ
ν
and the
ξ
µ
are a representation of the
Dirac algebra in flavor space, i.e.
ξ
µ
, ξ
ν
=
2
δ
µν
1
. Furthermore, we require
that
M
2
0 in the free-field case, so
M
is taken to be Hermitian and the mass
parameters are real. We note that in Eq.
(3.3.2)
we could in principle also add
similar terms multiplied by a
γ
5
ξ
5
, but we do not consider them in the following
discussion.
Let us now introduce the symmetric shift operator
E
µ
(
x, y
)
=
1
2
z
ζ
µ
U
µ
(
z
)
δ
(
x z
)
δ
(
y z a ˆµ
)
+ U
µ
(
z
)
δ
(
y z
)
δ
(
x z a ˆµ
)
,
(3.3.3)
which we expect to result in a Hermitian transfer matrix. We can implement
Eq. (3.3.2) on the lattice as
S
M
= a
4
x
m χ
(
x
)
χ
(
x
)
+
x, y
m
µ
χ
(
x
)
E
µ
(
x, y
)
χ
(
y
)
+
i
2
a
4
x, y, z
m
µν
χ
(
x
)
E
µ
(
x, y
)
E
ν
(
y, z
)
χ
(
z
)
i
6
a
4
w, x, y, z
m
5
µ
ε
µα βγ
χ
(
w
)
E
α
(
w, x
)
E
β
(
x, y
)
E
γ
(
y, z
)
χ
(
z
)
1
24
a
4
v,w, x, y, z
m
5
ε
α βγδ
χ
(
w
)
E
α
(
v, w
)
E
β
(
w, x
)
E
γ
(
x, y
)
E
δ
(
y, z
)
χ
(
z
)
.
(3.3.4)
In order to get a better understanding of the mass term in Eq.
(3.3.4)
, let us discuss
which of the symmetries in Sec. 3.2 it breaks. First we note that in this general
form the mass term is not invariant under lattice rotations. However, violations
of rotational invariance are expected to be of order
O
(
a
)
, so that the symmetry
is eventually restored in the continuum limit. For parity transformations we find
invariance of our mass term, while for the exchange symmetry
m
,
m
µ
and
m
5
change signs.
29
3 Staggered Wilson fermions
For numerical applications, we require a real determinant and, thus, we need
to retain
Hermiticity. Hence the flavor structure of the mass term needs to be
restricted to a sum of products of an even number of
ξ
µ
. Therefore from now on
we restrict ourselves to the case
m
µ
= m
5
µ
= 0, (3.3.5)
leaving us with the m, m
µν
and m
5
term.
In Sec. 3.4, we follow Adams original two-flavor proposal and construct a stag-
gered Wilson term using
m
and
m
5
, while in Sec. 3.5 we discuss Hoelblings one
flavor construction using m and m
µν
. We finish this chapter with generalizations
of these flavored mass terms in Sec. 3.6.
3.4 Adams’ mass term
Adams proposed a particular form of the flavored mass term, giving rise to the
original two-flavor staggered Wilson fermion formulation introduced in Refs. [44,
45]. In the following, we discuss its construction, interpretation and properties.
3.4.1 Construction
The staggered Wilson action reads
S
sw
= ¯χD
sw
χ = ¯χ
(
D
st
+ m
f
1
+ W
st
)
χ (3.4.1)
with the massless staggered Dirac operator
D
st
= η
µ
µ
. The staggered Wilson term
W
st
reads
W
st
=
r
a
(
1
Γ
55
Γ
5
)
(3.4.2)
with Wilson-like parameter
r >
0. Following Adams original notation, we intro-
duced the operator
Γ
55
χ
(
x
)
=
(
1
)
µ
x
µ
/a
χ
(
x
)
, (3.4.3)
where depending on the context we also use
(
x
)
=
(
1
)
µ
x
µ
/a
for the sign factor.
In Eq. (3.4.2) we introduced
Γ
5
= η
5
C , η
5
= η
1
η
2
η
3
η
4
, η
5
χ
(
x
)
=
(
1
)
(
x
1
+x
3
)
/a
χ
(
x
)
(3.4.4)
30
3.4 Adams mass term
and the symmetrized product of the C
µ
=
T
µ+
+ T
µ
/2 operators
C =
(
C
1
C
2
C
3
C
4
)
sym
1
4!
sym
α βγδ
C
α
C
β
C
γ
C
δ
. (3.4.5)
As pointed out earlier, Adams flavored mass term in Eq.
(3.4.2)
preserves
Her-
miticity of the Dirac operator, namely
D
sw
= D
sw
, and is itself Hermitian. This
implies a non-negative fermion determinant
det D
sw
0 for suitable choices of the
fermion mass
m
f
and ensures the applicability of importance sampling techniques
to Monte Carlo simulations.
3.4.2 Interpretation
To interpret Adams staggered Wilson term, we note that in the spin
flavor inter-
pretation [42] we find
Γ
55
γ
5
ξ
5
, Γ
5
γ
5
1
+ O
a
2
. (3.4.6)
While
Γ
5
corresponds to the usual
γ
5
up to discretization effects, the
Γ
55
is a flavored
γ
5
in the sense that it acts nontrivially in flavor space. As a result, we find for Adams
staggered Wilson term the spin flavor structure
W
st
r
a
1
(
1
ξ
5
)
+ O
(
a
)
. (3.4.7)
We note that
W
st
is of the form of a projector
P
±
=
(
1 ± ξ
5
)
/
2 in flavor space and
allows for a simple interpretation. The staggered Wilson term gives a mass
O
a
1
to the two negative flavor-chirality species, similar to the effect the usual Wilson
term has on the doubler modes. In the continuum limit
a
0, the fermion
doublers become heavy and decouple. On the other hand, the two positive flavor-
chirality species do not acquire additional mass contributions. Note that the minus
sign in Eq.
(3.4.2)
and Eq.
(3.4.7)
can be replaced by a plus sign, thus interchanging
the role of positive and negative flavor-chirality species.
3.4.3 Properties
An important observation is that the equivalent of
γ
5
, namely
Γ
5
, does not square
to the identity operator. While
γ
2
5
=
1
holds exactly, we find
Γ
2
5
=
1
+ O
a
2
,
1
.
This is at the heart of the problem why usual staggered fermions are not suitable
31
3 Staggered Wilson fermions
kernel operators for the overlap and domain wall constructions (see Sec. 6.2 for
a more in-depth discussion). However, there is another operator which squares
exactly to unity, namely
Γ
2
55
=
1
. While
Γ
55
has a spin
flavor interpretation of
γ
5
ξ
5
, for Adams construction we project out the positive flavor-chirality species
in the continuum limit and find
Γ
55
= Γ
5
+ O
a
2
γ
5
1
+ O
a
2
(3.4.8)
on the physical species. Therefore on the subspace of these modes one can use
Γ
55
instead of
Γ
5
. This crucial insight by Adams allows the construction of staggered
overlap and staggered domain wall fermions, which we discuss in Chapter 6, 7 and
8.
To summarize some interesting properties of two-flavor staggered Wilson fer-
mions, we recall that compared to usual staggered fermions we are left with a
reduced number of tastes, namely two instead of four (in
d =
4 dimensions). We
also find a smaller fermion matrix and expect a better condition number compared
to Wilson fermions, potentially increasing computational efficiency. In addition,
staggered Wilson fermions allow the construction of staggered versions of overlap
and domain wall fermions.
On the downside, we point out that the staggered Wilson term
W
st
breaks a
subset of the staggered fermion symmetries. In particular, the exact flavored chiral
symmetry is broken in the massless case, i.e.
D
0
sw
,
,
0. This gives rise to an
additive mass renormalization and there is the need to fine-tune the bare mass
m
f
like for Wilson fermions. Moreover, new fermionic counterterms are allowed [
9
,
45
],
but the only effect is a wave function renormalization for the physical species. On
the physical species the staggered Wilson terms vanished up to order
O
(
a
)
, hence
the order
O
a
2
discretization error of usual staggered fermions is lost. Finally,
we note that the
SU
(
2
)
vector and chiral symmetries of the two physical flavors
are broken, similarly to the broken
SU
(
4
)
symmetry of usual staggered fermions.
Nevertheless, the remaining symmetries, such as the flavored rotation symmetries,
are still enough to ensure e.g. a degenerate triplet of pions [
69
]. We also note
that the staggered Wilson term is a four-hop operator, making it potentially very
susceptible to fluctuations of the gauge field [
70
,
71
] and difficult to parallelize. We
discuss these aspects in more detail in Chapter 4.
32
3.5 Hoelbling’s mass term
3.4.4 Aoki phase
We conclude our discussion of the theoretical properties of staggered Wilson fermi-
ons with the investigations of the Aoki phase [
72
78
] carried out in Refs. [
79
81
]. In
these studies, the authors investigated strong-coupling lattice quantum chromo-
dynamics with staggered Wilson fermions using hopping parameter expansions
and effective potential analyses. The existence of a non-vanishing pion conden-
sate in certain mass parameter ranges could be established, while massless pions
and PCAC
1
relations were found around a second-order phase boundary. These
results are in support of the idea that, like with Wilson fermions, staggered Wilson
fermions can be applied to lattice quantum chromodynamics by taking a chiral
limit.
3.5 Hoelbling’s mass term
One can build upon Adams idea and construct flavored mass terms to completely
lift the degeneracy of staggered fermions. Originally proposed by Hoelbling in four
dimensions in Ref. [46], we are going to review the construction in the following.
3.5.1 Construction
A possible approach to construct single-flavor staggered fermions is to start from
Adams operator and add an additional term, which lifts the degeneracy of species
of the same flavor-chirality. A natural candidate for this is the
m
µν
term in Eq.
(3.3.4)
with a spin
flavor interpretation of
1
σ
µν
. A possible candidate for a one-flavor
action [46] is then given by
S = ¯χD χ ¯χ
D
st
+ m
f
1
+
r
a
(2 ·
1
+ W
st
+ M
µν
)
χ, (3.5.1)
where we introduced the mass term M
µν
= iη
µν
C
µν
(no sum) with operators
η
µν
χ
(
x
)
=
(
1
)
Σ
ν
ρ=µ+1
x
ρ
/a
χ
(
x
)
for µ < ν,
η
µν
= η
νµ
for µ ν
(3.5.2)
and
C
µν
=
C
µ
, C
ν
/
2. Here
W
st
refers to Adams staggered Wilson term as dis-
cussed in Sec. 3.4 and the choice of µ , ν is arbitrary.
1
PCAC: Partially Conserved Axial Current
33
3 Staggered Wilson fermions
While the Dirac operator in Eq.
(3.5.1)
lifts the degeneracy of the staggered fer-
mion flavors completely, one can consider a more symmetric combination of the
mass terms
M
µν
, resulting in better symmetry properties. Hoelblings proposal [
46
]
now takes the form
S
1f
= ¯χD
1f
χ ¯χ
D
st
+ m
f
1
+
r
a
(2 ·
1
+ M
1f
)
χ (3.5.3)
with the symmetrized flavored mass term
M
1f
=
[
s
12
(
s
1
s
2
M
12
+ s
3
s
4
M
34
)
+ s
13
(
s
1
s
3
M
13
+ s
4
s
2
M
42
)
+ s
14
(
s
1
s
4
M
14
+ s
2
s
3
M
23
)]
/
3 (3.5.4)
and
s
µ
= ±
1,
s
µν
= ±
1 being arbitrary sign factors [
46
]. We note that in numerical
applications a commonly employed form of Eq. (3.5.4) is given by
M
1f
=
(
M
12
+ M
34
+ M
13
M
24
+ M
14
+ M
23
)
/
3, (3.5.5)
see Refs. [
71
,
82
]. As
M
1f
already lifts the degeneracy completely by itself, there is
no need to introduce a W
st
term in Eq. (3.5.3).
When compared to Adams staggered Wilson operator
W
st
, the mass term
M
1f
has a smaller symmetry group. One can verify the invariance [
46
] under diagonal
shifts given by
x x ±
ˆ
1 ±
ˆ
2 ±
ˆ
3 ±
ˆ
4
, shifted axis reversals given by
x
µ
x
µ
+ a ˆµ
and double rotations given by
x R
α β
R
γδ
x
, where
α
,
β
,
γ
and
δ
are mutually
distinct.
3.5.2 Interpretation
To understand the effect of Hoelblings flavored mass term, we note [46] that
M
1f
r
a
1
ξ
f
+ O
(
a
)
,
ξ
f
= diag
(
4, 0, 2, 2
)
or diag
(
2, 2, 4, 0
)
(3.5.6)
in the spin
flavor interpretation. One can now see that in the presence of the
mass term M
1f
three flavors are lifted to two distinct doublers points, namely two
flavors to 2
r /a
and one to 4
r /a
. The remaining physical flavor remains massless
under the action of the mass term, while the doublers decouple in the continuum
limit. Similar to the two flavor case, the staggered overlap and staggered domain
wall constructions can be carried out for Hoelblings operator.
34
3.6 Generalized mass terms
3.5.3 Properties
The Dirac operator
D
1f
breaks several symmetries of the staggered fermion action.
Among them is rotational symmetry, where only a residual subgroup survives. It
turns out that there are no additional fermionic counterterms in the action, but new
gluonic counterterms appear as pointed out by Sharpe [
69
]. They arise from the
fermion loop contributions to the gluonic two-, three- and four-point functions, cf.
Ref. [
9
]. This means that in dynamical simulations one has to include and fine-tune
these terms, severely limiting the practical applicability of single flavor staggered
fermions.
For this reason, starting from Chapter 4, we restrict ourselves to the case of
Adams original two-flavor staggered Wilson kernel.
3.6 Generalized mass terms
In the following, we want to further generalize the flavored mass terms discussed
in Sec. 3.4 and Sec. 3.5 to allow for arbitrary mass splittings in arbitrary even
dimensions. In this section, we follow our discussion
2
given in Ref. [
2
] closely,
where we previously presented the following results.
3.6.1 The four-dimensional case
Up to discretization terms, the staggered Wilson term is a mass term which is
trivial in spin-space, but splits the different staggered flavors. We require that
the determinant of the lattice fermion Dirac operator is real, to allow the use of
importance sampling techniques in Monte Carlo simulations. Adopting a more
convenient notation for the following discussion, the original proposal by Adams
[44, 45] can be written as
W
st
=
r
a
(
1
+ Γ
1234
C
1234
)
. (3.6.1)
Here r > 0 is the Wilson-like parameter and we define the operators
Γ
1234
χ
(
x
)
=
(
1
)
µ
x
µ
/a
χ
(
x
)
, (3.6.2)
2
Discussion based with permission on C. Hoelbling and C. Zielinski, “Spectral properties and
chiral symmetry violations of (staggered) domain wall fermions in the Schwinger model, Phys.
Rev. D94 no. 1, (2016) 014501, arXiv:1602.08432 [hep-lat]. Copyright 2016 by the American
Physical Society.
35
3 Staggered Wilson fermions
C
1234
= η
1
η
2
η
3
η
4
(
C
1
C
2
C
3
C
4
)
sym
. (3.6.3)
We recall that this term has a spin
flavor interpretation of the form of Eq.
(3.4.7)
and splits the four flavors of staggered fermions into two pairs with opposite flavor
chirality, that is, the eigenbasis of
ξ
5
. For the following discussion, the notation
A B means that A has the spin flavor interpretation B up to proportionality.
It is also possible to split the flavors with respect to the eigenbasis of other ele-
ments of the Dirac algebra in flavor space [
42
,
46
]. We restrict the structure of the
mass term to a sum of products of an even number of
ξ
µ
matrices, so that
Her-
miticity and the reality of the fermion determinant is retained. In four dimensions
we can, thus, reduce the number of staggered fermion flavors to one by e.g.
W
st
=
r
a
2 ·
1
+ W
12
st
+ W
34
st
, (3.6.4)
W
µν
st
= i Γ
µν
C
µν
, (3.6.5)
where the operators Γ
µν
and C
µν
are defined as
Γ
µν
χ
(
x
)
= ε
µν
(
1
)
(
x
µ
+x
ν
)
/a
χ
(
x
)
, (3.6.6)
C
µν
= η
µ
η
ν
·
1
2
C
µ
C
ν
+ C
ν
C
µ
(
no sum
)
. (3.6.7)
Here
ε
µ
1
···µ
N
is the totally antisymmetric Levi-Civita symbol. In order to interpret
the mass term defined in Eq. (3.6.5), we note that
Γ
µν
γ
µ
γ
ν
ξ
µ
ξ
ν
, C
µν
γ
µ
γ
ν
1
, γ
5
ξ
5
, (3.6.8)
up to discretization terms. Therefore, we find that
W
µν
st
1
σ
µν
+ O
(
a
)
(3.6.9)
with
σ
µν
= iξ
µ
ξ
ν
. We now see that the number of physical flavors is reduced by
W
st
to one, i.e. all but a single flavor acquire a mass of O
a
1
.
36
3.6 Generalized mass terms
3.6.2 The d -dimensional case
We can now generalize our constructions to an arbitrary even number of dimen-
sions d. Here we write a single flavor mass term as
W
st
=
r
a
d/2
k =1
1
+ W
(
2k1
)(
2k
)
st
, (3.6.10)
so that Eq.
(3.6.4)
follows after the specialization to
d =
4. We can generalize this
mass term even further by introducing
W
µ
1
···µ
2n
st
= i
n
Γ
µ
1
···µ
2n
C
µ
1
···µ
2n
, (3.6.11)
for an arbitrary n d/2, where
Γ
µ
1
···µ
2n
χ
(
x
)
= ε
µ
1
···µ
2n
(
1
)
2n
i =1
x
µ
i
/a
χ
(
x
)
, (3.6.12)
C
µ
1
···µ
2n
= η
µ
1
···η
µ
2n
C
µ
1
···C
µ
2n
sym
. (3.6.13)
The spin flavor interpretations of these operators read
Γ
µ
1
···µ
2n
γ
µ
1
···γ
µ
2n
ξ
µ
1
··· ξ
µ
2n
, (3.6.14)
C
µ
1
···µ
2n
γ
µ
1
···γ
µ
2n
1
, (3.6.15)
W
µ
1
···µ
2n
st
1
i
n
ξ
µ
1
··· ξ
µ
2n
, (3.6.16)
up to discretization terms. These new mass terms are Hermitian as well, that is
M
= M , M W
µ
1
···µ
2n
st
, (3.6.17)
where
(
x
)
=
(
1
)
d
µ=1
x
µ
/a
. (3.6.18)
This allows us to introduce a very general mass term of the form
W
st
=
d/2
n=1
µ
n
r
µ
n
a
1
+ W
µ
n
st
(3.6.19)
with generalized Wilson-parameters
r
µ
n
0. Here the sum is over all multi-indices
µ
n
=
(
µ
1
, . . . , µ
2n
)
with 1
µ
i
d
for all
i
with 1
i
2
n
. We remark, that in
Eq.
(3.6.19)
not all of the possible combinations of mass terms are useful in practical
applications. To reproduce Adams staggered Wilson term in
d =
4 dimensions as
37
3 Staggered Wilson fermions
given in Eq. (3.6.1), we set r
1234
= r > 0 and r
µ
n
= 0 otherwise.
In Chapter 8 we deal with the
d =
2 case. We note that here the definition is
essentially unique and the mass term takes the form
W
st
=
r
a
1
+ W
12
st
, (3.6.20)
which reduces the number of staggered flavors from two to one.
We note that all possible
W
st
terms, like the Wilson term
W
w
, break chiral sym-
metry. Moreover, if too many of the symmetries of staggered fermions are broken,
some mass terms may give rise to additional counterterms, cf. Ref. [69].
38
C
C 
Staggered Wilson fermions and usual Wilson fermions share several technical
properties. Both are constructed by adding a momentum-dependent term to a
fermion action suffering from fermion doubling. This additional term breaks chiral
symmetry explicitly, resulting in an additive mass renormalization and the need
for fine-tuning the bare mass parameter.
Besides the technical aspects, comparing the computational properties of differ-
ent lattice fermion actions is important for practical applications. As the accuracy
of physical predictions in Monte Carlo simulations is limited by the available com-
putational resources, the question of computational efficiency is then of high
importance. In this chapter, we deal with the problem of quantifying and com-
paring the computational efficiency of two-flavor staggered Wilson fermions with
the one of usual Wilson fermions, where some of the results
1
were previously pre-
sented in Refs. [
4
,
7
,
8
]. We try to answer the question if staggered Wilson fermions
have a computational advantage, potentially bringing down the enormous costs of
state-of-the-art numerical simulations.
4.1 Introduction
Staggered Wilson fermions, like usual staggered fermions, are formulated using
one-component spinor fields. One might, thus, hope for a better computational
efficiency of staggered Wilson fermions compared to Wilson fermions due to a
smaller fermion matrix and a reduced condition number. The prospect of a kernel
operator with increased computational efficiency is especially interesting in the
context of the overlap and domain wall fermion formulations. These formulations
1
Parts of the discussion based on and some figures reprinted from D. H. Adams, D. Nogradi,
A. Petrashyk, and C. Zielinski, “Computational efficiency of staggered Wilson fermions: A first
look, PoS
LATTICE2013
(2014) 353,
arXiv:1312.3265 [hep-lat]
. Copyright owned by the
authors under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives
4.0 International License (CC BY-NC-ND 4.0).
39
4 Computational efficiency
respect chiral symmetry, but come at the price of high computational costs. This
limits their practical applicability despite many attractive theoretical properties.
In this chapter, we evaluate the computational efficiency of Adams two-flavor
staggered Wilson fermions. As the performance of overlap and domain wall fermi-
ons is tightly connected to the computational properties of their kernel operator,
this is also interesting with respect to our discussion of staggered overlap fermions
in Chapter 6 and 7 and staggered domain wall fermions in Chapter 8. We follow
our report given in Ref. [
4
] together with the updates given in Refs. [
7
,
8
], where we
discuss our results for the computational efficiency in a well-defined benchmark
setting.
We compare the computational costs of inverting the Dirac matrix on a source
for staggered Wilson fermions and usual Wilson fermions in quenched lattice
quantum chromodynamics. We do this while either keeping the physical volume
or the lattice spacing fixed. These inversions are typically the most time-consuming
part in realistic simulations and are, thus, expected to be a good proxy for studying
the computational performance. The resulting large and sparse linear systems
are solved using iterative methods, such as the conjugate gradient method [
83
,
84
]. We characterize the computational efficiency as the ratio of the computation
time for both fermion formulations. For this comparison to be meaningful, the
benchmarking has to be done at fixed values of some physical quantity, where in
general the measured efficiency depends on this fixed quantity. A natural candidate
for this is the pion mass
m
π
, which is easy to measure and tune with high accuracy.
This chapter is organized as follows. In Sec. 4.2, we discuss our strategy to quan-
tify the computational efficiency of staggered Wilson fermions and derive some
theoretical estimates. In Sec. 4.3, we present our numerical study and critically dis-
cuss our results. In Sec. 4.4, we characterize the memory bandwidth requirements
of staggered Wilson fermions by analyzing the arithmetic intensities of several
lattice fermion formulations. Finally, in Sec. 4.5 we summarize and conclude our
study.
4.2 Theoretical estimates
As we already pointed out, the computational costs of realistic simulations are
dominated by inverting the Dirac operator
D D
m
q
on a source
χ
, namely
solving the linear system
DΨ = χ
. As these linear systems are usually very large,
40
4.2 Theoretical estimates
one uses iterative solvers to find a numerical solution for
Ψ
. A traditional approach
to this is the use of the conjugate gradient method for the normal equation, where
one solves the system
D
D Ψ = χ
0
(4.2.1)
with
χ
0
D
χ
instead. We form these normal equations with the Hermitian and
positive (semi-)definite operator
D
D
to ensure the applicability of the conjugate
gradient method, see Ref. [85].
In the following, we want to derive a theoretical estimate for solving Eq.
(4.2.1)
with the conjugate gradient method. First we note that the cost can be characterized
by
cost =
(
#iterations
)
× (cost per iteration). (4.2.2)
For most iterative methods, such as the conjugate gradient method, the cost per
iteration is dominated by the cost of the matrix-vector multiplications. This gives
us
cost
(
#iterations
)
×
(
cost mat.-vec. mult.
)
. (4.2.3)
If we now take the ratio for the computational costs of Wilson and staggered Wilson
fermions, we find
cost
w
cost
sw
=
(
#iterations
)
w
(
#iterations
)
sw
×
(
cost mat.-vec. mult.
)
w
(
cost mat.-vec. mult.
)
sw
. (4.2.4)
When considering the fermion propagator, an additional factor of four should be
included on the right hand side because in the staggered Wilson case one has four
times fewer sources due to a reduced number of spinor components. To get an
estimate for the overall cost ratio of both fermion formulations, we have to estimate
both ratios on the right hand side of Eq.
(4.2.4)
. The first ratio in general depends
on the iterative solver and the fixed physical quantity, in our case the pion mass
m
π
. The second ratio depends on size, structure and sparsity of the corresponding
fermion matrices.
4.2.1 Number of iterations
We begin by further analyzing the first ratio in Eq.
(4.2.4)
. When using the conjugate
gradient method for a specified small residual ε, we make use of the relation
#iterations
κ log
2
ε
, (4.2.5)
41
4 Computational efficiency
where
κ =
|
λ
max
/λ
min
|
is the condition number of
D
D
, see e.g. Ref. [
85
]. Hence
one can approximate the ratio for the number of iterations by
(
#iterations
)
w
(
#iterations
)
sw
κ
w
κ
sw
, (4.2.6)
when keeping
ε
fixed. To verify this, we also numerically determine this ratio. In
the free-field case the ratio in Eq.
(4.2.6)
can be computed analytically for the same
bare mass m
q
for both formulations. We find
lim
m
q
0
κ
0
w
κ
0
sw
=
8
5
3.6, (4.2.7)
where κ
0
κ
0
m
q
refers to the free-field condition number at bare mass m
q
.
4.2.2 Cost of matrix-vector multiplications
Let us now analyze the second ratio on the right hand side of Eq.
(4.2.4)
. We es-
timate the costs for the matrix-vector multiplications from FLOP (floating-point
operations) counts. We find
(
cost mat.-vec. mult.
)
w
(
cost mat.-vec. mult.
)
sw
(
FLOPs
)
w
(
FLOPs
)
sw
,
4 ×
(
FLOPs/site
)
w
(
FLOPs/site
)
sw
=
4 × 1392 FLOPs/site
1743 FLOPs/site
3.2. (4.2.8)
Here the factor of four appears due to the four-component spinors in Wilsons
formulation, 1392 is the number of FLOPs per site needed for the application of
Wilsons Dirac operator and 1743 is the number of FLOPs per site for the staggered
Wilson Dirac operator. Note that Eq.
(4.2.8)
is an estimate and besides the number
of floating point operations also other factors come into play such as the computa-
tional intensities, which we discuss in Sec. 4.4.
Alternatively one can estimate the ratio in Eq.
(4.2.8)
with the approach of
Ref. [
70
], which we now briefly review. In practice Wilson fermions are used with the
spin-projection trick [
86
], effectively reducing the number of spinor components
from four to two. The Wilson action couples each lattice site to 2
×
8 other sites,
42
4.3 Numerical study
0
0.05
0.1
0.15
0.2
−0.82 −0.8 −0.78 −0.76 −0.74 −0.72 −0.7 −0.68
Pion mass m
π
2
Quark bare mass m
q
Pion mass m
π
at β = 6, 16
3
x 32 lattice
Wilson, linear fit
Wilson
Stag. Wilson, linear fit
Stag. Wilson
Figure 4.3.1:
Squared pion mass
m
2
π
as a function of the bare quark mass
m
q
with a
linear fit through the data points with 0.05 m
2
π
.
where the factor of two represents the effective number of spinor components and
8
=
4
×
2 comes from a forward and a backward hop in each space-time direction.
Staggered Wilson fermions are formulated with one-component spinors, but the
staggered Wilson term couples each site to all 2
4
=
16 corners of a four-dimensional
hypercube. This means that the staggered Wilson action couples each lattice site
to 8 + 16 nearby sites. We can now alternatively estimate the ratio by
4 × (8 × 2)
8 + 16
2.7, (4.2.9)
where we included the above-mentioned factor of four. This estimate is smaller
than the one we gave in Eq.
(4.2.8)
as we neglected here the cost of spin decompo-
sition and reconstruction for the spin projection trick.
In summary we conservatively estimate
(
cost mat.-vec. mult.
)
w
(
cost mat.-vec. mult.
)
sw
2 3 (4.2.10)
for the cost-ratio for the matrix-vector multiplications, in agreement with Ref. [
71
].
43
4 Computational efficiency
0
0.05
0.1
0.15
0.2
0.25
−0.82 −0.8 −0.78 −0.76 −0.74 −0.72 −0.7 −0.68
Pion mass m
π
2
Quark bare mass m
q
Pion mass m
π
at β = 6, 20
3
x 40 lattice
Wilson, linear fit
Wilson
Stag. Wilson, linear fit
Stag. Wilson
Figure 4.3.2:
Squared pion mass
m
2
π
as a function of the bare quark mass
m
q
with
a linear fit through the data points with 0.05 m
2
π
.
4.3 Numerical study
We did a numerical study with quenched QCD configurations on a 16
3
×
32 lattice
at
β =
6, a 20
3
×
40 lattice at
β =
6 and a 20
3
×
40 lattice at
β =
6
.
136716 with 200
configurations each. The 20
3
×
40 lattices are chosen in such a way to either keep
the lattice spacing (
β =
6) or the physical volume (
β =
6
.
136716) fixed compared to
the 16
3
×
32 lattice. For the simulations we used the Chroma/QDP software package
[87], which we extended by implementing staggered Wilson fermions.
While the use of preconditioning can speed up the use of Wilson fermions by
roughly a factor of two and preconditioning is in principle also possible for stag-
gered Wilson fermions, our study here deals with the unpreconditioned base case.
4.3.1 Results
In Figs. 4.3.1, 4.3.2 and 4.3.3, we find the squared pion mass
m
2
π
as a function of
the bare quark mass
m
q
for the different ensembles. The relation between the
pion mass and bare quark mass for staggered Wilson fermions was previously also
presented in Refs. [4, 88].
For the determination of the pion mass we follow the well-known standard
procedures, see Ref. [
17
]. By inspecting the effective mass plot of the propagator,
we fix a suitable fit range where the contributions of excited states is negligible and
44
4.3 Numerical study
0
0.05
0.1
0.15
0.2
0.25
−0.74 −0.72 −0.7 −0.68 −0.66 −0.64 −0.62
Pion mass m
π
2
Quark bare mass m
q
Pion mass m
π
at β = 6.14, 20
3
x 40 lattice
Wilson, linear fit
Wilson
Stag. Wilson, linear fit
Stag. Wilson
Figure 4.3.3:
Squared pion mass
m
2
π
as a function of the bare quark mass
m
q
with
a linear fit through the data points with 0.05 m
2
π
0.15.
the signal-to-noise ratio is acceptable. We then do an uncorrelated fit of the pion
propagator over the previously determined range using the function
p
(
n
t
)
= A cosh
m
π
n
t
1
2
N
t
(4.3.1)
by minimization of
χ
2
, where
t = n
t
a
and the fit quality varies in the range 0
.
01
.
χ/d.o.f. .
0
.
75. Here the fit parameter
A
is the amplitude,
m
π
the pion mass and
N
t
{
32, 40
}
is the extent of the corresponding lattice in temporal direction. The
errors are combined statistical and systematic errors and are determined with the
jackknife method [
89
91
], where we estimated the systematic error by varying the
range of the fit.
In agreement with chiral perturbation theory, we find an approximately linear
relationship between
m
2
π
and
m
q
. Moreover, the lowest achievable pion mass
for Wilson and staggered Wilson fermions is comparable. We also note that the
additive mass renormalization is smaller for staggered Wilson fermions compared
to Wilson fermions. This could be a (somewhat weak) indicator for a computational
advantage in the construction of overlap and domain wall fermions. Knowing
the relation between the pion mass and the bare quark mass, for a given value
of
m
π
we can use the respective bare mass
m
q
for Wilson and staggered Wilson
fermions. In the following, we fix five values of the pion mass, namely
m
2
π
{
0.04, 0.07, 0.10, 0.13, 0.16
}
, as in general the efficiency is a function of the fixed
45
4 Computational efficiency
0
2
4
6
8
10
12
14
16
18
1.9 1.95 2 2.05 2.1 2.15 2.2 2.25
Number of configurations
(κ
w
/ κ
sw
)
1/2
Distribution of condition number ratios
Figure 4.3.4:
Distribution of
κ
w
/κ
sw
for
m
2
π
=
0
.
10 on the 20
3
×
40 lattice at
β =
6.
0
0.5
1
1.5
2
2.5
3
0.04 0.06 0.08 0.1 0.12 0.14 0.16
Wilson / stag. Wilson
Pion mass m
π
2
Cond. num. vs. CG iterations (averaged ratios, point sources)
κ
1/2
ratio
CG ratio, ε = 10
−06
CG ratio, ε = 10
−10
CG ratio, ε = 10
−14
(a) Point sources
0
0.5
1
1.5
2
2.5
3
0.04 0.06 0.08 0.1 0.12 0.14 0.16
Wilson / stag. Wilson
Pion mass m
π
2
Cond. num. vs. CG iterations (averaged ratios, wall sources)
κ
1/2
ratio
CG ratio, ε = 10
−06
CG ratio, ε = 10
−10
CG ratio, ε = 10
−14
(b) Wall sources
Figure 4.3.5:
Averages for the ratios of the number of iterations and
κ
w
/κ
sw
as a
function of the squared pion mass m
2
π
on a 16
3
× 32 lattice at β = 6.
physical parameter.
Using Chroma, we calculated the lowest and highest eigenvalue on each config-
urations to determine the conditions number
κ
of
D
D
. With respect to Eq.
(4.2.6)
,
the ratio of the condition numbers is an important factor in judging the computa-
tional efficiency of staggered Wilson fermions. In Fig. 4.3.4, we give an example for
the distribution of
κ
w
/κ
sw
on one of our ensembles. As one can see, the square
root of the ratio is of order
O
(
2
)
and we conclude that the condition number of
Wilson fermions in typically a factor
O
(
4
)
higher compared to staggered Wilson
fermions.
In Figs. 4.3.5, 4.3.6 and 4.3.7, we find the results for the ratio of the number of
46
4.3 Numerical study
0
0.5
1
1.5
2
2.5
3
0.04 0.06 0.08 0.1 0.12 0.14 0.16
Wilson / stag. Wilson
Pion mass m
π
2
Cond. num. vs. CG iterations (averaged ratios, point sources)
κ
1/2
ratio
CG ratio, ε = 10
−06
CG ratio, ε = 10
−10
CG ratio, ε = 10
−14
(a) Point sources
0
0.5
1
1.5
2
2.5
3
0.04 0.06 0.08 0.1 0.12 0.14 0.16
Wilson / stag. Wilson
Pion mass m
π
2
Cond. num. vs. CG iterations (averaged ratios, wall sources)
κ
1/2
ratio
CG ratio, ε = 10
−06
CG ratio, ε = 10
−10
CG ratio, ε = 10
−14
(b) Wall sources
Figure 4.3.6:
Averages for the ratios of the number of iterations and
κ
w
/κ
sw
as a
function of the squared pion mass m
2
π
on a 20
3
× 40 lattice at β = 6.
0
0.5
1
1.5
2
2.5
3
0.04 0.06 0.08 0.1 0.12 0.14 0.16
Wilson / stag. Wilson
Pion mass m
π
2
Cond. num. vs. CG iterations (averaged ratios, point sources)
κ
1/2
ratio
CG ratio, ε = 10
−06
CG ratio, ε = 10
−10
CG ratio, ε = 10
−14
(a) Point sources
0
0.5
1
1.5
2
2.5
3
0.04 0.06 0.08 0.1 0.12 0.14 0.16
Wilson / stag. Wilson
Pion mass m
π
2
Cond. num. vs. CG iterations (averaged ratios, wall sources)
κ
1/2
ratio
CG ratio, ε = 10
−06
CG ratio, ε = 10
−10
CG ratio, ε = 10
−14
(b) Wall sources
Figure 4.3.7:
Averages for the ratios of the number of iterations and
κ
w
/κ
sw
as
a function of the squared pion mass
m
2
π
on a 20
3
×
40 lattice at
β =
6.136716.
conjugate gradient iterations as given in Eq.
(4.2.6)
. In the figures, we show both
our theoretical estimate given by
κ
w
/κ
sw
and the average measured ratios for the
actual number of iterations needed as a function of the pion mass. We did this
measurements for different values of the target residual
ε
to check for a possible
dependence. We also used two different kinds of sources, namely a point source
and a wall source.
First we note that the choice of the source has no noticeable effect on the relative
computational performance as the figures are nearly indistinguishable. Remark-
ably the dependence of the ratios on
m
π
is also weak, although the number of
conjugate gradient iterations grows strongly with decreasing pion masses. We can
see that our rough theoretical estimate
κ
w
/κ
sw
is a decent approximation for
47
4 Computational efficiency
the actual measured ratios. Moreover, for smaller values of ε the measured ratios
become closer to our estimate as expected. In all cases we see that Wilson fermions
need roughly
O
(
2
)
more iterations in the application of the conjugate gradient
method, which is in agreement with Ref. [71].
If we increase the physical volume in our simulation, i.e. when we compare the
16
3
×
32 lattice at
β =
6 with the 20
3
×
40 lattice at
β =
6, we find an improvement
at small
m
π
, but otherwise almost unchanged results. If, on the other hand, we
decrease the lattice spacing, i.e. when we compare the 16
3
×
32 lattice at
β =
6
with the 20
3
×
40 lattice at
β =
6
.
136716, we see a significant improvement in
efficiency over the whole range of
m
π
. As the staggered Wilson Dirac operator is
more sensitive to gauge fluctuations through its four-hop terms, we would expect
that on smoother gauge configurations its convergence properties improve more
compared to the Wilson Dirac operator.
4.3.2 Overall speedup
Combining the measured ratios from the previous discussion together with Eq.
(4.2.10)
,
the product in Eq. (4.2.4) takes the form
cost
W
cost
SW
=
(
#iterations
)
w
(
#iterations
)
sw
×
(
cost mat-vec mult.
)
w
(
cost mat-vec mult.
)
sw
2 × (2–3) = 4–6. (4.3.2)
This means that in our setting staggered Wilson fermions are a factor of 4
6 more
efficient than usual Wilson fermions for inverting the Dirac operator on a source.
When computing the quark propagator, an additional factor of four should be
included in Eq.
(4.3.2)
as for Wilson fermions one needs four times as many sources.
Before we move on with the discussion of arithmetic intensities, we add some
critical remarks for the interpretation of the estimate given in Eq.
(4.3.2)
. Our cost
ratio is for the case of inverting the Dirac operator on a source and does not include
the setup phase. For staggered Wilson fermions, one would typically prepare the
averaged link products in the staggered Wilson term before applying the conjugate
gradient method. The factor in Eq.
(4.3.2)
would therefore be a bit lower when
taking the cost for this setup phase into account. Also, if one intends to do hadron
spectroscopy with staggered Wilson fermions as described in Chapter 5, one has
to take the additional number of sources needed into account, effectively reducing
the computational advantage.
Moreover, when doing lattice field theoretical computations on high-perfor-
48
4.4 Arithmetic intensities
mance computing clusters, communication over the network can become a bottle-
neck for the performance. As large-scale simulations of lattice quantum chromody-
namics usually make use of a domain-decomposition approach and the staggered
Wilson term connects each site with all the corners of the surrounding hypercube
via four-hop terms, it is expected that the resulting network traffic will be signifi-
cant. How severely this limits the performance of staggered Wilson fermions in
large-scale simulations has yet to be seen. Until then, Eq.
(4.3.2)
should be taken
as an estimate for the achievable performance gain on shared memory machines.
4.4 Arithmetic intensities
On modern computer architectures the performance of lattice field theoretical
simulations is often not limited by the floating point throughput, but by limited
memory bandwidth. A measure for the bandwidth requirements relative to the
number of floating point operations (FLOPs) is the so called arithmetic intensity
I =
FLOPs
Memory transactions in byte
. (4.4.1)
If for a given application this ratio is lower than what the respective hardware can
provide, characterized by
I
hw
=
max. FLOPs / s
max. memory bandwidth in byte / s
, (4.4.2)
its performance is limited by the memory bandwidth. As a result, the computing
cores spend some time with idling while waiting for the completion of memory
access operations. On the other hand the performance of problems with high
arithmetic intensity is limited by the floating point throughput of the hardware.
Simulations in the context of lattice quantum chromodynamics tend to have
a rather low computational intensity, i.e. a relatively high number of memory
transactions per floating point operation. If optimized properly, lattice codes can
archive a high sustained performance on modern architectures. As the expected
performance depends on the arithmetic intensity of the lattice fermion formulation,
it is worth having a closer look.
In the following, we determine the arithmetic intensities for the application of
several lattice Dirac operators in the setting of quantum chromodynamics, where
we focus on the hopping-terms and exclude the trivial diagonal mass term. Within
49
4 Computational efficiency
derivations we abbreviate FLOPs with
F
. Before we begin, we note that for a com-
plex addition we need 2
F
and for a complex multiplication we need 6
F
. To multiply
a complex
SU
(
3
)
matrix with a complex three-component vector, we need 66
F
. Re-
garding memory transactions, we assume single-precision floating point numbers,
so-called floats, which are 32
bit
or 4
byte
. Where convenient, we abbreviate floats
by
fl
. Depending on the application and hardware, in practice one might also
use double-precision floating point numbers with 64
bit
or 8
byte
, lowering the
arithmetic intensity
2
by a factor of two (see Ref. [
92
]). In some cases even mixed-
precision is used, consisting of a combination of double-precision, single-precision
and half-precision floating point numbers.
4.4.1 Wilson fermions
We begin with Wilson fermions as defined in Subsec. 2.2.6. We note that some
fixed terms in the action can be precomputed, namely terms like
1
± γ
µ
, to save
arithmetic operations. We also have to take into account that a competitive im-
plementation of Wilson fermions employs the spin-projection trick [
86
], which
reduces the number of spin components by a factor of two.
Floating point operations.
The spin projection takes 12
×
2
×
2
F =
96
F
(twelve
operations
×
four directions
×
forward/backward) and for the link multiplications
we need 66
×
2
×
4
×
2
F =
1056
F
(matrix-vector multiplication
×
two spin compo-
nents
×
four directions
×
forward/backward). Accumulating the resulting vector
takes 12
×
2
×
7
F =
168
F
(twelve components
×
cost per complex addition
×
adding
eight vectors). In total we find that Wilson fermions take 1320
F/site
. This result is
in agreement with Refs. [9294].
Memory access operations.
We need to read one spinor per direction, i.e. 12
×
2
×
4
×
2
fl =
192
fl
(twelve complex numbers per spinor
×
floats per complex number
×
four directions
×
forward/backward) and one 3
×
3 gauge link per direction, i.e.
9
×
2
×
4
×
2
fl =
144
fl
(nine elements
×
floats per complex number
×
four directions
×
forward/backward). Writing the resulting spinor takes 12
×
2
fl =
24
fl
. In total
we find 360
fl/site =
1440
byte/site
of memory access operations per lattice site.
This result is also in agreement with Refs. [92, 94].
2
While in this case the arithmetic intensity decreases, the number of CPU cycles can potentially
increase.
50
4.4 Arithmetic intensities
Arithmetic intensity. We find
I
w
=
1320 FLOPs/site
1440 byte/site
0.92 FLOPs/byte (4.4.3)
for the arithmetic intensity of Wilson fermions.
4.4.2 Staggered fermions
We are repeating the previous exercise now for usual staggered fermions, which
we discussed in Subsec. 2.2.7.
Floating point operations.
For the link multiplications we have 66
×
4
×
2
F =
528
F
(matrix-vector multiplication
×
four directions
×
forward/backward). Accumulat-
ing the resulting vector takes 3
×
2
×
7
F =
42
F
(three color components
×
cost per
complex addition
×
adding eight vectors). In total we find that usual staggered
fermions need 570 F/site, which is in agreement with Ref. [95].
Memory access operations.
We read one spinor per direction, i.e. 3
×
2
×
4
×
2
fl =
48
fl
(three complex numbers per spinor
×
floats per complex number
×
four directions
×
forward/backward) as well as one gauge link per direction,
i.e. 9
×
2
×
4
×
2
fl =
144
fl
(nine elements
×
floats per complex number
×
four
directions
×
forward/backward). Writing the resulting spinor takes 3
×
2
fl =
6
fl
.
In total we have 198
fl/site =
792
byte/site
of memory access operations per lattice
site, which is also in agreement with Ref. [95].
Arithmetic intensity. We find
I
st
=
570 FLOPs/site
792 byte/site
0.72 FLOPs/byte (4.4.4)
for the arithmetic intensity of staggered fermions.
4.4.3 Staggered Wilson fermions
Compared to usual staggered fermions, staggered Wilson fermions have an addi-
tional term in the action. In the following, we only analyze the additional contri-
butions from the staggered Wilson term and then add the respective counts to the
numbers of usual staggered fermions.
51
4 Computational efficiency
Floating point operations.
Applying the
C
operator takes 66
×
16
F +
3
×
2
×
15
F =
1146
F
(matrix-vector multiplication
×
number of hybercube corners
+
three color
components
×
cost per complex addition
×
adding sixteen vectors). Applying
Γ
55
and
η
5
takes
(
3 + 6
)
F
. The
r
1
term can be combined with the diagonal mass term
m
1
and, thus, will be ignored as in the preceding calculations. Accumulating the
resulting vector takes 3
×
2
×
1
F =
6
F
(three color components
×
cost per complex
addition
×
adding two vectors). In total we find that staggered Wilson fermions
take an additional 1161
F/site
compared to staggered fermions, i.e. 1731
F/site
in
total. Note that when the diagonal
(
m + r
)
1
term is included we need additional
12 F, i.e. 1743 F/site in agreement with Ref. [4].
Memory access operations.
We need to read one additional spinor per hypercube
corner, i.e. 3
×
2
×
16
fl =
96
fl
(three complex numbers per spinor
×
floats per
complex number
×
number of corners) and one gauge link per hypercube corner,
i.e. 9
×
2
×
16
fl =
288
fl
(9 elements
×
floats per complex number
×
number of
corners) for the
C
operator. Write the resulting spinor takes 3
×
2
fl =
6
fl
. This
means that we have an additional 390
fl/site =
1560
byte/site
, i.e. in total we find
588 fl/site = 2352 byte/site for staggered Wilson fermions.
Arithmetic intensity. We find
I
stw
=
1731 FLOPs/site
2352 byte/site
0.74 FLOPs/byte (4.4.5)
for the arithmetic intensity of staggered Wilson fermions.
4.4.4 Summary
We summarize our findings in Table 4.4.1, where we added Wilson fermions with
a clover term (see e.g. Ref. [
16
]) as well as HISQ [
96
] (highly improved staggered
quarks) and asqtad [
97
] (
a
-squared tadpole improved) staggered fermions from the
literature. We note that all staggered formulations have a very similar arithmetic
intensity, while staggered Wilson fermions have slightly lower memory bandwidth
requirements compared to the others. This means, that despite the large number
of memory access operations per site needed for the application of the staggered
Wilson Dirac operator, we do not expect a more severe memory bandwidth bottle-
neck.
52
4.5 Conclusions
Formulation Operations I/O Intensity Source
Wilson 1320 1440 0.92 Here and Refs. [9294]
Wilson + Clover 1824 1728 1.06 Ref. [94]
Staggered 570 792 0.72 Here and Ref. [95]
HISQ / asqtad 1146 1560 0.73 Ref. [94]
Staggered Wilson 1731 2352 0.74 Here and Ref. [4]
Table 4.4.1:
Arithmetic intensities of several lattice Dirac operators. Operations are
in units of FLOPs / site, I/O in bytes / site and Intensity in FLOPs / byte.
One possibility to increase arithmetic intensities and to lower memory band-
width requirements for all fermion formulations is to only load the first two rows of
each
SU
(
3
)
matrix and then reconstruct the last one by using unitarity. This lowers
memory bandwidth requirements at the cost of additional FLOPs, thus increasing
arithmetic intensity.
4.5 Conclusions
We estimated the speedup factor of staggered Wilson fermions to be 4
6 compared
to usual Wilson fermions for inverting the Dirac matrix on a source. However, it is
important to keep the critical remarks from Subsec. 4.3.2 in mind when interpreting
this result.
To allow for an easier comparison, we discussed here the case of kernel operators
which are both unpreconditioned and unimproved. While for Wilson fermions
many preconditioned and improvement variants are known and in use, there are
many possibilities to further improve the computational efficiency of staggered
Wilson fermions as well. One could replace the usual staggered part of staggered
Wilson fermions by the HISQ action to reduce
O
a
2
effects and use link smearing.
While a certain form of a clover term for
O
(
a
)
improvement was studied in Ref. [
82
],
there is also the possibility for the introduction of other clover terms.
Finally, comparing the arithmetic intensities of the different fermion formula-
tions in Sec. 4.4, we note that the memory bandwidth requirements of staggered
Wilson fermions are slightly lower compared to other staggered formulations. As
memory bandwidth is usually a limiting factor for the performance of lattice codes,
53
4 Computational efficiency
this is important in the context of writing highly optimized implementations of
the Dirac operator.
54
C
P  
One of the traditional applications of lattice gauge theory is hadron spectroscopy.
In this setting not only computational efficiency, but also the usability of the for-
malism is important. To show the feasibility of doing spectroscopy with staggered
Wilson fermions, in this chapter we discuss the case of pions, i.e. pseudoscalar
mesons.
For usual staggered fermions the four tastes give rise to sixteen pions. In the
following we study the pion spectrum of staggered fermions when turning Adams
two-flavor staggered Wilson term on. The effect of this term is that two of the four
tastes acquire a mass of
O
a
1
and become heavy. As a result, we find that eight of
the pions become heavy as well, while the other eight remain light. In the physical
light part of the spectrum we find the degenerate two-flavor pion-triplet and the
flavor-singlet
η
meson with each coming in two copies. The
η
meson remains light
as we omit the disconnected piece of the propagator in our study. Some of the
results discussed here were previously presented in Refs. [9, 10].
5.1 Introduction
From a practitioners point of view, the usability of a lattice formulation is important.
To show that staggered Wilson fermions are not at a disadvantage compared to
other fermion formulations, we illustrate spectroscopy calculations as one of the
traditional applications of lattice quantum chromodynamics (QCD). Like for other
staggered formulations, spin and flavor degrees of freedom are distributed over
different lattice sites. We will see, that one can easily adapt the spectroscopy
methods and operators of usual staggered fermions to the two flavor case, as we
are going to illustrate for the case of pions.
In this work we deal with Adams original two flavor proposal [
44
,
45
], where the
action reads
S
sw
= ¯χ
D
st
+ m
q
+ W
st
χ, W
st
=
r
a
(
1
Γ
55
Γ
5
)
(5.1.1)
55
5 Pseudoscalar meson spectrum
and
D
st
= η
µ
µ
refers to the staggered Dirac operator,
m
q
to the bare quark mass,
W
st
to the staggered Wilson term as defined in Sec. 3.4,
a
to the lattice spacing and
r >
0 is the equivalent of the Wilson parameter. Here and in the following, the
ξ
µ
matrices refer to a representation of the Dirac algebra in flavor space and in figures
we use the shorthand notation ξ
µν
ξ
µ
ξ
ν
.
This chapter is organized as follows. In Sec. 5.2, we discuss pseudoscalar mesons
in the framework of usual staggered fermions and generalize the known methods
to the case of staggered Wilson fermions. In Sec. 5.3, we discuss our numerical
results and compare them to our theoretical predictions. Finally, in Sec. 5.4 we
summarize our results and end with some concluding remarks.
5.2 Pseudoscalar mesons
For usual staggered fermions, the meson and baryon rest-frame operators are
known [
98
] and we can adapt these operators to the two-flavor staggered Wilson
case. However, as two of the four tastes become heavy after the introduction of
the staggered Wilson term, the physical interpretation of these operators change.
In the following, we will discuss the case of pseudoscalar mesons both from a
theoretical and numerical point of view. We first begin with reviewing the case of
usual staggered fermions, before we move on to staggered Wilson fermions.
Staggered fermions have four (quark) tastes, giving rise to sixteen pseudoscalar
mesons. They have a spin flavor structure γ
5
ξ
F
with
ξ
F
1
, ξ
5
, ξ
µ
, ξ
µ
ξ
5
, ξ
µ
ξ
ν
. (5.2.1)
On the lattice the states fall into eight irreducible representations of the lattice
timeslice group [98], namely
ξ
4
ξ
5
, ξ
5
, ξ
i
ξ
5
, ξ
i
ξ
j
, ξ
i
, ξ
i
ξ
4
,
1
, ξ
4
. (5.2.2)
Here the
γ
5
ξ
5
state corresponds to the Goldstone pion for the axial symmetry at
vanishing
m
q
. While for staggered fermions the continuum
SU
(
4
)
flavor symmetry
is broken down to a discrete subgroup, by doing a joint expansion in
m
q
and
a
2
one
can show that the pion masses are
SO
(
4
)
symmetric up to leading discretization
errors, see Ref. [99].
Within these representations we find two distinct types of states: while some
states propagate with a
(
1
)
t
-factor in time, others do not. The timeslice operators,
56
5.2 Pseudoscalar mesons
Flavor matrix ξ
F
Composition Particles
ξ
k
ξ
4
qσ
k
q π triplet
ξ
i
ξ
j
ξ
5
qq η singlet
1
Table 5.2.1:
Pseudoscalar meson spectrum for staggered Wilson fermions (
i
,
j
and
k mutually distinct).
in general, excite both these kinds of states, i.e. if an operator couples to
γ
S
ξ
F
,
then it also couples to
γ
4
γ
5
γ
S
ξ
4
ξ
5
ξ
F
. In the case of pseudoscalar mesons we have
γ
S
{
γ
5
, γ
4
γ
5
}
. The fact that we excite two states simultaneously does not pose a
problem as both states can be easily separated by their distinct time-dependence.
We can parametrize the time-time correlation function by
f
R
±
,m
±
(
n
t
)
= R
+
cosh
m
+
n
t
1
2
N
t
+
(
1
)
n
t
R
cosh
m
n
t
1
2
N
t
, (5.2.3)
where
t = n
t
a
and we assume that one-particle states are dominant [
98
]. Here
R
±
and
m
±
are the amplitudes and masses of both respective states and
N
t
refers to
the number of timeslices of the lattice.
After the introduction of the staggered Wilson term, two tastes become heavy
and ξ
F
can be thought of having the structure
ξ
F
light + light light + heavy
heavy + light heavy + heavy
. (5.2.4)
In the continuum limit, the heavy quark content decouples and the “light + light
part of ξ
F
determines the physical interpretation. To give an example, consider
ξ
F
= ξ
1
ξ
2
=
iσ
3
iσ
3
. (5.2.5)
The physically relevant part is
iσ
3
, hence
i ¯χ
(
γ
5
ξ
3
ξ
4
)
χ
corresponds to a
π
0
operator, see also Ref. [9].
After analyzing all states, one finds that for
ξ
F
{
ξ
4
ξ
5
, ξ
i
ξ
5
, ξ
i
, ξ
4
}
there are
only contributions from heavy tastes. For the states with
ξ
F
ξ
5
, ξ
i
ξ
j
, ξ
i
ξ
4
,
1
,
we find a mixture of purely light and purely heavy quark content, where the heavy
57
5 Pseudoscalar meson spectrum
0
0.05
0.1
0.15
0.2
0.25
0 0.005 0.01 0.015 0.02 0.025 0.03
Quark bare mass m
q
Our γ
5
x ξ
5
Our γ
5
x ξ
i5
Our γ
5
x ξ
i4
Our γ
5
x ξ
4
Lit. γ
5
x ξ
5
Lit. γ
5
x ξ
i5
Lit. γ
5
x ξ
i4
Lit. γ
5
x ξ
4
(a) Part I
0
0.05
0.1
0.15
0.2
0.25
0 0.005 0.01 0.015 0.02 0.025 0.03
Quark bare mass m
q
Our γ
45
x 1
Our γ
45
x ξ
45
Our γ
45
x ξ
ij
Our γ
45
x ξ
i
Lit. γ
45
x 1
Lit. γ
45
x ξ
45
Lit. γ
45
x ξ
ij
Lit. γ
45
x ξ
i
(b) Part II
Figure 5.3.1:
Comparison between the pseudoscalar meson masses
m
2
m
q
at
r =
0 in our study with the literature values from Ref. [100].
0.1
0.2
0.3
0.4
0.5
0.6
−0.065 −0.06 −0.055 −0.05 −0.045 −0.04 −0.035 −0.03 −0.025 −0.02
Quark bare mass m
q
γ
5
x ξ
5
γ
45
x 1
γ
5
x ξ
i4
γ
45
x ξ
ij
(a) Physical pseudoscalar mesons
0.1
0.2
0.3
0.4
0.5
0.6
−0.065 −0.06 −0.055 −0.05 −0.045 −0.04 −0.035 −0.03 −0.025 −0.02
Quark bare mass m
q
γ
45
x ξ
45
γ
5
x ξ
i5
γ
5
x ξ
4
γ
45
x ξ
i
(b) Heavy pseudoscalar mesons
Figure 5.3.2: Pseudoscalar meson masses m
2
m
q
at r = 0. 1.
contributions decouple in the continuum limit. In total we find that from the
sixteen pseudoscalar mesons for usual staggered fermions only eight remain light
in the presence of the staggered Wilson term. Among these eight light pions we
find two copies of the two-flavor pion triplet, which only differ by their heavy quark
content. We note that the pion triplet for staggered Wilson fermions is degenerated
as discussed in Ref. [
69
]. In addition we find two copies of the flavor-singlet
η
meson. These light states and their physical interpretation are summarized in
Table 5.2.1, where q =
(
q
1
, q
2
)
|
denotes the two physical tastes.
58
5.3 Numerical results
0.4
0.42
0.44
0.46
0.48
0.5
0.52
0.54
0.56
0.58
0.6
10 20 30 40 50
Time slice t
Staggered Wilson fermions, ξ
F
= ξ
13
, β = 6, 16
3
x 64, m
q
= −0.6650
Fit: m = 0.511138 ± 0.001407 ± 0.000469, χ
2
/d.o.f. = 0.4704
Figure 5.3.3:
Effective mass plot
m
eff
(
t
)
together with the fitted mass
m
and its
statistical and systematic error at
r =
1 for the light pseudoscalar
meson with ξ
F
= ξ
1
ξ
3
.
5.3 Numerical results
To illustrate the practicality of spectrum calculations with staggered Wilson fermi-
ons and to verify the results from the previous section, we implemented staggered
Wilson fermions in the Chroma/QDP software package [
87
]. The software suite
comes with an application for spectrum calculations with staggered fermions
1
.
For the pseudoscalar mesons operators we used the operators that are already
implemented and which match the ones used in Ref. [100].
Our quenched QCD study was carried out using 200 gauge configurations at
a coupling of
β =
6 on a 16
3
×
64 lattice, where we fixed the configurations to
Coulomb gauge. On each configuration we evaluated eight staggered wall sources
2
,
i.e. wall sources where within the wall there is a one in a single corner of each of the
spatial cubes and zero everywhere else. By combining them appropriately, we can
project onto the irreducible representations of the timeslice group [
100
]. To extract
the pseudoscalar meson masses, we essentially follow the method described in
Subsec. 4.3.1. To compute effective masses, we can do a minimization of the form
1
The program spectrum_s.
2
Note that this additional number of sources effectively reduces the computational speedup factor
as presented in Chapter 4.
59
5 Pseudoscalar meson spectrum
0
0.05
0.1
0.15
0.2
0.25
0.3
−0.74 −0.73 −0.72 −0.71 −0.7 −0.69 −0.68 −0.67 −0.66
Pseudoscalar meson mass m
2
Quark bare mass m
q
γ
5
x ξ
5
γ
45
x 1
γ
5
x ξ
i4
γ
45
x ξ
ij
Figure 5.3.4:
Pseudoscalar meson spectrum for two-flavor staggered Wilson fermi-
ons (r = 1).
min
R
±
,m
±
F
R
±
,m
±
with
F
R
±
,m
±
=
f
R
±
,m
±
(
n
t
+ 0
)
p
(
n
t
+ 0
)
f
R
±
,m
±
(
n
t
+ 1
)
p
(
n
t
+ 1
)
f
R
±
,m
±
(
n
t
+ 2
)
p
(
n
t
+ 2
)
f
R
±
,m
±
(
n
t
+ 3
)
p
(
n
t
+ 3
)
, (5.3.1)
where
f
R
±
,m
±
(
n
t
)
refers to the fit function defined in Eq.
(5.2.3)
and
p
(
n
t
)
to the
corresponding time-time correlation function. In the final four parameter fit of
Eq.
(5.2.3)
we fixed the fit range from
n
1
=
10 to
n
2
=
25 together with the mirrored
range
N
t
n
2
to
N
t
n
1
for all pions, where
N
t
=
64 is the extent of the lattice
in temporal direction. Correlations are neglected in the
χ
2
-fit. The errors are
estimated in the same way as described in Subsec. 4.3.1.
As a check of the correctness of our method and fitting procedure, we reproduced
the known results from Ref. [
100
] (cf. Ref. [
101
]) for the case of usual staggered
fermions. By setting the staggered Wilson parameter to
r =
0 in Eq.
(5.1.1)
for this
test, we can calculate the pseudoscalar meson spectrum and find an excellent
match with the literature values, see Fig. 5.3.1. In this figure and the following
mass plots, for every pseudoscalar meson mass
m
we add a linear fit of the form
m
2
= c
1
m
q
+ c
2
.
As the negative flavor-chirality species acquire a mass of
O
(
r /a
)
in the presence
60
5.4 Conclusions
of the staggered Wilson term, we can use
r
to control the masses of the heavy
pseudoscalar mesons. We can then consider staggered Wilson fermions with a
small
r
to observe how a mass splitting between the light and heavy pions forms.
In Fig. 5.3.2, we can find the spectrum for
r =
0
.
1, where in the left part of the figure
we can find all light pions and in the right part the heavy pions. Using the same
scale for the vertical axes, one can see how the heavy pseudoscalar mesons are
systematically shifted upwards due to contributions from heavy tastes.
Moving on to staggered Wilson fermions with
r =
1, the initially small mass gap
between the light and heavy pion states becomes large. In Fig. 5.3.3, we can see a
representative effective mass plot for the state with
ξ
F
= ξ
1
ξ
3
, which remains light
after the introduction of the staggered Wilson term in agreement with Table 5.2.1
and the discussion in Sec. 5.2. On the other hand, the heavy states now have a
mass of
m = O
a
1
. As the condition
am
1 is violated, one would not expect
to be able to reliably describe these states on our lattices.
Finally, focusing on the physical (light) part of the pseudoscalar mesons, we find
the resulting spectrum of pseudoscalar mesons for two-flavor staggered Wilson
fermions in Fig. 5.3.4. As predicted by chiral perturbation theory, we find that the
squared pseudoscalar meson masses
m
2
are approximately linear in the bare quark
mass
m
q
. Referring to Table 5.2.1, in the continuum limit the pairs
{
ξ
5
,
1
}
and
ξ
k
ξ
4
, ξ
i
ξ
j
are expected to be degenerated as the heavy quark content decou-
ples. In the present setting with a finite lattice spacing these degeneracies only
hold approximately, but already describe the observed spectrum well. Moreover,
we observe that both pairs separate well. We note that in our calculations the
flavor-singlet
η
remains light due to the omission of the disconnected piece of the
propagator.
5.4 Conclusions
The aim of this study was to demonstrate the feasibility of spectrum calculations
with Adams two-flavor staggered Wilson fermions. To this end we determined how
the pseudoscalar spectrum can be computed and changes with the presence of
the staggered Wilson term.
In the physical part of the spectrum we find two copies for both the degenerate
two-flavor pion-triplet as well as the flavor-singlet
η
meson. We conclude that
two-flavor staggered Wilson fermions work as expected for studying pseudoscalar
61
5 Pseudoscalar meson spectrum
mesons. In future work one can investigate spectroscopy of other hadrons, in
particular also baryons, where our approach is expected to generalize.
62
C
S  
Chiral symmetry plays an important role in understanding the low-energy regime
of quantum chromodynamics, in particular hadron phenomenology. As discussed
in Subsec. 2.2.4, the Nielsen-Ninomiya theorem [
30
33
] complicates the implemen-
tation of chiral symmetry on the lattice. With the overlap construction [
102
108
] this
problem could eventually be overcome and one can formulate a lattice Dirac oper-
ator with exact chiral symmetry [
109
111
]. While the overlap formalism has many
attractive theoretical properties, its practical applications are limited due to the fact
that overlap fermions typically require a factor of O
(
10–100
)
more computational
resources compared to Wilson fermions.
The prospect of a staggered overlap operator becomes interesting as it can po-
tentially bring down the enormous costs of lattice theoretical simulations in which
chiral symmetry is of importance. Usual staggered fermions do not define a suit-
able kernel operator for the overlap formalism due to the fact that the staggered
equivalent of
γ
5
, namely the
Γ
5
γ
5
1
+ O
a
2
operator, does not square to
the identity, i.e.
Γ
2
5
,
1
. As Adams pointed out in Refs. [
44
,
45
,
112
], this problem
can be solved with the help of the staggered Wilson construction, where on the
physical species
Γ
55
can be used for
γ
5
. As
Γ
2
55
=
1
holds exactly, this allows the
construction of staggered variants of overlap and domain wall fermions.
In this chapter, we discuss Neuberger’s overlap fermions in Sec. 6.1 and then
introduce Adams staggered overlap construction in Sec. 6.2. This is followed by a
review of the index theorem in Sec. 6.3. In Sec. 6.4, we verify that the continuum
limit of the index and axial anomaly for the staggered overlap Dirac operator is cor-
rectly reproduced. This is meant as a first step to place staggered overlap fermions
on solid theoretical foundations. Our analytical work was first presented in Ref. [
3
],
where later a more comprehensive discussion was given by Har in Ref. [11].
63
6 Staggered overlap fermions
6.1 Neuberger’s overlap fermions
We begin by discussing the Ginsparg-Wilson relation in Subsec. 6.1.1 and Neu-
bergers overlap construction in Subsec. 6.1.2, following the outline of Ref. [16].
6.1.1 The Ginsparg-Wilson relation
The Nielsen-Ninomiya theorem prevents a straight-forward implementation of
chiral symmetry on the lattice of the form
{
D, γ
5
}
=
0. To circumvent this problem,
Ginsparg and Wilson proposed in Ref. [
109
] to replace the continuum expression
by
{
D, γ
5
}
= aD γ
5
D. (6.1.1)
For this so-called Ginsparg-Wilson relation in Eq.
(6.1.1)
we see that the anticommu-
tator is not vanishing on the lattice, but is instead an order
O
(
a
)
term, eventually
shrinking to zero in the continuum limit. A remarkable observation is now that,
while the continuum form of chiral symmetry is recovered in the
a
0 limit,
even at a finite lattice spacing a Dirac operator
D
satisfying Eq.
(6.1.1)
has an exact
chiral symmetry. However, for a long time no solution to the Ginsparg-Wilson rela-
tion was known and the possibility of finding a closed solution in the interacting
case was doubted. Only much later solutions were explicitly constructed, namely
overlap fermions [108] and fixed points fermions [113].
Assuming that a Dirac operator
D
satisfies the Ginsparg-Wilson relation, one
can define an infinitesimal chiral transformation of the form
δΨ = Ψγ
5
, δΨ = γ
5
(
1
aD
)
Ψ, (6.1.2)
where the corresponding finite transformation takes the form
Ψ Ψe
iαγ
5
, Ψ e
iαγ
5
(
1
aD
)
Ψ (6.1.3)
with
α R
being a continuous parameter. We note that these transformations can
also be written in a symmetric form, where both
Ψ
and
Ψ
transform with a factor
of
γ
5
(
1
aD/2
)
. One can easily verify the invariance of the action
S = ΨD Ψ
under
these chiral transformations, where in the
a
0 limit the continuum expressions
are reproduced.
Now let us impose that
D
is also
γ
5
Hermitian, i.e.
D
= γ
5
Dγ
5
. This ensures that
complex eigenvalues come in conjugated pairs and
det D R
. Another important
64
6.1 Neubergers overlap fermions
consequence is that for an eigenvector
v
λ
of
D
, corresponding to the eigenvalue
λ
,
one can verify that
h
v
λ
, γ
5
v
λ
i
= 0, (6.1.4)
unless
λ R
. Here
h
v, w
i
= v
w
denotes the usual inner product of
C
n
. This
means that only eigenmodes with real eigenvalues can have definite chirality.
Going back to the Ginsparg-Wilson relation in Eq.
(6.1.1)
, we can see that by
multiplying with
γ
5
from either the left or the right side and using
γ
5
Hermiticity,
one finds that
aD
D = D
+ D = D + D
= aDD
, (6.1.5)
so D is a normal operator satisfying
D
, D
= 0. As a consequence one finds
λ + λ
= a λλ
(6.1.6)
for the eigenvalues. Writing the eigenvalues explicitly with real and imaginary part,
this equation takes the form
Re λ
1
a
2
+
(
Im λ
)
2
=
1
a
2
(6.1.7)
and one can see that they lie on a circle in the complex plane, commonly referred
to as the Ginsparg-Wilson circle.
As the Ginsparg-Wilson circle is going through the origin,
D
can have exact zero
modes. If v
0
denotes such a zero mode, one can verify that
Dv
0
= γ
5
Dv
0
= D γ
5
v
0
= 0. (6.1.8)
Thus on the eigenspace for the eigenvalue zero the Dirac operator
D
and the
chirality operator
γ
5
commute. One can now choose the eigenvectors so that they
are simultaneously eigenmodes of both D and γ
5
. We can then write
γ
5
v
0
= ±v
0
, (6.1.9)
where the eigenvalues are
±
1 due to
γ
2
5
=
1
. This means that the zero modes of
D
have definite chirality. If we find a plus sign in Eq.
(6.1.9)
, we say that
v
0
has positive
chirality or call it right-handed, otherwise we say that
v
0
has negative chirality or
call it left-handed.
65
6 Staggered overlap fermions
6.1.2 The overlap operator
While for a long time the search for an explicit solution to the Ginsparg-Wilson
relation was not successful, eventually a solution in the form of the overlap operator
was found. The overlap formalism was developed over the course of several papers
[
102
108
] and finally brought into its present day form by Neuberger [
106
108
]. The
Dirac operator reads
D
ov
=
1
a
1
+ A
A
A
1
2
=
1
a
1
+ γ
5
H /
H
2
=
1
a
(
1
+ γ
5
sign H
)
(6.1.10)
with kernel operator
A = D
w
(
M
0
)
, Hermitian operator
H = γ
5
D
w
(
M
0
)
and
negative mass parameter
M
0
. In the free-field case valid choices are in the range
M
0
(
0, 2r
)
, where
r
denotes the Wilson parameter, to give rise to a one flavor
theory. In some contexts there is an additional rescaling of
D
ov
by some factor
ρ
,
see our discussion in Chapter 8. In principle one can replace the Wilson Dirac
operator
D
w
also by some other suitable
γ
5
Hermitian kernel operator. Using the
fact that
sign
2
H =
1
, one can easily verify that the overlap Dirac operator satisfies
the Ginsparg-Wilson relation in Eq. (6.1.1).
The overlap Dirac operator in Eq.
(6.1.10)
is not ultralocal due to the presence
of the 1
/
H
2
term. Here we call a lattice Dirac operator
D
ultralocal if
l
0 so
that
|
D
(
x, α, a; y, β, b
)|
=
0 for
k
x y
k
1
> l
, where
D
(
x, α, a; y, β, b
)
denotes
the matrix element of
D
connecting the space-time coordinates
x
and
y
with
respective spin components
α
,
β
and color components
a
,
b
. Under some general
assumptions this lack of ultralocality holds in fact for all solutions of the Ginsparg-
Wilson relation [
114
]. However, locality of
D
ov
could be established in Ref. [
115
] in
the form of
|
D
ov
(
x, α, a; y, β, b
)|
λ e
γ/a·
k
xy
k
1
. (6.1.11)
Here
λ
0 and
γ >
0 are constants independent of the gauge field. As a result, the
decay constant
a/γ
vanishes in the continuum limit
a
0 and a local quantum
field theory is recovered.
The presence of the
sign
-function in Eq.
(6.1.10)
makes the application of the
overlap Dirac operator computational expensive. As mentioned earlier, one typ-
ically needs an factor of
O
(
10–100
)
more computational resources compared to
Wilson fermions. As a consequence the use of overlap fermions is limited in practi-
cal applications and a lot of past work deals with the efficient numerical treatment
of the sign-function, see e.g. Refs. [116120].
66
6.2 Staggered overlap fermions
6.2 Staggered overlap fermions
When applying the overlap construction to the staggered Dirac operator, one en-
counters some severe problems. For the resulting operator the remnant chiral
symmetry is lost, the Ginsparg-Wilson relation does not hold and exact zero modes
are absent [45].
These problems can be traced back to the fact that the overlap construction
relies on the property
γ
2
5
=
1
, while this does not hold for the staggered equivalent
Γ
5
γ
5
1
+ O
a
2
. The eigenvalues of
γ
5
are
±
1, while the eigenvalues of
Γ
5
are
distributed throughout the interval
[
1, 1
]
and include zero as can be seen from
the momentum representation
µ
cos
2
p
µ
of
Γ
2
5
in the free-field case [
112
]. Finally,
we note that the operator
Γ
5
D
st
, unlike its Wilson equivalent
γ
5
D
w
, is even lacking
Hermiticity [
112
]. For these reasons the overlap construction with a usual staggered
kernel does not give rise to a proper lattice fermion formulation.
All these problems could be eventually overcome by Adams construction in-
troduced in Refs. [
44
,
45
,
112
]. The idea is that, after the introduction of the stag-
gered Wilson term, one can use the
Γ
55
operator instead of
Γ
5
. Although
Γ
55
has
a spin
flavor interpretation [
42
] of
Γ
55
γ
5
ξ
5
, the staggered Wilson term
projects out the negative flavor-chirality species and we find
Γ
55
Γ
5
+ O
a
2
on
the physical species. Furthermore,
Γ
2
55
=
1
holds exactly and
H
sw
= Γ
55
D
sw
defines
a Hermitian operator.
As a consequence, one can apply the overlap construction to the staggered
Wilson kernel with the help of a simple replacement rule given by
D
w
D
sw
, γ
5
Γ
55
(6.2.1)
as derived in Refs. [
44
,
45
,
112
]. Explicitly, the staggered overlap operator
D
so
reads
D
so
=
1
a
1
+ A
A
A
1
2
=
1
a
1
+ Γ
55
H
sw
H
2
sw
=
1
a
(
1
+ Γ
55
sign H
sw
)
(6.2.2)
with kernel operator
A = D
sw
(
M
0
)
, Hermitian operator
H = Γ
55
D
sw
(
M
0
)
and
negative mass parameter
M
0
within the range
M
0
(
0, 2r
)
in the free-field case. We
find that the action
S
so
= χD
so
χ
is invariant under the infinitesimal transformation
δ χ = χ Γ
55
, δ χ = Γ
55
(
1
aD
so
)
χ, (6.2.3)
cf. Eq.
(6.1.2)
for the usual overlap operator. Furthermore, the staggered overlap
67
6 Staggered overlap fermions
operator satisfies the Ginsparg-Wilson relation
{
Γ
55
, D
so
}
=
a
r
D
so
Γ
55
D
so
. (6.2.4)
As in four space-time dimensions
D
sw
describes two tastes, the staggered overlap
operator
D
so
represents two degenerate chiral fermion species on the lattice, which
can be used to describe e.g. the up and down quark.
6.3 The index theorem and the axial anomaly
With respect to the topological aspects we are discussing throughout this thesis,
let us introduce some important concepts from topology. If
n
denote the number
of left-handed and right-handed zero modes, then we define the integer-valued
index of the Dirac operator D as
index D = n
n
+
. (6.3.1)
In the
d
-dimensional (
d
even) continuum setting of quantum chromodynamics,
the famous Atiyah-Singer index theorem [121124] states that
n
n
+
=
(
1
)
d/2
Q
cont
, (6.3.2)
where Q
cont
is the so-called topological charge given by
Q
cont
=
d
4
x q
cont
(
x
)
, q
cont
(
x
)
=
1
32π
2
ε
µν ρσ
tr
F
µν
(
x
)
F
ρσ
(
x
)
(6.3.3)
and
F
µν
(
x
)
is the corresponding field strength tensor. The index theorem in Eq.
(6.3.2)
is remarkable given the fact that the right hand side appears to be a continuous
function of the gauge field, while the left hand side is an integer number. On the
lattice one can formulate a similar index theorem [110], which reads
n
n
+
=
(
1
)
d/2
Q (6.3.4)
with
Q = a
4
x Λ
q
(
x
)
, q
(
x
)
=
1
2a
3
tr
[
γ
5
D
(
x, x
)]
. (6.3.5)
We note that in practice one never finds both
n
and
n
+
to be non-vanishing at
the same time on thermalized configurations, i.e. one finds either
n
=
(
1
)
d/2
Q
or
n
+
=
(
1
)
1+d/2
Q
. Moreover, in two dimensions one can show the Vanishing
68
6.4 Continuum limit of the index and axial anomaly
Theorem [
125
127
]. That is, one can prove that if
Q ,
0, then either
n
or
n
+
necessarily vanishes.
While a chiral flavor-singlet rotation is a symmetry of the action for a Dirac
operator satisfying the Ginsparg-Wilson relation, the path integral itself is not
invariant due to the nontrivial transformation of the integration measure [
110
,
111
],
cf. Refs. [
128
132
] for the continuum case. This so-called axial anomaly results in an
anomalous continuity equation for the singlet axial vector current. The presence
of the axial anomaly implies that there is no flavor-singlet Goldstone particle and
explains why the
η
-meson is significantly heavier than the pions assuming an
approximate chiral symmetry for two flavors, see e.g. Ref. [
16
]. In the three flavor
case the Witten-Veneziano formula [
133
140
] connects the mass of the
η
0
-meson to
the topological susceptibility χ
top
=
Q
2
/V , where V is the space-time volume.
6.4 Continuum limit of the index and axial anomaly
Massless overlap fermions have exact zero modes with definite chirality and one
can define an index for the overlap Dirac operator as discussed in Subsec. 6.3.
Overlap fermions implement an exact chiral symmetry on the lattice, which is
anomalously broken in the quantum case and gives rise to the axial anomaly. This
axial anomaly and the index density of the overlap Dirac operator are proportional,
while the continuum axial anomaly is proportional to the topological charge density
and by the Atiyah-Singer index theorem [
121
124
] we find equality between the
topological charge and the index.
In the setting of lattice quantum chromodynamics a natural question is then if
the continuum axial anomaly and index are recovered in the continuum limit in
sufficiently smooth gauge background fields. In the past several investigations [
109
,
110
,
141
145
] dealt with this question. Here, as a first step to secure the theoretical
foundations for Adams construction, we want to check that the continuum axial
anomaly and index is correctly reproduced for the staggered overlap Dirac operator.
Our strategy for evaluating the continuum limit is similar to the Wilson case [
144
,
145
], although one encounters some technical complications due to the nontrivial
spin-flavor structure of staggered fermions. In Ref. [
3
], we found that the index
correctly reproduces the index in the continuum (cf. the numerical studies in
Refs. [
70
,
71
,
146
]), while for the anomaly first an averaging over the sites of a lattice
hypercube is needed. This is not unexpected as in the staggered formalism spin
69
6 Staggered overlap fermions
and flavor degrees of freedom are distributed over lattice hypercubes, while the
distance between neighboring lattice sites is of order
O
(
a
)
, shrinking to zero in
the continuum limit.
In this section, we briefly review our discussion
1
given in Ref. [
3
], where we
previously presented the following results. For a more elaborate discussion, we
refer the reader to the thesis in Ref. [11].
6.4.1 The index and axial anomaly
We now give an overview of the derivation of the continuum limit of the index
and anomaly of the staggered overlap operator in the setting of lattice quantum
chromodynamics. We begin by recalling the definition of the staggered overlap
operator as defined in Eq. (6.2.2) for the case of r = 1, which reads
D
so
=
1
a
1
+ Γ
55
H
H
2
, H = Γ
55
D
sw
(
m
)
. (6.4.1)
Using the index formula in Ref. [110], we can write
index D
so
=
1
2
Tr
H
H
2
. (6.4.2)
In the following, our setting is a four-dimensional box of length
L
with lattice
spacing
a
and
N
sites in each direction. The lattice transcripts of a smooth con-
tinuum gauge field
A
µ
(
x
)
are taken as the link variables, taking values in the Lie
algebra
su
(
3
)
of
SU
(
3
)
. Furthermore, we impose boundary conditions as described
in Ref. [145].
Let us now express Eq. (6.4.2) as a sum over the index density
index D
so
= a
4
x
q
(
x
)
, q
(
x
)
=
1
2
tr
c
H
H
2
(
x, x
)
. (6.4.3)
Here
tr
c
denotes a color trace and
Θ
(
x, y
)
is defined for an operator
Θ
by
Θ χ
(
x
)
=
y
Θ
(
x, y
)
χ
(
y
)
. Like in the case of usual overlap fermions, we find that the index
density
q
(
x
)
is related to the axial anomaly via
A
(
x
)
=
2
iq
(
x
)
, where
A
(
x
)
is the
divergence of the axial current.
1
Discussion based on D. H. Adams, R. Har, Y. Jia, and C. Zielinski, “Continuum limit of the axial
anomaly and index for the staggered overlap Dirac operator: An overview,” PoS
LATTICE2013
(2014) 462,
arXiv:1312.7230 [hep-lat]
. Copyright owned by the authors under the terms of
the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License
(CC BY-NC-ND 4.0).
70
6.4 Continuum limit of the index and axial anomaly
In the case of the overlap construction with a Wilson kernel, the central problem
in verifying the correct continuum limits of the index and axial anomaly is to show
that
lim
a0
q
(
x
)
=
1
32π
2
ε
µνσ ρ
tr
c
F
µν
(
x
)
F
σ ρ
(
x
)
, (6.4.4)
where
tr
c
denotes a trace in color space, as it was done in Refs. [
144
,
145
]. The
a
0
limit is taken by consecutive symmetric refinements of the underlying lattice. As
mentioned earlier, in the case of the staggered overlap operator Eq.
(6.4.4)
only
holds if we average over the sites of a lattice hypercube which contains
x
. For the
remainder of this section we are going to outline the proof of Eq.
(6.4.4)
, where we
have to include the averaging procedure and a factor of two on the right hand side
to account for the two physical tastes of the staggered Wilson kernel.
6.4.2 Continuum limit of the index density
Following the derivation for the usual overlap case given in Ref. [
145
], we use the
identity
lim
a0
H
H
2
(
x, x
)
= lim
a0
3π/2a
π/2a
d
4
p e
ipx
H
H
2
e
ipx
, (6.4.5)
where expressions like
px
are understood as scalar products. Let us begin by intro-
ducing some useful notation. We write the lattice momentum
p
µ
[
π/2, 3π/2
]
as
p
µ
=
π
a
B
µ
+ q
µ
, where
B
µ
{
0, 1
}
and
q
µ
[
π/2, π/2
]
. Furthermore, we define
the sum of two vectors
A, B
{
0, 1
}
4
as the componentwise sum
mod
2 and let
e
B
(
x
)
exp
(
iπBx/a
)
for
B
{
0, 1
}
4
. On the vector space
V
spanned by the
e
B
, we
can define two commuting representations of the Dirac algebra [42] given by
ˆ
Γ
µ
AB
=
(
1
)
A
µ
δ
A,B+η
µ
, (6.4.6)
ˆ
Ξ
µ
AB
=
(
1
)
A
µ
δ
A,B+ζ
µ
, (6.4.7)
where
η
µ
σ
=
1 for
σ < µ
,
ζ
µ
σ
=
1 for
σ > µ
and all remaining components
being zero. Here we use the notation with hat as the versions without the hat
denote the extensions from the space
V
to the one-component spinor field space.
In the following, we use the notation
Ξ
5
Γ
55
Γ
5
, C
5
sym
α βγδ
C
α
C
β
C
γ
C
δ
/4!, (6.4.8)
where
C
5
is the symmetric product of the
C
µ
=
T
µ
+ T
µ
/
2 operators. In our
71
6 Staggered overlap fermions
notation, the staggered Wilson term reads
W
st
=
r
a
(
1
Ξ
5
)
, cf. Sec. 3.4. Writing a
plane wave as
exp
(
ipx
)
= exp
(
i
[
πB/a + q
]
x
)
= e
B
(
x
)
exp
(
iqx
)
, (6.4.9)
we can understand the action of the staggered Wilson kernel D
sw
in AB-language
to be
D
sw
e
ipx
= D
sw
e
B
(
x
)
e
iqx
= e
A
(
x
)
ˆ
Γ
µ
AB
µ
+
r
a
δ
AB
ˆ
Ξ
5
AB
C
e
iqx
= e
A
(
x
)
ˆ
Γ
µ
µ
+
r
a
1
ˆ
Ξ
5
C
AB
e
iqx
e
A
(
x
)
˜
D
sw
AB
e
iqx
, (6.4.10)
where we sum over all sixteen possible values of the repeated
A
-index. Similarly,
we find for the Γ
55
operator
Γ
55
e
ipx
= Γ
55
e
B
(
x
)
e
iqx
= e
A
(
x
)
ˆ
Γ
5
ˆ
Ξ
5
AB
e
iqx
. (6.4.11)
We then conclude that
H
H
2
e
ipx
=
H
H
2
e
B
(
x
)
e
iqx
= e
A
(
x
)
˜
H
˜
H
2
AB
e
iqx
, (6.4.12)
where
˜
H
follows from
H
by replacing
D
sw
˜
D
sw
and
Γ
55
ˆ
Γ
5
ˆ
Ξ
5
. This allows us
to rewrite Eq. (6.4.5) as
lim
a0
H
H
2
(
x, x
)
= lim
a0
B
π/2a
π/2a
d
4
q e
iqx
e
B
(
x
)
H
H
2
e
B
(
x
)
e
iqx
= lim
a0
A,B
e
iπ
(
AB
)
x/a
π/2a
π/2a
d
4
q e
iqx
˜
H
˜
H
2
AB
e
iqx
.(6.4.13)
In the last sum in Eq.
(6.4.13)
, we now separately analyze the contributions from
the case
A = B
and
A , B
. While the former reproduces the continuum topological
charge density
q
(
x
)
, the latter contributions vanish after averaging over a lattice
hypercube containing x .
72
6.4 Continuum limit of the index and axial anomaly
The A = B case
Let us begin with the A = B case in Eq. (6.4.13). In this case we find
lim
a0
π/2a
π/2a
d
4
q e
iqx
Tr
˜
H
˜
H
2
e
iqx
, (6.4.14)
where the trace is taken in the vector space
V
. Following Ref. [
42
], there exist an
isomorphism such that
ˆ
Γ
µ
γ
µ
1
and
ˆ
Ξ
ν
1
ξ
ν
. In a basis where
ξ
5
is diagonal,
we find
˜
D
sw
=
γ
µ
1
µ
+
(
1
1
)
r
a
(
1 C
5
)
(6.4.15)
with signs
for the flavors on which
ξ
5
= ±
1
. On each of these two flavor subspaces
˜
D
sw
is a hypercubic lattice Dirac operator as discussed in Ref. [
147
] and
Γ
55
γ
5
ξ
5
= ±
(
γ
5
1
)
. Noting that the momentum representation of
C
5
in the free-
field case reads
C
5
(
aq
)
=
µ
cos
aq
µ
= 1 + O
a
2
, (6.4.16)
we see that
˜
D
sw
describes a single physical flavor for each of the two species of
the subspace where
ξ
5
=
1
. Finally, note that in this case Eq.
(6.4.14)
equals
to
tr
H /
H
2
(
x, x
)
with
H =
(
γ
5
1
)
˜
D
sw
m
. Using the general result of
Ref. [
147
] and the fact that the trace over the flavor subspace produces a factor of
two, we find that
q
(
x
)
yields the continuum topological charge density with an
additional factor of two due to the two physical species of the staggered Wilson
kernel.
At the same time, the contribution from the other flavor subspace, where
ξ
5
=
1
, vanishes as there are no physical fermion species, which again follows from
Ref. [147].
The A , B case
We are left with analyzing the
A , B
contributions in Eq.
(6.4.13)
when using the
mentioned averaging procedure. Rewriting the prefactor
e
iπ
(
AB
)
x/a
=
(
1
)
(
AB
)
n
with
n = x /a
, we see that the sum over lattice sites gives zero if not all components
of A B are zero. Hence the problem is reduced to showing that
π/2a
π/2a
d
4
q e
iqx
˜
H
˜
H
2
AB
e
iqx
(6.4.17)
73
6 Staggered overlap fermions
changes by order
O
(
a
)
when
x
is moved within a lattice hypercube. This can be
shown by expanding the integrand in powers of the continuum gauge field as done
in Refs. [
144
,
145
]. While the integral formally diverges as
O
a
4
, one can show
that the lowest non-vanishing contribution in the expansion is of order
O
a
4
. As
the smooth continuum gauge field varies by
O
(
a
)
when moving
x
inside a lattice
hypercube, we conclude that Eq.
(6.4.17)
changes by an order
O
(
a
)
term as well,
completing our proof.
Further discussion
For more details on this derivation we refer the reader to Refs. [
3
,
11
]. While Ref. [
3
]
deals with the above derivation, the thesis in Ref. [
11
] also discusses the index of
lattice Dirac operators in a wider setting.
74
C
S 
An important motivation for studying staggered overlap fermions [
44
,
45
,
112
] is
the prospect of bringing down the enormous computational costs of simulations
with lattice fermions with exact chiral symmetry. A first study on this was done in
Refs. [
70
,
71
], where in the free-field limit a large speedup factor of
O
(
10
)
compared
to usual overlap fermions was found. However, after the introduction of a ther-
malized QCD background field this factor dropped to
O
(
2
)
. At the same time, we
observed a large speedup factor of 4
6 for the staggered Wilson kernel compared
to the Wilson kernel as discussed in Chapter 4.
In this chapter, we try to explain the somewhat surprising observation that
the overlap construction shows a much smaller speedup factor compared to the
underlying kernel operator. We try to explain this discrepancy with the spectral and
chiral properties of the staggered Wilson Dirac operator for different lattices sizes.
To this end, we note that the studies of staggered overlap fermions in Refs. [
70
,
71
]
were carried out on relatively small lattices, where the eigenvalue spectrum has a
small gap and diffuse branches. Our investigations of the computational efficiency
of the staggered Wilson kernel operator in Chapter 4, on the other hand, were done
on significantly larger lattices.
In the following, we present indicators in favor of the idea that the spectral and
chiral properties of the staggered Wilson kernel improve on larger lattices and, thus,
the performance of staggered overlap fermions on small lattices is not representa-
tive of the expected performance in a realistic setting. As only an actual benchmark
on larger lattices will eventually clarify this point, the following discussion should
be taken as a proposed explanation. Some of the results discussed in this chapter
were previously presented in Refs. [4, 7].
7.1 Introduction
While usual overlap fermions [
102
108
] are very attractive due to the presence of
chiral zero modes and having a well-defined index, their high computational costs
75
7 Spectral properties
make their practical use for high-precision studies limited. By using staggered
overlap fermions, which use the staggered Wilson Dirac operator as a kernel oper-
ator, one can potentially reduce these costs significantly. Initial numerical tests
confirmed this expectation only partially [
70
,
71
], see also Ref. [
82
]. While in the
free-field case an impressive speedup factor of
O
(
10
)
compared to usual overlap
fermions was observed, in a quenched QCD background field this factor reduced
to O
(
2
)
.
Naturally the computational performance of the overlap operator depends on
the spectral and chiral properties of the underlying kernel operator. In particular,
a more chiral kernel is expected to result in lowered computational costs [
148
]. The
reduction of the speedup factor for staggered overlap fermions, when moving from
the free-field case to thermalized configurations, was explained in Refs. [
70
,
71
]
by the spectral properties of the staggered Wilson kernel. While in the free-field
limit the eigenvalue spectrum has a shape close to the Ginsparg-Wilson circle,
in the case of a
β =
6 background field the spectrum contracts, the gap of the
spectrum becomes narrow and the branches become diffuse. At
β =
5
.
8, these
effects become even more severe. Due to these changes in the eigenvalue spectrum
significantly more computational time is needed, partly offsetting the advantage
of a smaller fermion matrix.
At the same time we could show in Refs. [
4
,
7
], that at
β =
6 the staggered
Wilson kernel has a significantly reduced condition number and is 4
6 times more
efficient for inverting the Dirac operator on a source compared to the Wilson kernel.
The difference in the relative performance of staggered overlap vs. usual overlap
fermions compared to the one of their kernel operators as well as the connection
to spectral properties then has to be clarified.
Looking at Refs. [
70
,
71
], we note the use of small lattices of up to a size of 12
4
for
benchmarking the staggered overlap operator. Generally, the use of large lattices
for overlap fermions is difficult due to the high computational costs and might be
difficult to justify for exploratory studies. On the other hand, the studies of the
staggered Wilson Dirac operator in Refs. [
4
,
7
] were carried out on lattice sizes of
up to 20
3
× 40.
This gives rise to the question if the observed large differences of the computa-
tional efficiency of overlap vs. staggered overlap fermions on small lattices com-
pared to the respective kernel operators on larger lattices are connected to changes
in the eigenvalue spectrum and chiral properties when varying the lattice size.
To answer this question, we investigate the eigenvalue spectrum of the staggered
76
7.2 Free-field case
Wilson Dirac operator. We investigate how the spectrum changes when moving
from small to larger lattices and discuss how this can impact the performance of
staggered overlap fermions.
This chapter is organized as follows. In Sec. 7.2, we begin by discussing the
free-field eigenvalue spectrum of the relevant fermion discretizations. In Sec. 7.3,
we investigate eigenvalue spectra in the case of quenched QCD background fields
and discuss how changes in the spectra affect computational performance. Finally,
in Sec. 7.4 we make some concluding remarks.
7.2 Free-field case
We begin by examining the free-field eigenvalue spectra of usual Wilson and stag-
gered Wilson fermions. Throughout this chapter, we set the lattice spacing to
a =
1
and, in order to avoid ambiguities, write summations explicitly.
7.2.1 Wilson fermions
Wilson fermions differ from the naïve discretization by the introduction of the Wil-
son term as discussed in Subsec. 2.2.6. The Wilson term is an additional covariant
Laplacian term and we can write the resulting Dirac operator in the free-field case
as
D
w
(
x, y
)
=
(
m + 4r
)
δ
x, y
1
1
2
µ
r
1
γ
µ
δ
x+ ˆµ, y
+
r
1
+ γ
µ
δ
x ˆµ, y
. (7.2.1)
Here
m
refers to the bare fermion mass,
r
(
0, 1
]
to the Wilson parameter,
1
to a
4 × 4 unit matrix in spinor space, the γ
µ
matrices to a representation of the Dirac
algebra
γ
µ
, γ
ν
= δ
µ, ν
1
and
δ
a,b
to the Kronecker delta. Going to momentum
space we find
D
w
(
p
)
=
(
m + 4r
)
1
+ i
µ
γ
µ
sin p
µ
r
ν
cos p
ν
1
. (7.2.2)
The eigenvalues of this matrix can be easily computed as
λ
w
(
p
)
= m + 2r
µ
sin
2
p
µ
/2
± i
ν
sin
2
p
ν
. (7.2.3)
77
7 Spectral properties
On a finite lattice the values of p
µ
are discretized. The allowed values are
p
µ
= 2π
n
µ
+ ε
µ
/N
µ
(7.2.4)
with
n
µ
=
0
,
1
, . . . , N
µ
1. In the case of periodic boundary conditions we have
ε
µ
=
0, for antiperiodic boundary conditions
ε
µ
=
1
/
2. In both cases
N
µ
denotes
the number of slices in µ-direction.
7.2.2 Staggered Wilson fermions
Two-flavor staggered Wilson fermions were discussed in Sec. 3.4. In the free-field
limit, the Dirac operator has a spin flavor interpretation [42, 112] of
D
sw
(
p
)
= m
(
1
1
)
+ i
µ
sin p
µ
γ
µ
1
+ r
1
1
γ
5
ν
cos p
ν
. (7.2.5)
Explicitly multiplying out the Kronecker products we find a 16
×
16 matrix, whose
eigenvalues can be calculated as
λ
sw
(
p
)
= m ± i
µ
sin
2
p
µ
+ r
1 ±
ν
cos p
ν
. (7.2.6)
Here the two
±
are to be chosen independently and the allowed range for
p
µ
is
as for Wilson fermions, but with
n
µ
=
0
,
1
, . . . , N
µ
/
2
1. This expression was also
derived in Ref. [71].
For Wilson and staggered Wilson fermions, the free-field eigenvalue spectrum
can be found in Fig. 7.2.1. As one can see, the staggered Wilson spectrum is close to
the Ginsparg-Wilson circle and we have a clear separation between the physical and
the doubler branch. Together with the fact that the staggered Wilson fermion matrix
is smaller than the Wilson fermion matrix, one expects a significant computational
advantage for staggered overlap fermions. This was confirmed by the free-field
speedup factor of O
(
10
)
in the studies of Refs. [70, 71].
78
7.3 Quenched quantum chromodynamics
Figure 7.2.1:
Eigenvalue spectrum of the staggered Wilson (red) and the Wilson
Dirac operator (blue) in the free-field case together with the Ginsparg-
Wilson circle (green).
7.3 Quenched quantum chromodynamics
Let us now consider the case of (unsmeared) quenched QCD background fields.
The calculation of eigenvalue spectra on large lattices is numerically and computa-
tionally challenging. The Dirac operator is represented by a large, nonsymmetric,
complex-valued matrix of size
N
c
N
d
N
3
s
N
t
2
, where for quantum chromodynamics
the number of colors is
N
c
=
3 and the number of spinor components is given by
N
d
=
4 for Wilson and by
N
d
=
1 for staggered Wilson fermions. In particular, for
larger lattices such as a
N
s
× N
t
=
16
3
×
32 lattice, we have more than 1
.
5
×
10
6
eigenvalues for the Wilson Dirac operator and more than 5
×
10
5
for the staggered
Wilson Dirac operator. As a result, we can compute complete eigenvalue spectra
only for small lattice sizes. For larger lattices, we have to restrict ourselves to sub-
sets of the spectrum. Of particular interest is here the physically relevant low-lying
part around the physical branch.
7.3.1 Implementation
To determine the eigenvalues of interest, our numerical implementation works
as follows. For small lattices we compute complete eigenvalue spectra using the
 [
149
] library. The library function
zgeev
computes all eigenvalues of a
79
7 Spectral properties
(a) 32
4
lattice, free field (b) 8
4
lattice at β = 6 (c) 10
4
lattice at β = 6
Figure 7.3.1: The staggered Wilson eigenvalue spectrum at β = 6.
complex nonsymmetric matrix in double precision using the
QR
-algorithm [
150
152
]. As the function requires the input matrix to be constructed explicitly (i.e.
densely) in memory, its use for larger lattices is limited by quickly growing memory
requirements. As a consequence we use this method to compute complete spectra
for lattices of up to a size of 6
4
for the Wilson Dirac operator and up to a size of 10
4
in the case of the staggered Wilson Dirac operator.
For larger lattices we use  [
153
], a library to solve large-scale eigenvalue
problems involving sparse and structured matrices. It uses the implicitly restarted
Arnoldi method [
154
] to compute the low-lying part of the eigenvalue spectrum.
Here low-lying refers to eigenvalues with smallest absolute values. The implicitly
restarted Arnoldi method is an iterative method and is not suitable for the compu-
tation of complete spectra, hence our choice of  for smaller lattices. It has
nontrivial convergence behavior, in particular choosing the number of Lanczos
basis vectors has significant impact on the performance. The
znaupd
function
deals with the case of a complex nonsymmetric matrix in double precision. By
choosing the bare mass parameter
m
appropriately, one can shift the part of the
spectrum of interest to the origin. With our setup we can compute typically a few
hundred eigenvalues for lattice sizes of up to 16
3
×
32 in quenched QCD background
fields.
80
7.3 Quenched quantum chromodynamics
(a) 32
4
lattice, free field (b) 8
4
lattice at β = 5.8 (c) 10
4
lattice at β = 5.8
Figure 7.3.2: The staggered Wilson eigenvalue spectrum at β = 5.8.
In the following, all reported numerical results are for periodic boundary condi-
tions, use the canonical choice
r =
1 and are shifted back to
m =
0. All numerical
calculations are done in double precision as the use of single precision resulted in
erroneous eigenvalues due to accumulated rounding errors.
7.3.2 Eigenvalue spectra
We now begin our discussion of the eigenvalue spectra by comparing the free-field
case with a
β =
6 quenched QCD background field on small lattices. In Fig. 7.3.1, we
can find two typical examples for the spectrum for a 8
4
and a 10
4
lattice. Compared
to the free-field spectrum in the same figure, we can see how the whole spectrum
is contracted noticeably and the gap is narrow. In Ref. [
70
], this degradation of
the spectrum was attributed to the four-hop terms in the staggered Wilson action,
which raise gauge fluctuations to the fourth power. The physical branch and the
doubler branch get closer as the spectrum collapses into a vertical stripe and one
would expect a reduced computational efficiency of staggered Wilson fermions.
For
β =
5
.
8 the spectrum degrades further. In Fig. 7.3.2, we can find two exem-
plary eigenvalue spectra. We see that the branches become even more diffuse and
the gap is very narrow. As a result, we expect a further decreased performance
81
7 Spectral properties
−0.5
−0.4
−0.3
−0.2
−0.1
0.0
0.1
0.2
0.3
0.4
0.5
0.6 0.7 0.8 0.9
Im λ
Re λ
8
4
lattice, β = 6
(a) 8
4
lattice at β = 6
−0.5
−0.4
−0.3
−0.2
−0.1
0.0
0.1
0.2
0.3
0.4
0.5
0.6 0.7 0.8 0.9
Im λ
Re λ
10
4
lattice, β = 6
(b) 10
4
lattice at β = 6
−0.5
−0.4
−0.3
−0.2
−0.1
0.0
0.1
0.2
0.3
0.4
0.5
0.6 0.7 0.8 0.9
Im λ
Re λ
12
3
×16 lattice, β = 6
(c) 12
3
× 16 lattice at β = 6
Figure 7.3.3:
Close-up of the physical branch of the staggered Wilson Dirac operator
for increasing lattice sizes.
compared to the β = 6 case.
The situation changes when we move to larger lattices. In Fig. 7.3.3, we can find
a close-up of the physical branch over an increasing range of lattice sizes from 8
4
to 12
3
×
16 at
β =
6. We can observe how for larger lattices the physical branch
becomes increasingly sharp. Finally, in Fig. 7.3.4 we can find a direct comparison
of the physical branch of both the Wilson and staggered Wilson Dirac operator
up to the largest lattices we can access, i.e. 16
3
×
32. Here we can see very clearly
how the spectrum improves from small to large lattices. While on small lattices
the branches are diffuse and not well-separated, on larger lattices the situation
improves remarkably.
82
7.4 Conclusions
−0.2
−0.1
0.0
0.1
0.2
0.7 0.8 0.9
Im λ
Re λ
12
3
×16 lattice, β = 6
(a) 12
3
× 16 lattice at β = 6
−0.2
−0.1
0.0
0.1
0.2
0.7 0.8 0.9
Im λ
Re λ
12
3
×32 lattice, β = 6
(b) 12
3
× 32 lattice at β = 6
−0.2
−0.1
0.0
0.1
0.2
0.7 0.8 0.9
Im λ
Re λ
16
3
×32 lattice, β = 6
(c) 16
3
× 32 lattice at β = 6
Figure 7.3.4:
Comparison of the staggered Wilson (red) and Wilson Dirac spectrum
(blue).
7.4 Conclusions
Our observations give a possible explanation for the modest speedup factors for
staggered overlap fermions observed in Refs. [
70
,
71
] compared to the large speedup
factors for staggered Wilson fermions found in Refs. [
4
,
7
]. While the former study
was carried out on small lattices up to a size of 12
4
, the latter one was done on
lattices sizes of up to a size of 20
3
× 40.
As we confirmed, on small lattices the eigenvalue spectrum has a small gap and
is lacking a clear separation of the branches. Here one can often find eigenvalues
close to the center of the “belly” of the spectrum, making the overlap construc-
tion computationally more expensive. The situation quickly improves on both
the infrared and ultraviolet part of the spectrum when increasing the lattice size,
see Fig. 7.3.3. While this also holds for Wilson fermions, it appears to be more
pronounced for the staggered Wilson Dirac operator, which is likely linked to the
presence of four-hop terms in the staggered Wilson term.
That the eigenvalue spectrum of the staggered Wilson Dirac operator is well-
behaved on larger lattices can also be seen from the condition number. The diffuse
spectrum could in principle give rise to almost vanishing eigenvalues when the pion
mass is small. However, as we observed the condition number in our numerical
studies in Sec. 4.3 to be typically a factor of
O
(
4
)
smaller compared to Wilson
83
7 Spectral properties
fermions, we conclude that these pathological cases are rare.
Finally the smaller additive mass renormalization of staggered Wilson fermions
is suggesting improved chiral properties compared to Wilson fermions, which are
expected to have a positive impact on the computational efficiency.
As a consequence, we believe that the speedup factors reported in Refs. [
70
,
71
]
are not representative for the actual performance gains achievable with staggered
overlap fermions on larger lattices. We would expect a higher computational effi-
ciency in cases where the staggered Wilson eigenvalue spectrum shows a wide gap
and well-separated, sharp branches. As our study suggests, this can be (at least
partially) achieved by moving to larger lattices, although only a comprehensive
numerical study can clarify this point in the end. Alternatively, as suggested in
Ref. [
82
], smearing also significantly improves the eigenvalue spectrum and, thus,
is expected to result in an improved computational performance.
84
C
S   
Besides staggered overlap fermions, the staggered Wilson kernel allows the con-
struction of a related chiral lattice formulation, namely staggered domain wall
fermions as proposed by Adams in Refs. [
44
,
45
]. In the following, we investigate
spectral properties and quantify chiral symmetry violations of various formulations
of staggered and usual domain wall fermions. We do this in the free-field case, on
quenched thermalized background fields in the setting of the Schwinger model and
on smooth topological configurations. Moreover, we present first results for four-
dimensional quantum chromodynamics. We closely follow the discussion from
our original
1
and subsequent
2
reports given in Refs. [
1
,
2
,
12
], where we previously
presented our results.
8.1 Introduction
For our understanding of the low-energy dynamics of quantum chromodynamics,
such as hadron phenomenology, chiral symmetry plays an essential role. While
the proper implementation of chiral symmetry on the lattice turned out to be a
notorious problem, it was eventually overcome with the overlap construction as
discussed in Chapter 6. Overlap fermions obey an exact chiral symmetry, thus
evading the Nielsen-Ninomiya theorem (cf. Subsec. 2.2.4). From a theoretical
perspective overlap fermions have many desirable properties, but their practical
use is limited due to the fact that they typically require a factor of
O
(
10–100
)
more
computational resources compared to Wilson fermions. Moreover, even at moder-
ate lattice spacings tunneling between topological sectors is strongly suppressed
1
Discussion based on and figures reprinted with permission from C. Hoelbling and C. Zielin-
ski, “Spectral properties and chiral symmetry violations of (staggered) domain wall fermions
in the Schwinger model, Phys. Rev.
D94
no. 1, (2016) 014501,
arXiv:1602.08432 [hep-lat]
.
Copyright 2016 by the American Physical Society.
2
Discussion based on and figures reprinted from C. Hoelbling and C. Zielinski, “Staggered do-
main wall fermions, PoS
LATTICE2016
(2016) 254,
arXiv:1609.05114 [hep-lat]
. Copyright
owned by the authors under the terms of the Creative Commons Attribution-NonCommercial-
NoDerivatives 4.0 International License (CC BY-NC-ND 4.0).
85
8 Staggered domain wall fermions
[
155
158
], so that adequate coverage of the different topological sectors becomes a
concern.
An alternative to overlap fermions is given by domain wall fermions [
159
161
],
which implement an approximate chiral symmetry in
d
dimensions by means of a
theory of massive coupled fermions in
d +
1 dimensions (
d =
2
,
4). The extent of
this extra dimension controls the degree of chiral symmetry violations and in the
limit of an infinite extent one recovers the overlap operator with its exact chiral
symmetry. The case of a finite extent can be interpreted as a truncation of overlap
fermions.
Implementations of domain wall fermions can be easily parallelized and bear
the potential of reduced computational costs. While they only have an approxi-
mate chiral symmetry, violations of chiral symmetry are well-controlled. Based on
theoretical grounds, these violations are expected to be exponentially suppressed
[
160
,
162
164
]. In practice, however, this suppression typically still requires large
extents of the extra dimension [
165
169
]. At the same time, these violations allow
for an easier tunneling between different topological sectors.
Typically domain wall fermions are formulated with a Wilson-like kernel operator.
However, recently Adams laid the foundations [
44
,
45
] for the use of staggered
kernels in the domain wall construction. This was enabled by the introduction of a
flavored mass term as described in Sec. 3.4. While in previous studies the properties
of the staggered Wilson kernel [
4
,
71
,
79
82
] and the staggered overlap construction
[
3
,
45
,
70
] were investigated, the staggered domain wall fermion formulation was
largely ignored in the literature. For this reason, we investigate Adams proposal in
this chapter. We give explicit constructions of these lattice fermions in different
variants, study their spectral properties and compare the degree of chiral symmetry
breaking with the one of usual domain wall fermions in the setting of the Schwinger
model [170].
The Schwinger model, i.e. quantum electrodynamics (QED) in two dimensions,
serves us a toy-model for quantum chromodynamics (QCD). Like quantum chro-
modynamics itself, it shows confinement and has topological structure. At the
same time its numerical treatment is simpler and allows the computation of com-
plete eigenvalue spectra of lattice Dirac operators on nontrivial background fields.
Moreover, the study of the Schwinger model is of interest in its own right as it is
relevant for the description of conducting electrons in metals in the low-energy
regime, see e.g. Ref. [171].
This chapter is organized as follows. In Sec. 8.2, we give a short overview of the
86
8.2 Kernel operators
kernel operators we use in our study. In Sec. 8.3, we discuss the construction of
(staggered) domain wall fermions and their variants. In Sec. 8.4, we introduce
effective low-energy Dirac operators and discuss their relations to the overlap
formalism. In Sec. 8.5, we introduce the setting of our numerical calculations. In
Sec. 8.6, we discuss the numerical results in detail. In Sec. 8.7, we present first
results for four-dimensional quantum chromodynamics. In Sec. 8.8, we provide
some exemplary weight factors for the optimal domain wall construction and
finally in Sec. 8.9, we end with some concluding remarks.
8.2 Kernel operators
Throughout this chapter we consider domain wall fermions with two different
kernel operators, i.e. Wilson and staggered Wilson fermions. In the following, we
adopt the notation previously used in Chapters 2, 3 and 6. In the case of
d =
2
dimensions the
γ
µ
C
2×2
matrices (
µ =
1
,
2) correspond to a representation of
the Dirac algebra
γ
µ
, γ
ν
=
2
δ
µν
1
and we denote the chirality matrix as
γ
3
. At
the same time, the
ξ
µ
C
2×2
matrices are a representation of the Dirac algebra in
flavor space.
Wilson fermions.
We introduced and discussed the Wilson Dirac operator in Sub-
sec. 2.2.6. The Dirac operator is defined in Eq.
(2.2.21)
, where we consider the case
of two space-time dimensions for our numerical studies.
Staggered Wilson fermions.
The second kernel operator of interest is Adams stag-
gered Wilson Dirac operator. In Sec. 3.4 and Sec. 3.6, we discussed its construction
and properties in detail. In the numerical part of this chapter we use the formu-
lation in
d =
2 dimensions as defined in Eq.
(3.6.20)
. We note that in principle
also the use of the other flavored mass terms in Chapter 3 can give rise to suitable
kernel operators.
8.3 Domain wall fermions
Domain wall fermion were originally proposed by Kaplan in Ref. [
159
] and subse-
quently refined by Shamir and Furman in Refs. [
160
,
161
]. By means of a
(
d + 1
)
-
dimensional theory, the domain wall formulation realizes approximately massless
87
8 Staggered domain wall fermions
fermions in
d
dimensions. Alternatively, one can interpret the construction as a
tower of N
s
fermions in d dimensions with a specific flavor structure.
We begin by introducing the
(
d + 1
)
-dimensional bulk operators, where we fix
the lattice spacing to
a =
1 in the first
d
dimensions and the (staggered) Wilson
parameter to
r =
1. Although for the numerical work we are interested in the
d =
2 setting, we keep the following discussion general and consider the case of
even
d
dimensions, where
γ
d+1
refers to the corresponding chirality matrix in that
dimension.
8.3.1 Standard construction
We begin with the originally proposed construction. We define
D
±
w
= a
d+1
D
w
(
M
0
)
±
1
(8.3.1)
and explicitly keep the lattice spacing
a
d+1
in the extra dimension. The mass
parameter
M
0
is commonly referred to as the domain wall height and has to be
suitably chosen to give rise to a one flavor theory. For the free-field case, its values
are restricted to the range M
0
(
0, 2r
)
.
We can write down the Dirac operator, which reads
ΨD
dw
Ψ =
N
s
s=1
Ψ
s
D
+
w
Ψ
s
P
Ψ
s+1
P
+
Ψ
s1
(8.3.2)
with chiral projectors
P
±
=
(
1
± γ
d+1
)
/
2 and
(
d + 1
)
-dimensional fermion fields
Ψ
,
Ψ
. Throughout this chapter, the index
s =
1
, . . . , N
s
refers to the additional spatial
coordinate (or equivalently to the flavor). Along the extra spatial coordinate, the
gauge links are taken to be the unit matrix. Furthermore, for finite extents of the
s-coordinate we must impose boundary conditions, which we take to be
P
+
Ψ
0
+ mΨ
N
s
= 0,
P
Ψ
N
s
+1
+ mΨ
1
= 0.
(8.3.3)
Here the parameter
m
is related to the bare fermion mass, see Eq.
(8.4.8)
. In the
special case of
m =
0 the condition reduces to Dirichlet type boundary conditions,
while for
m = ±
1 we find (anti-)periodic boundary conditions. Writing the Dirac
88
8.3 Domain wall fermions
operator explicitly in the extra dimension, we find
D
dw
=
D
+
w
P
mP
+
P
+
D
+
w
P
.
.
.
.
.
.
.
.
.
P
+
D
+
w
P
mP
P
+
D
+
w
. (8.3.4)
The d-dimensional fermion fields can be defined from the boundary as
q = P
+
Ψ
N
s
+ P
Ψ
1
,
q = Ψ
1
P
+
+ Ψ
N
s
P
.
(8.3.5)
Now let us also introduce the reflection operator in the extra dimension
R =
1
.
.
.
1
. (8.3.6)
We can easily verify that D
dw
is R γ
d+1
Hermitian, that is
D
dw
= R γ
d+1
· D
dw
· R γ
d+1
. (8.3.7)
This property implies the reality of the fermion determinant, i.e. det D
dw
R.
Besides this standard formulation, several variants of domain wall fermions have
been proposed after the original proposal.
8.3.2 Boriçi’s construction
Among them we can find Boriçi’s construction [
172
], which follows from the stan-
dard construction by the replacements
P
+
Ψ
s1
D
w
P
+
Ψ
s1
,
P
Ψ
s+1
D
w
P
Ψ
s+1
(8.3.8)
and is an
O
(
a
d+1
)
modification. After this replacement, the Dirac operator takes
the form
ΨD
dw
Ψ =
N
s
s=1
Ψ
s
D
+
w
Ψ
s
+ D
w
P
Ψ
s+1
+ D
w
P
+
Ψ
s1
(8.3.9)
89
8 Staggered domain wall fermions
or explicitly
D
dw
=
D
+
w
D
w
P
mD
w
P
+
D
w
P
+
D
+
w
D
w
P
.
.
.
.
.
.
.
.
.
D
w
P
+
D
+
w
D
w
P
mD
w
P
D
w
P
+
D
+
w
. (8.3.10)
Moreover, the generalization of Eq. (8.3.5) reads
q = P
+
Ψ
N
s
+ P
Ψ
1
,
q = Ψ
1
D
w
P
+
Ψ
N
s
D
w
P
(8.3.11)
and the equivalent of Eq. (8.3.7) takes the form
D
1
D
dw
= R γ
d+1
·
D
1
D
dw
· R γ
d+1
,
D =
1
N
s
D
w
(8.3.12)
(see Ref. [
173
]). Boriçi referred to this formulation as truncated overlap fermions due
to the fact that the corresponding
d
-dimensional effective operator, as discussed
in Sec. 8.4, corresponds to the polar decomposition approximation [
107
,
174
] of
Neubergers overlap operator of order N
s
/2 (where N
s
is even).
8.3.3 Optimal construction
Another modification of domain wall fermions is the optimal construction as pro-
posed by Chiu [
175
,
176
]. Here one modifies
D
dw
so that the effective Dirac operator
is formulated via Zolotarev’s optimal rational function approximation of the
sign
function [
177
179
], see also Refs. [
117
,
180
]. We now summarize the construction
given in Ref. [175].
As a starting point we use Boriçi’s formulation of domain wall fermions. We
generalize the Dirac operator by introducing weight factors
ΨD
dw
Ψ =
N
s
s=1
Ψ
s
D
+
w
(
s
)
Ψ
s
+ D
w
(
s
)
P
Ψ
s+1
+ D
w
(
s
)
P
+
Ψ
s1
(8.3.13)
via
D
±
w
(
s
)
= a
d+1
ω
s
D
w
(
M
0
)
±
1
. (8.3.14)
90
8.3 Domain wall fermions
The weight factors ω
s
take the values
ω
s
=
1
λ
min
1 κ
02
sn
(
v
s
, κ
0
)
(8.3.15)
with
sn
(
v
s
, κ
0
)
being the respective Jacobi elliptic function with the argument
v
s
and modulus κ
0
. The modulus is taken as
κ
0
=
1 λ
2
min
/λ
2
max
(8.3.16)
and λ
2
min
(λ
2
max
) denotes the smallest (largest) eigenvalue of H
2
w
with
H
w
= γ
d+1
D
w
(
M
0
)
. (8.3.17)
The argument v
s
reads
v
s
=
(
1
)
s1
M sn
1
1 + 3λ
(
1 + λ
)
3
,
1 λ
2
+
s
2
2K
0
N
s
, (8.3.18)
where
λ =
N
s
`=1
Θ
2
(
2`K
0
/N
s
, κ
0
)
Θ
2
((
2` 1
)
K
0
/N
s
, κ
0
)
, (8.3.19)
M =
b
N
s
/2
c
`=1
sn
2
((
2` 1
)
K
0
/N
s
, κ
0
)
sn
2
(
2`K
0
/N
s
, κ
0
)
. (8.3.20)
Here
b
·
c
denotes the floor function,
K
0
= K
(
κ
0
)
is the complete elliptic integral of
the first kind with
K
(
k
)
=
π/2
0
dθ
1 k
2
sin
2
θ
. (8.3.21)
We also introduced the elliptic Theta function via
Θ
(
w, k
)
= ϑ
4
πw
2K
, k
(8.3.22)
with
K = K
(
k
)
and elliptic theta functions
ϑ
i
, see e.g. Ref. [
181
]. To allow for the
verification of implementations of these so-called optimal domain wall fermions,
we provide some reference values for the weight factors ω
s
in Sec. 8.8.
91
8 Staggered domain wall fermions
In the optimal construction, Eq. (8.3.5) is replaced by
q = P
+
Ψ
N
s
+ P
Ψ
1
,
q = Ψ
1
D
w
(
1
)
P
+
Ψ
N
s
D
w
(
N
s
)
P
(8.3.23)
and Eq. (8.3.7) again takes the form of Eq. (8.3.12), but now with
D = diag
[
D
w
(
1
)
, . . . , D
w
(
N
s
)]
(8.3.24)
as pointed out in Ref. [
173
]. Optimal domain wall fermions are one of the most
popular domain wall fermion formulations and were used in numerous studies. In
Refs. [182187] some of these works can be found.
Besides the original optimal domain wall fermion formulation, there is also a
modified version [
188
] which is reflection-symmetric along the extra dimension.
Finally, we note that all of the preceding domain wall fermion variants can be
interpreted as special cases of Möbius domain wall fermions [173, 189, 190].
8.3.4 Staggered formulations
As recently clarified by Adams, it is possible to employ a staggered kernel for the for-
mulation of domain wall fermions [
44
,
45
]. Following this idea, the Dirac operator
takes the form
ΥD
sdw
Υ =
N
s
s=1
Υ
s
D
+
sw
Υ
s
P
Υ
s+1
P
+
Υ
s1
, (8.3.25)
where
Υ
refers to the
(
d + 1
)
-dimensional staggered fermion field. Similarly to the
Wilson case, we let
D
±
sw
= a
d+1
D
sw
(
M
0
)
±
1
(8.3.26)
with
D
sw
being the staggered Wilson Dirac operator. The chiral projectors are given
by
P
±
=
(
1
±
)
/
2, where
is defined in Eq.
(3.6.18)
and we note that
2
=
1
. Recall
that
γ
d+1
ξ
d+1
, which reduces to
γ
d+1
1
on the physical species. The
Rγ
d+1
Hermiticity of
D
dw
generalizes to a
R
Hermiticity of
D
sdw
. With our sign
convention
D
sdw
is in full analogy with
D
dw
, while a slightly different convention
was used in the original proposal in Ref. [45].
In Ref. [
45
], a replacement rule was specified, which can be written in our general
92
8.4 Effective Dirac operator
d-dimensional setting as
γ
d+1
, D
w
D
sw
(8.3.27)
and describes how
D
sdw
can be constructed from
D
dw
. Using Eq.
(8.3.27)
, it is
possible for us to generalize also Boriçi’s and Chiu’s work to the case of a staggered
Wilson kernel. This allows us to write down the Dirac operator for what we are
going to denote as truncated staggered domain wall fermions
3
and which reads
ΥD
sdw
Υ =
N
s
s=1
Υ
s
D
+
sw
Υ
s
+ D
sw
P
Υ
s+1
+ D
sw
P
+
Υ
s1
. (8.3.28)
Moreover, the Dirac operator of optimal staggered domain wall fermions, a gener-
alization of Chius construction, takes the form
ΥD
sdw
Υ =
N
s
s=1
Υ
s
D
+
sw
(
s
)
Υ
s
+ D
sw
(
s
)
P
Υ
s+1
+ D
sw
(
s
)
P
+
Υ
s1
. (8.3.29)
Here we let
D
±
sw
(
s
)
= a
d+1
ω
s
D
sw
(
M
0
)
±
1
and the weight factors
ω
s
are given by
Eq. (8.3.15) for the kernel H
sw
= D
sw
(
M
0
)
.
8.4 Effective Dirac operator
We now discuss the effective low-energy
d
-dimensional Dirac operator derived
in Refs. [
172
,
191
193
] (cf. Refs. [
194
,
195
]). This will shine some light on the rela-
tion between the light
d
-dimensional
q
,
q
fields at the boundary and the
(
d + 1
)
-
dimensional theory. We follow Refs. [
172
,
192
,
195
] in our brief discussion of the
construction.
8.4.1 Derivation
After integrating out the
N
s
1 heavy modes in the theory, one can define an
effective low-energy d-dimensional action
S
eff
=
x
q
(
x
)
D
eff
q
(
x
)
, (8.4.1)
3
We chose the name “truncated staggered domain wall fermions in Refs. [
1
,
2
] rather than “trun-
cated staggered overlap fermions” in order to emphasize the fact that we are dealing with a
(
d + 1
)
-dimensional domain wall fermion formulation, which is motivated by a truncation of the
staggered overlap operator.
93
8 Staggered domain wall fermions
where the effective Dirac operator is defined by means of the propagator of light
modes
D
1
eff
(
x, y
)
=
h
q
(
x
)
q
(
y
)i
. (8.4.2)
In the
(
d + 1
)
-dimensional theory we find one light and
N
s
1 heavy Dirac fermions
(for suitable choices of M
0
).
If we take the chiral limit, where at fixed bare coupling
β
we take the extent of
the extra dimension to infinity
N
s
, the contribution from the heavy modes
diverges. This bulk contribution can be canceled by the introduction of pseud-
ofermionic fields, where the pseudofermion action is typically taken to be the
fermion action with m = 1.
Now let us define the Hermitian operators
H
w
= γ
d+1
D
w
(
M
0
)
, H
m
= γ
d+1
D
m
(
M
0
)
, (8.4.3)
where for standard domain wall fermions the kernel operator reads
D
m
(
M
0
)
=
D
w
(
M
0
)
2 ·
1
+ a
d+1
D
w
(
M
0
)
. (8.4.4)
The transfer matrix in the extra dimension takes the form
T =
T
T
+
, T
±
=
1
± a
d+1
H, (8.4.5)
where the Hermitian operator H depends on the construction given by
H =
H
m
for standard construction,
H
w
for Boriçi’s construction.
(8.4.6)
One can derive the explicit form [192] of the effective Dirac operator to find
D
eff
=
1 + m
2
1
+
1 m
2
γ
d+1
T
N
s
+
T
N
s
T
N
s
+
+ T
N
s
. (8.4.7)
Rewriting Eq. (8.4.7) as
D
eff
=
(
1 m
)
D
eff
(
0
)
+
m
1 m
, (8.4.8)
where
D
eff
(
0
)
refers to the effective operator
D
eff
at
m =
0, we find a bare mass of
m/
(
1 m
)
for a given choice of the parameter m, see Ref. [196].
94
8.4 Effective Dirac operator
The effective operator in Eq. (8.4.7) can also be written in the compact form
D
eff
=
P
|
D
1
1
D
m
P
1, 1
(8.4.9)
as shown in Refs. [172, 195], where the matrix P reads
P =
P
P
+
P
P
+
.
.
.
.
.
.
P
P
+
P
+
P
(8.4.10)
and
P
1
= P
|
. To simplify the notation we introduced
D
m
D
dw
(
m
)
and the
(
1, 1
)
-index refers to the corresponding s -block of the matrix product.
For Chius optimal domain wall fermions the derivation follows Boriçi’s case,
where the additional weight factors
ω
s
have to be taken into account [
175
]. By con-
struction, we find that in this case the
sign
-function approximation in the effective
Dirac operator equals Zolotarev’s optimal rational function approximation.
8.4.2 The N
s
limit
For the following discussion we restrict ourselves to the
m =
0 case. First, we
remark that Eq. (8.4.7) can be written as
T
N
s
+
T
N
s
T
N
s
+
+ T
N
s
=
N
s
/2
(
a
d+1
H
)
, (8.4.11)
where
N
s
/2
is Neuberger’s polar decomposition approximation [
107
,
174
] of the
sign
-function. One can easily see that in the
N
s
limit we recover the overlap
Dirac operator,
D
ov
= lim
N
s
→∞
D
eff
=
1
2
1
+
1
2
γ
d+1
sign H
=
1
2
1
+ D
M
0
D
M
0
D
M
0
1
2
, (8.4.12)
95
8 Staggered domain wall fermions
with H given in Eq. (8.4.6), using the shorthand notation D
M
0
= D
(
M
0
)
and
D =
D
m
for standard construction,
D
w
for Boriçi’s/Chius construction.
(8.4.13)
The Ginsparg-Wilson equation
{
γ
d+1
, D
ov
}
= 2D
ov
γ
d+1
D
ov
(8.4.14)
can be shown to hold for
D
ov
, thus implementing an exact chiral symmetry and
implying the normality of the operator. We note that compared to the discussion
in Subsec. 6.1.1, we find an additional factor of two in the Ginsparg-Wilson relation
due to a different normalization of the overlap operator.
If we now compare Eq. (8.4.12) with the standard definition
D
ov
= ρ
1
+ D
ρ
D
ρ
D
ρ
1
2
(8.4.15)
of the overlap operator and we use the relation for the effective negative mass
parameter
ρ =
M
0
a
d+1
2
M
2
0
for standard construction,
M
0
for Boriçi’s/Chius construction,
(8.4.16)
this would result in a restriction on the domain wall height
M
0
from
ρ =
1
/
2. This
limitation can be overcome by rescaling
D
eff
by a factor
% =
2
ρ
, so that in the
free-field case the low-lying eigenvalues of the kernel operator remain invariant
(up to discretization effects) when the effective operator projection is applied. This
can also be seen in Fig. 8.6.7, which we later discuss in Sec. 8.6.1. In all numerical
investigations of this chapter this rescaling is taken into account.
8.4.3 Approximate sign functions
The effective operator, as previously derived, is an approximation to the overlap
operator. Explicitly, the sign-function is approximated by the rational function
r
(
z
)
=
Π
+
(
z
)
Π
(
z
)
Π
+
(
z
)
+ Π
(
z
)
(8.4.17)
96
8.5 Setting
−1.0
−0.5
0.0
0.5
1.0
−4.0 −3.0 −2.0 −1.0 0.0 1.0 2.0 3.0 4.0
Polar
Optimal
(a) N
s
= 2
0.99996
0.99998
1.00000
1.00002
1.00004
1.0 1.5 2.0 2.5 3.0
Polar
Optimal
(b) N
s
= 6
Figure 8.4.1:
Approximations of the
sign
-function by
r
(
z
)
. The optimal construc-
tion is illustrated for the case of λ
min
= 1 and λ
max
= 3.
with
Π
±
(
z
)
=
(
1 ± z
)
N
s
for standard/Boriçi’s construction,
s
(
1 ± ω
s
z
)
for Chiu’s construction.
(8.4.18)
As expected, we can verify that
r
(
z
)
sign
(
z
)
for
N
s
. We note that for the
standard construction and Boriçi’s construction the approximation is identical,
but is applied to different kernel operators, namely
H
m
or
H
w
respectively. To il-
lustrate the approximations in Eq.
(8.4.18)
, we compare the polar decomposition
approximation in Boriçi’s formulation with the optimal approximation in Chius
construction in Fig. 8.4.1. As we can see, the approximation error over the construc-
tion range is significantly lower for the optimal construction. We recall that the
coefficients ω
s
are related to Zolotarev’s coefficients, cf. Refs. [117, 177, 180].
Staggered formulations
The discussion in this section naturally generalizes to the case of a staggered Wilson
kernel, when the replacement rule in Eq.
(8.3.27)
is applied. In particular, the
effective and overlap operators follow from this replacement and we can interpret
staggered domain wall fermions as a truncation of staggered overlap fermions [
45
].
8.5 Setting
We now move on to the numerical part of the study and begin by discussing our
setting. In particular, we elaborate on our method to quantify chiral symmetry
violations to benchmark the various domain wall fermion constructions.
97
8 Staggered domain wall fermions
8.5.1 Setting the domain wall height
To give rise to a one-flavor theory, in the free-field case the domain wall height
M
0
has to be chosen in the range 0
< M
0
<
2
r
. When moving to a nontrivial
background field, this interval generally shifts and contracts. In less rigor terms,
the parameter
M
0
has to be chosen so that the origin is shifted inside the leftmost
“belly” of the eigenvalue spectrum of the kernel operator. Although even in the
free-field case there is no optimal choice for
M
0
as discussed in Ref. [
197
], one
typically uses the canonical choice
M
0
=
1 for the free field, which is at the center
of the eigenvalue belly. More rigorously speaking, valid choices of
M
0
are restricted
by the position of the mobility edge [51, 198200].
In the Schwinger model, for reasonable choices of the inverse coupling
β
the
eigenvalue spectrum is not distorted too far away from the free-field spectrum.
As a consequence, valid choices for
M
0
remain close to the free-field case. While
in four-dimensional quantum chromodynamics one often uses
M
0
=
1
.
8 (see e.g.
Ref. [
201
]), in our present case
M
0
=
1 remains a sensible and simple choice, which
we consequently use for all our numerical work.
8.5.2 Effective mass
Commonly employed measures in the literature to quantify the degree of chiral
symmetry breaking of domain wall fermions are the residual mass
m
res
[
161
,
167
,
201
,
202
] and the effective mass
m
eff
[
203
205
]. While the residual mass makes use of
the explicit mass dependence in the chiral Ward-Takahashi identities, the effective
mass is defined as the lowest eigenvalue (by magnitude) of the Hermitian operator
in a background field with nontrivial topology. Although these measures are not
identical, they usually agree within a factor of order
O
(
1
)
. Consequently both define
valid measures to quantify the chirality of a given lattice fermion formulation.
Because of its conceptual simplicity, we use the effective mass
m
eff
as one of
our measures for the numerical studies. To give a precise definition, we consider
a given Dirac operator
D
on a topologically nontrivial background configuration
with periodic boundary conditions and define the effective mass
m
eff
as the lowest
eigenvalue of the Hermitian kernel operator H . Explicitly, we let
m
eff
= min
λspec H
|
λ
|
= min
Λspec D
D
Λ, (8.5.1)
where we used the fact that
H
2
= D
D
. For a normal operator
D
, this definition
98
8.5 Setting
simplifies to
m
eff
= min
λspec D
|
λ
|
. For a non-normal operator, however, the
eigenvalues of D and H are not directly related.
For gauge configurations with topological charge
Q ,
0, the existence of zero
modes of
H
is guaranteed in the continuum by the Atiyah-Singer index theorem
[
121
124
]. Similarly, the presence of zero modes of the overlap operator as defined
in Eq.
(8.4.12)
can be shown as well [
109
111
]. For the effective Dirac operators these
zero modes are approximate, but become exact in the
N
s
limit. The absolute
value of the eigenvalues of these approximate zero modes can then serve as a
measure of chirality.
In the context of a gauge theory, the Atiyah-Singer index theorem takes the
following form. Let
n
refer to the number of left-handed and right-handed zero
modes, then their difference is related to the topological charge by
n
n
+
=
(
1
)
d/2
Q, (8.5.2)
see Ref. [
44
]. In Eq.
(8.5.6)
, we will give a precise definition of
Q
. We note that one
can show the Vanishing Theorem [
125
127
] in our two-dimensional setting. The
theorem states that on gauge configurations with
Q ,
0, either
n
or
n
+
vanishes,
i.e. n
and n
+
are not simultaneously nonzero.
8.5.3 Normality and Ginsparg-Wilson relation
The continuum Dirac operator, the naïve and overlap lattice Dirac operators are all
normal operators. At the same time, the effective Dirac operators are not normal.
This makes operator normality an interesting aspect in the context of chirality,
especially as it has been shown that chiral properties imply the normality of the
Dirac operator [206], see also Refs. [207, 208].
By definition a normal operator
D
satisfies
D, D
=
0. Deviation from normal-
ity can be quantified by the norm of this commutator and we define
N
=
D, D
, (8.5.3)
where
k
·
k
is the matrix norm induced by the
L
-norm. As previously discussed,
the normality measure
N
vanishes for the overlap operator and, thus, for the
effective Dirac operators in the limit N
s
.
Besides normality, violations of the Ginsparg-Wilson relation as given in Eq.
(8.4.14)
99
8 Staggered domain wall fermions
are a natural measure for chirality of a lattice fermion formulation. We let
GW
=
{
γ
3
, D
}
ρ
1
Dγ
3
D
(8.5.4)
and find that for the effective operators
GW
0 for
N
s
. For a staggered
Wilson kernel, the chirality matrix
γ
3
has to be replaced by the staggered
in the
definition of
GW
. We note that the additional factor of 1
/ρ
in Eq.
(8.5.4)
is due
to the different scales of the effective Dirac operators. The measures
N
and
GW
were previously considered in Refs. [
197
,
209
] and together with
m
eff
serve here as
a measure for chiral symmetry violations.
8.5.4 Topological charge
To determine the topological charge of a gauge configuration, we use the index
formula for the standard overlap operator,
Q =
1
2
Tr
H
w
/
H
2
w
, (8.5.5)
and its staggered equivalent,
Q =
1
2
Tr
H
sw
/
H
2
sw
, (8.5.6)
with
H
sw
= D
sw
(
M
0
)
as discussed in Ref. [
44
]. We numerically verified that these
two definition yield the same topological charge on a set of sample configurations.
This observation is in agreement with analytical results [
3
] and other numerical
studies [71, 210].
8.6 Numerical results
For the numerical part of this work, we calculate complete eigenvalue spectra and
evaluate our measures of chirality for the previously discussed Dirac operators. We
do this on various gauge configurations in the Schwinger model and cover the free-
field case, thermalized configurations and the smooth topological configurations
discussed in Ref. [211].
Throughout this section, we set the lattice spacings in all dimensions to
a =
a
d+1
=
1 and fix the (staggered) Wilson parameter to
r =
1. For the extra dimension
we consider extents within the range 2
N
s
8. In order to allow the calculation
100
8.6 Numerical results
of the effective mass
m
eff
as discussed in Sec. 8.5.2, we impose periodic boundary
conditions in all dimensions.
We determine complete eigenvalue spectra using the  [
149
] library, while
extremal eigenvalues, such as the effective mass, are computed with  [
153
].
For all numerical calculations we use double precision. In the figures, the standard
construction is abbreviated with std”, Boriçi’s construction with Bor” and Chiu’s
optimal construction with opt”. When discussing overlap operators, the construc-
tion with kernel
H
m
is referred to as “DW, Neubergers overlap with kernel
H
w
is
abbreviated as Neub and Adams staggered overlap with kernel
H
sw
is denoted
as Adams.
8.6.1 Free-field case
In our discussion we begin with the simplest case of the free field. Periodicity allows
a Fourier transformation of the kernel operators and simplifies the computation
of the eigenvalue spectra, cf. Chapter 7. In the momentum representation, the
Wilson kernel takes the form of the following 2 × 2 linear map
D
w
=
(
m
f
+ 2r
)
1
+ i
µ
γ
µ
sin p
µ
r
ν
cos p
ν
1
, (8.6.1)
where
p
µ
=
2
πn
µ
/N
µ
with
n
µ
=
0
,
1
, . . . , N
µ
1 and
N
µ
denotes the number of
slices in
µ
-direction. The staggered Wilson kernel is represented by the 4
×
4 linear
map
D
sw
= m
f
(
1
1
)
+ i
µ
sin p
µ
γ
µ
1
+ r
1
1
+ ξ
3
ν
cos p
ν
(8.6.2)
with n
µ
= 0, 1, . . . , N
µ
/2 1, cf. Ref. [71].
For the three-dimensional bulk operator, the extra dimension is lacking periodic-
ity and we leave the corresponding coordinate in the position-space representation.
By employing a momentum-space representation for the kernel operators, we re-
duce the dimensionality of the problem and avoid numerical instabilities otherwise
encountered in the free-field case. All free-field results discussed in the following
are for the case of a N
s
× N
t
= 20 × 20 lattice.
Kernel operators.
In Fig. 8.6.1, we show the free-field eigenvalue spectra of the Wil-
son and staggered Wilson Dirac operator. While in four dimensions Wilson fermi-
101
8 Staggered domain wall fermions
0
0.5
1
1.5
2
2.5
3
3.5
4
Re(λ)
-1.5
-1
-0.5
0
0.5
1
1.5
Im(λ)
Wilson
0
0.5
1
1.5
2
2.5
3
3.5
4
Re(λ)
-1.5
-1
-0.5
0
0.5
1
1.5
Im(λ)
staggered Wilson
Figure 8.6.1: Spectrum of the kernel operators in the free field case.
ons have four doubler branches, in our setting of the two-dimensional Schwinger
model only two doubler branches are present. Adams staggered Wilson term,
however, always splits the tastes into two groups with positive and negative flavor-
chirality. Compared to Wilson fermions, we observe how the spectrum of the
staggered Wilson kernel resembles the Ginsparg-Wilson circle much more closely.
On sufficiently smooth configurations one can then hope for lower computational
costs for the construction of chiral fermion formulations.
We note that the Wilson and staggered Wilson eigenvalue spectrum differs mostly
in their ultraviolet parts and very little in the physically relevant low-lying part.
Nevertheless, these discretization effects alter the properties and shape of the bulk,
effective and overlap operators and in general have an impact on computational
efficiency and chiral properties.
Bulk operators.
Moving on to the
(
2 + 1
)
-dimensional bulk operators, in Figs. 8.6.2
and 8.6.3 we show the eigenvalue spectra in the standard, Boriçi’s and Chius con-
struction. To illustrate the effect of changing boundary conditions in the extra
dimension, we show in Fig. 8.6.4 the case of periodic (
m =
1) and antiperiodic
(
m =
1) boundary conditions, which can be compared to the Dirichlet (
m =
0
)
case.
In the spectrum of the bulk operator with the standard construction, we find
(two) three doubler branches for the case of a (staggered) Wilson kernel. For Boriçi’s
102
8.6 Numerical results
-2 0 2 4
6
Re(λ)
-2
0
2
Im(λ)
(a) Standard construction
-2 0 2 4
6
Re(λ)
-2
0
2
Im(λ)
(b) Boriçi’s construction
-2 0 2 4
6
Re(λ)
-2
0
2
Im(λ)
(c) Optimal construction
Figure 8.6.2:
Spectrum of
D
dw
with Wilson kernel for
m =
0 at
N
s
=
8 in the free
field case.
-2 0 2 4
6
Re(λ)
-2
0
2
Im(λ)
(a) Standard construction
-2 0 2 4
6
Re(λ)
-2
0
2
Im(λ)
(b) Boriçi’s construction
-2 0 2 4
6
Re(λ)
-2
0
2
Im(λ)
(c) Optimal construction
Figure 8.6.3:
Spectrum of
D
sdw
with staggered Wilson kernel for
m =
0 at
N
s
=
8 in
the free field case.
construction the number of branches is reduced by one. In both cases we find 2
N
s
exact zero modes in the Dirichlet case [
160
,
203
], which disappear for non-vanishing
values of
m
. For Chiu’s optimal construction we only find two approximate zero
modes and observe that the real eigenvalues are spread along the real axis. In the
case of Boriçi’s and the optimal construction with a Wilson kernel, we notice how
the spectrum of the bulk operators lost its resemblance to the three-dimensional
Wilson operator.
Effective operators.
In Figs. 8.6.5 and 8.6.6, we can find the eigenvalue spectra
of the effective Dirac operators
%D
eff
with different kernel operators as defined
in Eq.
(8.4.7)
. The spectra resemble the Ginsparg-Wilson circle and rapidly con-
verge towards it for increasing values of
N
s
. As the free field is smooth, this fast
convergence is not unexpected. Already for small values of
N
s
(such as
N
s
=
8),
it is almost impossible to visually distinguish the spectrum from the one of the
associated overlap operator. We also point out the rapid convergence of Boriçi’s
and Chiu’s construction, in particular with a staggered Wilson kernel.
103
8 Staggered domain wall fermions
-2 0 2 4
6
Re(λ)
-2
0
2
Im(λ)
(a) Periodic case (m = 1)
-2 0 2 4
6
Re(λ)
-2
0
2
Im(λ)
(b) Antiperiodic case (m = 1)
Figure 8.6.4:
Spectrum of
D
dw
with Wilson kernel in the standard construction at
N
s
=
8 in the free field case with different boundary conditions (cf.
Fig. 8.6.2)
By construction, the optimal construction shows significantly improved chiral
properties. For a given spectral range
I =
[
λ
min
, λ
max
]
of the Hermitian kernel op-
erator, the optimal rational function approximation
r
opt
(
z
)
minimizes the maximal
deviation from the sign function
δ
max
= max
z ∈−I I
sign
(
z
)
r
opt
(
z
)
= 1 r
opt
(
±λ
min
)
. (8.6.3)
For a normal operator, such as in the free-field case, the eigenvalues are confined in
a tube around the Ginsparg-Wilson circle with radius
δ
max
. As Zolotarevs optimal
approximation to the
sign
-function minimizes the
k
·
k
-norm, we find a point of
maximal deviation at both
λ
min
and
λ
max
, resulting in the absence of an exact zero
mode. However, this approximate zero mode is decreasing rapidly in magnitude
with increasing values of
N
s
due to the rapid convergence of the underlying
sign
-
function approximation.
Overlap operators.
Taking the
N
s
limit of the effective Dirac operators, we
find an overlap operator. The Hermitian kernel operator
H
of the overlap operator
as given in Eq.
(8.4.6)
depends on the choice of the domain wall construction. In
Fig. 8.6.7, we show the eigenvalue spectrum of
%D
ov
together with the stereographic
projection π of the eigenvalues, given by the map
π
(
λ
)
=
λ
1
1
%
λ
. (8.6.4)
104
8.6 Numerical results
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
std Bor
opt
(a) Wilson kernel
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
std Bor
opt
(b) Staggered Wilson kernel
Figure 8.6.5:
Spectrum of
%D
eff
for the standard (std), Boriçi (Bor) and optimal (opt)
construction at N
s
= 2.
The result is the projection of the eigenvalues onto the imaginary axis, where the
physical low-lying eigenvalues line up as required due to the introduction of the
scale factor in Eq. (8.4.16).
For Adams overlap operator the eigenvalue spectrum is highly symmetric. By
construction, the effective operators in Boriçi’s and the optimal construction con-
verge, depending on the kernel operator, towards Neubergers and Adams overlap
operator for
N
s
. For the standard construction we find an overlap operator
with a modified kernel operator.
8.6.2 U
(
1
)
gauge field case
After the discussion of the free-field case, we now consider nontrivial gauge config-
urations. The Schwinger model is an Abelian gauge theory with symmetry group
U
(
1
)
and in the following we consider quenched thermalized gauge configurations
with the setup of Refs. [
212
,
213
]. Although we restrict ourselves to quenched con-
figurations, it is in principle possible to reweight to an unquenched ensemble
with arbitrary mass [
212
215
]. We begin our discussion with the analysis of the
eigenvalue spectra on a few selected 20
2
configurations at an inverse coupling of
β = 5.
105
8 Staggered domain wall fermions
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
std Bor
opt
(a) Wilson kernel
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
std Bor
opt
(b) Staggered Wilson kernel
Figure 8.6.6:
Spectrum of
%D
eff
for the standard (std), Boriçi (Bor) and optimal (opt)
construction at N
s
= 8.
Kernel operators.
An example of a kernel eigenvalue spectrum on a gauge con-
figuration with topological charge
Q =
1 can be found in Fig. 8.6.8. Compared to
the case of four-dimensional quantum chromodynamics, the branches are much
sharper and better separated, cf. Chapter 7 and Refs. [4, 7, 8, 70, 71, 82].
Bulk operators.
For the eigenvalue spectra of the bulk operators in Figs. 8.6.9 and
8.6.10 we use the same gauge background field as in Fig. 8.6.8. Being a topologically
nontrivial configuration, the effective operator has
|
Q
|
approximate zero modes
which become exact in the limit
N
s
. For the bulk operators, there are
N
s
·
|
Q
|
eigenvalues in the neighborhood of the origin. It is an interesting observation to
see how the eigenvalue spectrum is distorted in the optimal construction in order
to improve chiral properties.
Effective operators.
Again using the same gauge configuration, in Figs. 8.6.11 and
8.6.12 we find the eigenvalue spectra of the effective operators. Here Boriçi’s con-
struction shows superior chiral properties compared to the standard construction
for N
s
4 with respect to our measures m
eff
,
N
and
GW
.
The optimal construction further improves upon that and shows an ever lower
degree of chiral symmetry violations. In the case of Fig. 8.6.12b, the effective mass
m
eff
is already so small that it is in the order of the round-off error and, thus, we are
only able to quote an upper bound. In this nontrivial background field the kernel
106
8.6 Numerical results
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
DW Neub
(a) Wilson kernel
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
DW Adams
(b) Staggered Wilson kernel
Figure 8.6.7:
Spectrum of
%D
ov
with its stereographic projection for the domain
wall (DW) kernel and the standard overlap (Neub/Adams) kernel.
operators lose their normality and the maximum deviation
δ
max
of eigenvalues
from the Ginsparg-Wilson circle in the optimal construction does not have to be
saturated. We also note that a smaller
δ
max
does not necessarily imply a smaller
value of
GW
, but that also the distribution of eigenvalues of
H
is important. There
is, however, a strong correlation between larger
N
s
and smaller values of
GW
, where
the violations of the Ginsparg-Wilson relation rapidly tend to zero for N
s
.
Chius construction is only in a very particular sense optimal, namely the min-
imization of the deviation
δ
max
from the
sign
-function as defined in Eq.
(8.6.3)
.
With respect to other criteria, such as the number of iterations required for solving
a linear system, optimal domain wall fermions are typically not optimal [
189
]. In
theory, it is possible to formulate variants of domain wall fermions which are opti-
mal with respect to other measures such as
GW
. However, such a construction
would likely require more knowledge about the eigenvalue spectrum of H .
For the standard construction
m
eff
,
N
and
GW
have typically the same order of
magnitude for both a Wilson and a staggered Wilson kernel. On the other hand, for
Boriçi’s and the optimal construction a staggered Wilson kernel results in reduced
chiral symmetry violations compared to the Wilson kernel.
An interesting, although rather artificial case is
N
s
=
2. Here we observe that the
performance of all constructions is comparable in the case of a staggered Wilson
kernel, while for the case of a Wilson kernel the standard construction performs
107
8 Staggered domain wall fermions
0
0.5
1
1.5
2
2.5
3
3.5
4
Re(λ)
-1.5
-1
-0.5
0
0.5
1
1.5
Im(λ)
7.85×10
-2
2.23×10
0
9.31×10
0
8.73×10
-2
1.78×10
0
2.62×10
0
m
eff
N
GW
Wil
stag
Figure 8.6.8: Spectrum of kernel operators in a U
(
1
)
background field.
better than Boriçi’s, which itself clearly outperforms the optimal construction.
Overlap operators.
In Fig. 8.6.13, we again show the spectrum of the associated
overlap operators together with a stereographic projection of their eigenvalues. We
omit the chirality measures
m
eff
,
N
and
GW
in the figure labels as they vanish up
to round-off errors.
8.6.3 Smearing
Smearing is a common technique in the simulations of lattice gauge theories. It is
expected to be especially beneficial for the staggered Wilson kernel (in particular in
3
+
1 dimensions) as discussed in Ref. [
82
]. For this reason we consider gauge fields
with three-step  smearing [
216
,
217
]. For the smearing parameter, we choose
the maximal value
α =
0
.
5 in two dimensions within the perturbatively reasonable
range [218].
To test the effects of smearing on the violations of chiral symmetry, in Figs. 8.6.14
and 8.6.15 we directly compare the effective operators on an unsmeared and smeared
version of a gauge configuration at
N
s
=
4. We observe that after three smearing
steps, chiral symmetry violations as measured by
m
eff
,
N
and
GW
are reduced
108
8.6 Numerical results
-2 0 2 4
6
8
Re(λ)
-4
-2
0
2
4
Im(λ)
(a) Standard construction
-2 0 2 4
6
8
Re(λ)
-4
-2
0
2
4
Im(λ)
(b) Boriçi’s construction
-2 0 2 4
6
8
Re(λ)
-4
-2
0
2
4
Im(λ)
(c) Optimal construction
Figure 8.6.9: Spectrum of D
dw
with Wilson kernel for m = 0 at N
s
= 8.
-2 0 2 4
6
8
Re(λ)
-4
-2
0
2
4
Im(λ)
(a) Standard construction
-2 0 2 4
6
8
Re(λ)
-4
-2
0
2
4
Im(λ)
(b) Boriçi’s construction
-2 0 2 4
6
8
Re(λ)
-4
-2
0
2
4
Im(λ)
(c) Optimal construction
Figure 8.6.10: Spectrum of
D
sdw
with staggered Wilson kernel for
m =
0 at
N
s
=
8.
significantly. In some cases we observe a reduction by up to two order of magni-
tude with the exception of the effective mass for the standard construction with a
Wilson kernel.
8.6.4 Topology
On topologically nontrivial configurations with
Q ,
0, the Atiyah-Singer index
theorem ensures that the continuum Dirac operator has exact zero modes. On the
lattice the same can be shown for the overlap operator, while for the effective Dirac
operators we only find approximate zero modes as illustrated in Fig. 8.6.16. These
modes come with a multiplicity of
|
Q
|
as expected from the Vanishing Theorem.
By using the technique outlined in Ref. [
211
] for constructing a smooth configura-
tion for any given topological charge
Q
, we can study topological aspects also in a
more abstract setting. The gauge configurations are maximally smooth in the sense
that they are invariant under the application of the  smearing prescription. We
109
8 Staggered domain wall fermions
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
1.75×10
-2
2.38×10
-1
8.05×10
-1
1.95×10
-2
8.84×10
-1
1.44×10
0
1.12×10
-1
1.43×10
0
1.80×10
0
m
eff
N
GW
std
Bor
opt
(a) Wilson kernel
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
2.01×10
-2
2.40×10
-1
7.74×10
-1
2.19×10
-2
6.68×10
-1
7.04×10
-1
4.57×10
-2
5.59×10
-1
5.55×10
-1
m
eff
N
GW
std
Bor
opt
(b) Staggered Wilson kernel
Figure 8.6.11:
Spectrum of
%D
eff
for the standard (std), Boriçi (Bor) and optimal
(opt) construction at N
s
= 2.
measure
m
eff
,
N
and
GW
for the effective Dirac operators on a wide range of
smooth topological configurations with various
Q
. Our measures of chirality are
identical for the configurations with ±Q and only depend on
|
Q
|
.
Due to the smoothness of the configurations, the measures
m
eff
,
N
and
GW
are all small compared to the typical values one measures on thermalized config-
urations at reasonable values of
β
. We also observe that with increasing values
of
|
Q
|
the chiral symmetry violations become larger. Finally, we show two eigen-
value spectra of bulk operators on a 20
2
lattice in Fig. 8.6.17, revealing a very clear
structure of the spectrum on a topologically nontrivial configuration.
8.6.5 Spectral flow
Besides the study of the eigenvalue spectra, topological aspects—such as the index
theorem—can be studied on the lattice also with the help of spectral flows [
105
,
219
].
Commonly used with a Wilson kernel, one studies the eigenvalues
λ
(
m
f
)
of
the Hermitian operator
H
w
(
m
f
)
as a function of
m
f
. One finds a direct one-to-
one correspondence between the eigenvalue crossings of
H
w
(
m
f
)
and the real
eigenvalues of
D
w
(
m
f
)
. Moreover, the low-lying real eigenvalues of the Wilson
110
8.6 Numerical results
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
2.00×10
-4
4.05×10
-2
1.08×10
-1
4.74×10
-6
5.39×10
-3
1.92×10
-2
3.36×10
-6
2.57×10
-4
3.39×10
-4
m
eff
N
GW
std
Bor
opt
(a) Wilson kernel
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
2.82×10
-4
3.26×10
-2
9.12×10
-2
3.55×10
-6
1.03×10
-3
1.01×10
-3
<3
×10
-8
5.03×10
-6
5.71×10
-6
m
eff
Δ
N
Δ
GW
std
Bor
opt
(b) Staggered Wilson kernel
Figure 8.6.12:
Spectrum of
%D
eff
for the standard (std), Boriçi (Bor) and optimal
(opt) construction at N
s
= 8.
Dirac operator
D
w
(
m
f
)
correspond to the would-be zero modes [
211
]. For each
respective mode, the slope of the eigenvalue crossings at small values of
m
f
equals
minus its chirality. By studying the eigenvalue crossings of the Hermitian operator
in a background field we can, thus, determine the respective topological charge.
Although originally spectral flows of Wilson-like kernels were studied, Adams
generalized the method to staggered kernels [
44
] allowing the study of topological
aspects within the staggered framework [
70
,
71
,
210
,
220
]. While in the literature
these spectral flows have been investigated before, here we want to show the
effectiveness of smearing when studying topological aspects on the lattice.
To this end we show an example for the flow
λ
(
m
f
)
of the lowest 50 eigenval-
ues for the Wilson kernel
H
w
(
m
f
)
and the staggered Wilson kernel
H
sw
(
m
f
)
in
Figs. 8.6.18 and 8.6.19. The 12
2
gauge configuration under consideration was gener-
ated at
β =
1
.
8 and has a topological charge of
Q =
2. We compare the eigenvalue
flow on both the unsmeared and the three-step  smeared (
α =
0
.
5) version
of the configuration and also added the smooth
Q =
2 topological configuration
described in Sec. 8.6.4. While the determination of the topological charge
Q
via the
spectral flow is not easily possible for the rough and unsmeared configuration, the
use of smearing allows an unambiguous identification of the topological sector. In
111
8 Staggered domain wall fermions
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
DW Neub
(a) Wilson kernel
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
DW Adams
(b) Staggered Wilson kernel
Figure 8.6.13:
Spectrum of
%D
ov
with its stereographic projection for the domain
wall (DW) kernel and the standard overlap (Neub/Adams) kernel.
the case of the smooth topological configuration, the two eigenvalue crossings lie
essentially on top of each other. The direct comparison between the figures shows
the effectiveness of smearing for the study of topological aspects on the lattice.
8.6.6 Approaching the continuum
To evaluate the chiral properties of the various effective Dirac operators when
approaching the continuum, we evaluated them on seven ensembles with 1000
configurations each. While keeping the physical volume fixed, we consider the
following ensembles: 8
2
at
β =
0
.
8, 12
2
at
β =
1
.
8, 16
2
at
β =
3
.
2, 20
2
at
β =
5
.
0, 24
2
at
β =
7
.
2, 28
2
at
β =
9
.
8 and 32
2
at
β =
12
.
8. Besides the unsmeared configurations,
we also consider the smeared versions, resulting in
N =
14 000 gauge configurations
in total.
While we are interested in the comparison of the different formulations on finer
lattices, we do not attempt a rigorous continuum limit analysis. The chirality
measures on our finest lattice at
β =
12
.
8 serve us as an indicator for the relative
performance of the different domain wall formulations. In Table 8.6.1, we quote
the median values for the parameter range
N
s
{
2, 4, 6
}
. For
N
s
8 some values
already become so small, that they are comparable with the round-off error and,
thus, cannot be considered reliable. For the standard construction we find that
the chiral symmetry violations at large values of
β
are comparable for the cases of
112
8.6 Numerical results
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
3.04×10
-3
1.72×10
-1
5.11×10
-1
6.87×10
-4
1.42×10
-1
3.30×10
-1
1.12×10
-3
7.89×10
-2
1.12×10
-1
m
eff
N
GW
std
Bor
opt
(a) No smearing
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
5.22×10
-5
2.41×10
-2
2.27×10
-1
4.59×10
-6
2.96×10
-2
2.78×10
-1
1.42×10
-5
1.67×10
-2
3.04×10
-2
m
eff
N
GW
std
Bor
opt
(b) Three smearing iterations
Figure 8.6.14:
Spectrum of
%D
eff
with Wilson kernel for the standard (std), Boriçi
(Bor) and optimal (opt) construction at
N
s
=
4 on a smeared config-
uration with α = 0.5.
a Wilson and a staggered Wilson kernel. At the same time, in the case of Boriçi’s
and the optimal construction we often find differences between the two kernel
operators of one to two orders of magnitude.
In practical applications the spectral bounds λ
min
and λ
max
of optimal domain
wall fermions are set on a per ensemble basis. As especially for small
β
and large en-
sembles one frequently observes values of
λ
min
close to zero, one typically projects
out the lowest lying eigenvalues and treats them exactly in order to arrive at a
kernel operator with a narrower spectrum. As in our study we want to compare
the “baseline performance of the various domain wall fermion formulations, we
decided to not add such an additional algorithmic component mixing into the
results of our benchmarks. However, as we also investigate large ensembles at
small
β
, the observed ratios of
λ
2
min
/λ
2
min
can be very small in these cases, ren-
dering the optimal
L
-approximation of the
sign
-function inadequate and of no
practical use. To mitigate this phenomenon, we decided to construct the
sign
-
function approximation on a per-configuration rather than a per-ensemble basis.
This means that a small eigenvalue on a given configuration does not render the
sign
-function approximation inadequate for the whole ensemble, but only on that
113
8 Staggered domain wall fermions
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
3.71×10
-3
1.51×10
-1
4.67×10
-1
7.69×10
-4
8.73×10
-2
8.59×10
-2
2.90×10
-4
1.17×10
-2
1.39×10
-2
m
eff
N
GW
std
Bor
opt
(a) No smearing
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
5.19×10
-5
2.29×10
-2
1.97×10
-1
1.16×10
-6
1.12×10
-3
5.97×10
-3
1.47×10
-5
4.40×10
-4
6.06×10
-4
m
eff
N
GW
std
Bor
opt
(b) Three smearing iterations
Figure 8.6.15:
Spectrum of
%D
eff
with staggered Wilson kernel for the standard (std),
Boriçi (Bor) and optimal (opt) construction at
N
s
=
4 on a smeared
configuration with α = 0.5.
particular configuration. On the other hand this implies that in our benchmarks
the performance of optimal domain wall fermions tend to be overestimated and
can only be considered indicative of the performance of Chiu’s formulation of
domain wall fermions.
In general, we find that the standard construction is outperformed by Boriçi’s
construction with respect to our chirality measures, while optimal domain wall
fermions show the best overall performance. A notable exception is the effective
mass
m
eff
, where the optimal construction with a Wilson kernel is not so clearly
outperforming the competing variants. This is not surprising, considering that the
optimal rational function approximation of the
sign
-function deviates the most
at
λ
min
(and
λ
max
) and, thus, low-lying eigenvalues are typically not mapped very
accurately. However, as the size of the error decreases rapidly with increasing
values of N
s
, this problem is only present for small values of N
s
.
In this context the distribution of the lowest eigenvalue
λ
min
= min
λspec H
|
λ
|
(8.6.5)
of the respective kernel operators
H = H
w
and
H = H
sw
is of interest. In Fig. 8.6.20,
114
8.6 Numerical results
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
7.37×10
-2
1.82×10
-1
5.25×10
-1
8.26×10
-2
1.46×10
-1
1.46×10
-1
8.27×10
-2
1.74×10
-2
1.99×10
-2
m
eff
N
GW
std
Bor
opt
(a) Configuration with Q = 0
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
3.71×10
-3
1.51×10
-1
4.67×10
-1
7.69×10
-4
8.73×10
-2
8.59×10
-2
2.90×10
-4
1.17×10
-2
1.39×10
-2
m
eff
N
GW
std
Bor
opt
(b) Configuration with Q = 1
0
0.5
1
1.5
2
Re(λ)
-1
-0.5
0
0.5
1
Im(λ)
2.79×10
-3
1.08×10
-1
4.19×10
-1
4.41×10
-4
3.17×10
-2
3.17×10
-2
6.54×10
-5
7.37×10
-3
7.90×10
-3
m
eff
N
GW
std
Bor
opt
(c) Configuration with Q = 3
Figure 8.6.16:
Spectrum of
%D
eff
with staggered Wilson kernel for the standard (std),
Boriçi (Bor) and optimal (opt) construction at
N
s
=
4 on configura-
tions with various Q .
-2 0 2 4
6
8
Re(λ)
-4
-2
0
2
4
Im(λ)
(a) Wilson kernel
-2 0 2 4
6
8
Re(λ)
-4
-2
0
2
4
Im(λ)
(b) Staggered Wilson kernel
Figure 8.6.17:
Spectrum of
D
dw
in Boriçi’s construction at
N
s
=
4 on a smooth
topological configuration with Q = 3.
we show the numerically determined cumulative distribution functions (CDFs)
of
λ
min
for both the Wilson and staggered Wilson kernel at different values of
β
.
When increasing
β
, we observe how the distribution moves toward larger values of
λ
min
and pathologically small values become extremely unlikely. This is important,
as rational approximations to the
sign
-function become inaccurate for small argu-
ments. We also note that the corresponding CDFs for the Wilson and staggered
Wilson kernel are very similar.
115
8 Staggered domain wall fermions
−1
−0.5
0
0.5
1
−3 −2.5 −2 −1.5 −1 −0.5 0 0.5 1
λ(m)
Fermion mass m
f
(a) Unsmeared configuration
−1
−0.5
0
0.5
1
−3 −2.5 −2 −1.5 −1 −0.5 0 0.5 1
λ(m)
Fermion mass m
f
(b) Smeared configuration
−1
−0.5
0
0.5
1
−3 −2.5 −2 −1.5 −1 −0.5 0 0.5 1
λ(m)
Fermion mass m
f
(c) Topological configuration
Figure 8.6.18: Spectral flow of the Wilson kernel for configurations with Q = 2.
−1
−0.5
0
0.5
1
−3 −2.5 −2 −1.5 −1 −0.5 0 0.5 1
λ(m)
Fermion mass m
f
(a) Unsmeared configuration
−1
−0.5
0
0.5
1
−3 −2.5 −2 −1.5 −1 −0.5 0 0.5 1
λ(m)
Fermion mass m
f
(b) Smeared configuration
−1
−0.5
0
0.5
1
−3 −2.5 −2 −1.5 −1 −0.5 0 0.5 1
λ(m)
Fermion mass m
f
(c) Topological configuration
Figure 8.6.19:
Spectral flow of the staggered Wilson kernel for configurations with
Q = 2.
Finally, in Figs. 8.6.21 and 8.6.22 we show two examples of the
β
-dependence of
our chiral symmetry measures
N
and
GW
. In the plot we show the median value
and the width of the distribution. Fixing the value
q =
1
2
1 erf
1
2
, (8.6.6)
where
erf
(
x
)
refers to the error function, then we define the width using the
q
-
quantile and the
(
1 q
)
-quantile. With our choice of
q
, we find 68
.
3 % of the values
within the range. Our figures illustrate the superior chiral properties of the effective
operator with a staggered Wilson kernel at sufficiently large values of β. With the
use of smearing, the chiral properties are even further improved.
116
8.6 Numerical results
Kernel Construction N
s
m
eff
N
GW
2 5.47 × 10
3
1.53 × 10
1
6.72 × 10
1
Standard 4 5.90 × 10
4
6.87 × 10
2
3.10 × 10
1
6 1. 00 × 10
4
2.30 × 10
2
1.02 × 10
1
2 5.73 × 10
3
3.71 × 10
1
1.16 × 10
0
Wilson Boriçi 4 8.56 × 10
5
6.64 × 10
2
3.00 × 10
1
6 5.42 × 10
6
1.57 × 10
2
7.60 × 10
2
2 8.30 × 10
3
7.51 × 10
1
1.16 × 10
0
Optimal 4 2.11 × 10
3
3.59 × 10
2
4.78 × 10
2
6 8.71 × 10
6
1.21 × 10
3
1.74 × 10
3
2 6.22 × 10
3
1.48 × 10
1
6.49 × 10
1
Standard 4 7.18 × 10
4
5.88 × 10
2
2.88 × 10
1
6 1.34 × 10
4
2.01 × 10
2
9.72 × 10
2
2 6.38 × 10
3
2.02 × 10
1
2.92 × 10
1
Staggered Wilson Boriçi 4 5.13 × 10
5
5.45 × 10
3
8.68 × 10
3
6 5.91 × 10
7
1.53 × 10
4
2.69 × 10
4
2 1.72 × 10
2
2.32 × 10
1
2.66 × 10
1
Optimal 4 3.35 × 10
5
2.23 × 10
3
2.63 × 10
3
6 5.02 × 10
8
2.01 × 10
5
2.36 × 10
5
Table 8.6.1:
Median values for the various measures of chirality on unsmeared 32
2
configurations generated at
β =
12
.
8. For
m
eff
only configurations with
Q , 0 are considered.
117
8 Staggered domain wall fermions
0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
CDF(λ
min
)
λ
min
β = 3.2
β = 5.0
β = 12.8
(a) Wilson kernel
0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
CDF(λ
min
)
λ
min
β = 3.2
β = 5.0
β = 12.8
(b) Staggered Wilson kernel
Figure 8.6.20:
Cumulative distribution function
CDF
of
λ
min
for the Hermitian ker-
nel operators.
0
1
2
3
4
5
6
7
8
0 0.2 0.4 0.6 0.8 1 1.2 1.4
N
1 / β
Wilson
Staggered Wilson
(a) Without smearing
0
0.5
1
1.5
2
2.5
0 0.2 0.4 0.6 0.8 1 1.2 1.4
N
1 / β
Wilson
Staggered Wilson
(b) With smearing
Figure 8.6.21:
Violation of normality
N
as a function of 1
/β
for
%D
eff
in the optimal
construction at N
s
= 2 (this figure is not part of Ref. [2]).
0
0.5
1
1.5
2
2.5
3
0 0.2 0.4 0.6 0.8 1 1.2 1.4
GW
1 / β
Wilson
Staggered Wilson
(a) Without smearing
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 0.2 0.4 0.6 0.8 1 1.2 1.4
GW
1 / β
Wilson
Staggered Wilson
(b) With smearing
Figure 8.6.22:
Violation of the Ginsparg-Wilson relation
GW
as a function of 1
/β
for %D
eff
in Boriçi’s construction at N
s
= 4.
118
8.7 Quenched quantum chromodynamics
0
0.5
1
1.5
2
2.5
3
Re(λ)
-1.5
-1
-0.5
0
0.5
1
1.5
Im(λ)
4.23×10
-1
1.42×10
0
2.38×10
0
3.99×10
0
2.57×10
0
2.60×10
0
N
GW
std
Bor
opt
(a) Wilson kernel
0
0.5
1
1.5
2
2.5
3
Re(λ)
-1.5
-1
-0.5
0
0.5
1
1.5
Im(λ)
8.07×10
-1
2.12×10
0
1.29×10
0
1.43×10
0
2.60×10
-1
4.07×10
-1
N
GW
std
Bor
opt
(b) Staggered Wilson kernel
Figure 8.7.1:
Spectrum of
%D
eff
at
N
s
=
4 on a smeared 6
4
configuration in QCD
4
at β = 6.
8.7 Quenched quantum chromodynamics
Although our study focuses on the setting of the Schwinger model, we also in-
vestigated the case of four-dimensional quantum chromodynamics. Our results
presented here are exploratory in nature, go beyond our discussion in Ref. [
2
] and
were first presented in Ref. [1].
In Figs. 8.7.1 and 8.7.2 we can find the eigenvalue spectrum of the effective Dirac
operator in the standard, Boriçi’s and the optimal construction at
N
s
{
4, 8
}
on a smeared four-dimensional 6
4
lattice. The
SU
(
3
)
gauge configuration under
consideration was generated at
β =
6 and then smeared with one step of the
 smearing prescription using a smearing parameter of
α =
0
.
65. As it is a
topologically trivial configuration, the effective mass
m
eff
is omitted in the figure
labels. Taking smearing into account, for the domain wall height we use M
0
= 1.4
in the case of the Wilson kernel and M
0
= 1 for the staggered Wilson kernel.
While in the case of the standard construction with a Wilson kernel we actually
observe smaller chiral symmetry violations compared to the staggered Wilson
kernel, for Boriçi’s and the optimal construction the use of the staggered Wilson
kernel results in clearly superior chiral properties. We often observe a reduction in
119
8 Staggered domain wall fermions
0
0.5
1
1.5
2
2.5
3
Re(λ)
-1.5
-1
-0.5
0
0.5
1
1.5
Im(λ)
8.33×10
-2
2.73×10
-1
5.39×10
-1
1.10×10
0
2.78×10
-2
3.15×10
-2
N
GW
std
Bor
opt
(a) Wilson kernel
0
0.5
1
1.5
2
2.5
3
Re(λ)
-1.5
-1
-0.5
0
0.5
1
1.5
Im(λ)
3.41×10
-1
7.71×10
-1
8.66×10
-2
8.94×10
-2
6.37×10
-4
9.55×10
-4
N
GW
std
Bor
opt
(b) Staggered Wilson kernel
Figure 8.7.2:
Spectrum of
%D
eff
at
N
s
=
8 on a smeared 6
4
configuration in QCD
4
at β = 6.
N
and
GW
by more than an order of magnitude.
Although without considering more samples it is too early to judge the potential
benefits of using staggered domain wall fermions in the setting of four-dimensional
quantum chromodynamics, our first results are very encouraging.
8.8 Optimal weights
Before ending this chapter with our conclusions, we quote some example weight
factors
ω
s
for optimal domain wall fermions as defined in Sec. 8.3.3. We consider
the free-field case in two dimensions and set
M
0
=
1 and
N
s
=
8. In the case of the
Wilson kernel we then fix
λ
min
=
1 and
λ
max
=
3, while for the staggered Wilson
kernel we let
λ
min
=
1 and
λ
max
=
2
. The resulting weights
ω
s
can be found in
Table 8.8.1.
120
8.9 Conclusions
s ω
s
(
λ
min
= 1, λ
max
= 3
)
ω
s
λ
min
= 1, λ
max
=
2
1 0.989011284192743 0.996659816028010
2 0.908522120246430 0.971110743060917
3 0.779520722603956 0.925723982088869
4 0.641124364053574 0.869740043520870
5 0.519919928211430 0.813009342796322
6 0.427613177773962 0.763841917102527
7 0.366896221792507 0.728142270322088
8 0.337036936444470 0.709476563432225
Table 8.8.1: Weight factors ω
s
for the optimal domain wall fermion construction.
8.9 Conclusions
In this chapter, we explicitly constructed various formulations of staggered domain
wall fermions. We investigated their chiral properties in the free-field case, on
quenched thermalized background fields in the Schwinger model and on smooth
topological configurations. With respect to our study, staggered domain wall fer-
mions work as advertised and generalize the domain wall fermion construction to
staggered kernels.
We also went beyond Adams original proposal of staggered domain wall fermi-
ons and generalized existing variants to the staggered case, such as Boriçi’s and
Chius construction. This resulted in the formulation of truncated staggered do-
main wall fermions and optimal staggered domain wall fermions, which were pre-
viously not considered in the literature. Especially these new variants of staggered
domain wall fermions show significantly improved chiral properties compared to
the traditional Wilson-based constructions, at least in the context of the Schwinger
model. In a setting where chiral symmetry plays an important role, the use of these
novel lattice fermion formulations can be potentially very advantageous.
While our results in the setting of a U
(
1
)
gauge theory are very encouraging,
the case of a
SU
(
3
)
gauge group is not fully understood yet. As our first results in
four-dimensional quantum chromodynamics are promising, they warrant further
investigations.
121
C
C
In this thesis we aimed for a better understanding of the theoretical foundations,
computational properties and the practical applications of staggered Wilson, stag-
gered domain wall and staggered overlap fermions.
9.1 Summary
To summarize some of the core parts of this thesis we recall that in Chapter 3,
we gave an extensive review of the symmetries of staggered fermions and the
construction of flavored mass terms. We discussed both Adams proposal for a
two-flavor and Hoelblings proposal for a one-flavor staggered Wilson term. In Sec.
3.6, we built upon these ideas and generalized the staggered Wilson term to allow
for arbitrary mass splittings in arbitrary even dimensions.
In Chapter 4, we estimated the gains in computational efficiency when using
staggered Wilson fermions instead of usual Wilson fermions for inverting the Dirac
matrix on a source, where we found a speedup factor of 4
6 in our particular
benchmark setting. We also analyzed the memory bandwidth requirements of
staggered Wilson fermions, quantified by the arithmetic intensity, and found that
they are comparable to other common staggered formulations.
In Chapter 5, we demonstrated the feasibility of spectroscopy calculations with
staggered Wilson fermions by studying the case of pseudoscalar mesons. We
adapted the operators used for spectroscopy with usual staggered fermions and
did a numerical study of the pion spectrum when turning the two-flavor staggered
Wilson term on.
In Chapter 6, we reviewed the overlap construction and discussed Adams gen-
eralization to the case of a staggered kernel. In this context we did an analytical
verification of the correctness of the continuum limit of the index and axial anomaly
for the staggered overlap Dirac operator in Sec. 6.4.
In Chapter 7, we analyzed the eigenvalue spectrum of the staggered Wilson
Dirac operator and related it to the computational efficiency of staggered overlap
123
9 Conclusions
fermions. We proposed an explanation to resolve the apparent discrepancies in
the speedup factors for the staggered Wilson kernel and for the staggered overlap
formulation.
Finally, in Chapter 8 we studied Adams proposal of staggered domain wall fer-
mions. In the process we generalized some variants of domain wall fermions to a
staggered kernel, investigated spectral properties and quantified chiral symmetry
violations of various domain wall formulations. We found that in the setting of
the Schwinger model the use of a staggered Wilson kernel results in significantly
improved chiral properties.
9.2 Outlook
While staggered Wilson fermions and their derived formulations have both ad-
vantages and disadvantages compared to more established lattice fermion for-
mulations, we believe that their use can be beneficial in some settings. In four
dimensions, the two flavors of Adams staggered Wilson fermions can be naturally
used as a discretization of the light up and down quarks. Moreover, compared to
Wilson fermions, staggered Wilson fermions are at no disadvantage with respect to
flavor-singlet physics [
4
]. Possible applications include the high precision determi-
nation of the
η
0
mass and the calculation of bulk quantities in thermodynamics
[
4
,
9
]. Furthermore, one can test universality, i.e. the independence of physical
results of the choice of the lattice action, with the help of new lattice fermion
formulations.
While the staggered Wilson kernel defines an interesting lattice fermion dis-
cretization in its own right, its derived chiral fermion formulations bear the poten-
tial of reducing the computational costs of simulations with chiral lattice fermions
significantly. While staggered overlap fermions obey an exact chiral symmetry,
staggered domain wall fermions implement an approximate chiral symmetry in
a controlled manner. While the achievable speedup factor of staggered overlap
fermions compared to Neubergers overlap fermions has yet to be determined in a
realistic setting (cf. Chapter 7), for staggered domain wall fermions an improve-
ment of chirality by often more than an order of magnitude could be already shown
in the Schwinger model.
Besides Adams formulation, one may also ask about the potential of Hoelblings
single flavor staggered action (see Sec. 3.5). As already pointed out in Ref. [
46
], it is
124
9.2 Outlook
important to note that in this case the hypercubic rotational symmetry is broken,
although the double rotation symmetry is preserved. This results in the need for two
gluonic counterterms to restore rotational invariance in unquenched simulations
[
69
] as the terms
G
2
12
+ G
2
34
,
G
2
13
+ G
2
24
and
G
2
14
+ G
2
23
can all appear with different
coefficients, where
G
µν
denotes the clover version of the gauge field-strength tensor
(see e.g. Ref. [16]).
While this fine-tuning is in principle possible, the simultaneous fine-tuning of
coefficients requires significant effort and makes the use of this action unlikely in
the context of practical applications in the near future. Perturbative methods allow
studying this fine-tuning on an analytical level, although we expect the necessary
calculations to quickly become tedious due to the nontrivial structure of terms
of the form
G
2
α β
+ G
2
γδ
with
α
,
β
,
γ
,
δ
pairwise different. While these issues do
not render Hoelbling’s formulation invalid, one can get an idea of the challenges
ahead from the many-year efforts needed in studying similar issues in the context
of minimally doubled fermions of the Karsten-Wilczek kind [
221
,
222
], both on an
analytical [223] and a numerical level [224].
For the future, we hope that the practical potential of these new novel fermion
formulations is explored further, also in the case of full dynamical simulations. As
the technical properties of staggered Wilson fermions mimic the ones of Wilson
fermions (with the substitution
D
w
D
sw
and
γ
5
), it is expected that the
commonly employed algorithms for Wilson fermions can be adapted to staggered
Wilson fermions.
In particular rational and polynomial hybrid Monte Carlo methods [
225
227
]
would then allow the generation of two-flavor full QCD configurations and render
any rooting procedure—together with any potential related problems—unnecessary.
However, practical implementations of these algorithms together with in-depth
studies of their theoretical and computational properties are still outstanding and
left for future work.
125
B
[1] C. Hoelbling and C. Zielinski, “Staggered domain wall fermions, PoS
LATTICE2016 (2016) 254, arXiv:1609.05114 [hep-lat].
[2] C. Hoelbling and C. Zielinski, “Spectral properties and chiral symmetry
violations of (staggered) domain wall fermions in the Schwinger model,
Phys. Rev. D94 no. 1, (2016) 014501, arXiv:1602.08432 [hep-lat].
[3] D. H. Adams, R. Har, Y. Jia, and C. Zielinski, “Continuum limit of the axial
anomaly and index for the staggered overlap Dirac operator: An overview,
PoS LATTICE2013 (2014) 462, arXiv:1312.7230 [hep-lat].
[4] D. H. Adams, D. Nogradi, A. Petrashyk, and C. Zielinski, “Computational
efficiency of staggered Wilson fermions: A first look, PoS LAT TICE2013
(2014) 353, arXiv:1312.3265 [hep-lat].
[5] K. G. Wilson, “Confinement of quarks, Phys. Rev. D10 (1974) 2445–2459.
[6] K. G. Wilson, “Quarks and strings on a lattice, in New Phenomena in
Subnuclear Physics: Part A, A. Zichichi, ed., pp. 69–142. Springer US, Boston,
MA, 1977. Proceedings of the First Half of the International School of
Subnuclear Physics, Erice, Sicily, July 11 August 1, 1975.
[7] C. Zielinski, “Computational Efficiency and Spectrum Results for Staggered
Wilson Fermions. Talk given at the 32nd International Symposium on
Lattice Field Theory, New York City, USA, 2014.
[8] C. Zielinski, “More Results on the Computational Efficiency of Staggered
Wilson Fermions. Talk given at the 9th International Conference on
Computational Physics, Singapore, Singapore, 2015.
[9]
D. H. Adams, “Update on staggered Wilson fermions. Talk given at the 32nd
International Symposium on Lattice Field Theory, New York City, USA, 2014.
[10]
C. Zielinski, “Pion spectrum for the 2-flavor staggered Wilson fermion. Talk
given at the 33rd International Symposium on Lattice Field Theory, Kobe,
Japan, 2015.
[11] R. Har, Index of lattice Dirac operators. PhD thesis, Nanyang Technological
University, 2014. https://dr.ntu.edu.sg/handle/10356/61803.
127
Bibliography
[12] C. Zielinski, “On the Eigenvalue Spectrum of (Staggered) Domain Wall
Fermions. Talk given at the Spring Meeting of the German Physical Society,
Hamburg, Germany, 2016.
[13] S. Weinberg, The Quantum theory of fields. Vol. 1: Foundations. Cambridge
University Press, 2005.
[14] S. Weinberg, The quantum theory of fields. Vol. 2: Modern applications.
Cambridge University Press, 2013.
[15] M. E. Peskin and D. V. Schroeder, An introduction to quantum field theory.
Advanced book program. Westview Press Reading, 1995.
[16] C. Gattringer and C. Lang, Quantum Chromodynamics on the Lattice: An
Introductory Presentation. Lecture Notes in Physics. Springer Berlin
Heidelberg, 2009.
[17] T. DeGrand and C. DeTar, Lattice Methods for Quantum Chromodynamics.
World Scientific, 2006.
[18] I. Montvay and G. Munster, Quantum fields on a lattice. Cambridge
University Press, 1997.
[19] H. J. Rothe, Lattice Gauge Theories: An Introduction. World Scientific, 2005.
[20] P. A. M. Dirac, “The quantum theory of the electron, Proc. Roy. Soc. Lond.
A117 (1928) 610–624.
[21]
F. A. Berezin, The method of second quantization. Pure and applied physics.
Academic Press, 1966.
[22] C.-N. Yang and R. L. Mills, “Conservation of Isotopic Spin and Isotopic
Gauge Invariance, Phys. Rev. 96 (1954) 191–195.
[23] R. P. Feynman, “Space-time approach to nonrelativistic quantum
mechanics, Rev. Mod. Phys. 20 (1948) 367–387.
[24] R. P. Feynman, “Space-time approach to quantum electrodynamics, Phys.
Rev. 76 (1949) 769–789.
[25] S. Albeverio, R. Høegh-Krohn, and S. Mazzucchi, Mathematical Theory of
Feynman Path Integrals: An Introduction. Lecture Notes in Mathematics.
Springer Berlin Heidelberg, 2008.
[26]
L. D. Faddeev and V. N. Popov, “Feynman Diagrams for the Yang-Mills Field,
Phys. Lett. B25 (1967) 29–30.
128
Bibliography
[27] T. Aoyama, M. Hayakawa, T. Kinoshita, and M. Nio, “Tenth-Order QED
Contribution to the Electron g 2 and an Improved Value of the Fine
Structure Constant, Phys. Rev. Lett. 109 (2012) 111807, arXiv:1205.5368
[hep-ph].
[28] G. C. Wick, “Properties of Bethe-Salpeter Wave Functions, Phys. Rev. 96
(1954) 1124–1134.
[29] N. Sadooghi and H. J. Rothe, “Continuum behavior of lattice QED,
discretized with one sided lattice differences, in one loop order, Phys. Rev.
D55 (1997) 6749–6759, arXiv:hep-lat/9610001 [hep-lat].
[30] H. B. Nielsen and M. Ninomiya, “No Go theorem for regularizing chiral
fermions, Phys. Lett. B105 (1981) 219.
[31]
H. B. Nielsen and M. Ninomiya, “Absence of neutrinos on a lattice: (I). Proof
by Homotopy theory, Nucl. Phys.
B185
(1981) 20. Erratum: Nucl. Phys. B195,
541 (1982).
[32] H. B. Nielsen and M. Ninomiya, “Absence of neutrinos on a lattice: (II).
Intuitive topological proof, Nucl. Phys. B193 (1981) 173.
[33] D. Friedan, “A proof of the Nielsen-Ninomiya theorem, Commun. Math.
Phys. 85 (1982) 481–490.
[34] J. B. Kogut and L. Susskind, “Hamiltonian formulation of Wilsons lattice
gauge theories, Phys.Rev. D11 (1975) 395.
[35]
T. Banks, L. Susskind, and J. B. Kogut, “Strong coupling calculations of lattice
gauge theories: (1+1)-dimensional exercises, Phys.Rev. D13 (1976) 1043.
[36] Cornell-Oxford-Tel Aviv-Yeshiva Collaboration, T. Banks, S. Raby,
L. Susskind, J. Kogut, D. R. T. Jones, P. N. Scharbach, and D. K. Sinclair,
“Strong coupling calculations of the hadron spectrum of quantum
chromodynamics, Phys.Rev. D15 (1977) 1111.
[37] L. Susskind, “Lattice fermions, Phys.Rev. D16 (1977) 3031–3039.
[38] H. Kluberg-Stern, A. Morel, O. Napoly, and B. Petersson, “Flavors of
Lagrangian Susskind Fermions, Nucl. Phys. B220 (1983) 447–470.
[39] F. Gliozzi, “Spinor algebra of the one component lattice fermions, Nucl.
Phys. B204 (1982) 419–428.
[40] H. S. Sharatchandra, H. J. Thun, and P. Weisz, “Susskind Fermions on a
Euclidean Lattice, Nucl. Phys. B192 (1981) 205–236.
[41] C. van den Doel and J. Smit, “Dynamical Symmetry Breaking in Two Flavor
SU(N ) and SO(N ) Lattice Gauge Theories, Nucl. Phys. B228 (1983) 122–144.
129
Bibliography
[42] M. F. L. Golterman and J. Smit, “Selfenergy and flavor interpretation of
staggered fermions, Nucl. Phys. B245 (1984) 61.
[43] D. Daniel and T. D. Kieu, “On the Flavor Interpretations of Staggered
Fermions, Phys. Lett. B175 (1986) 73–76.
[44]
D. H. Adams, “Theoretical Foundation for the Index Theorem on the Lattice
with Staggered Fermions, Phys. Rev. Lett. 104 (2010) 141602,
arXiv:0912.2850 [hep-lat].
[45]
D. H. Adams, “Pairs of chiral quarks on the lattice from staggered fermions,
Phys. Lett. B699 (2011) 394–397, arXiv:1008.2833 [hep-lat].
[46] C. Hoelbling, “Single flavor staggered fermions, Phys. Lett. B696 (2011)
422–425, arXiv:1009.5362 [hep-lat].
[47]
Y. Iwasaki and T. Yoshie, “Renormalization Group Improved Action for SU(3)
Lattice Gauge Theory and the String Tension, Phys. Lett. B143 (1984)
449–452.
[48]
Y. Iwasaki, K. Kanaya, T. Kaneko, and T. Yoshie, “Scaling in SU(3) pure gauge
theory with a renormalization group improved action, Phys. Rev.
D56
(1997)
151–160, arXiv:hep-lat/9610023 [hep-lat].
[49] QCD-TARO Collaboration, P. de Forcrand, M. G. Perez, T. Hashimoto,
S. Hioki, H. Matsufuru, O. Miyamura, A. Nakamura, I. O. Stamatescu,
T. Takaishi, and T. Umeda, “Renormalization group flow of SU(3) lattice
gauge theory: Numerical studies in a two coupling space, Nucl. Phys.
B577
(2000) 263–278, arXiv:hep-lat/9911033 [hep-lat].
[50] A. Borici and R. Rosenfelder, “Scaling in SU(3) theory with a MCRG
improved lattice action, Nucl. Phys. Proc. Suppl. 63 (1998) 925–927,
arXiv:hep-lat/9711035 [hep-lat].
[51]
M. Golterman, Y. Shamir, and B. Svetitsky, “Localization properties of lattice
fermions with plaquette and improved gauge actions, Phys. Rev.
D72
(2005)
034501, arXiv:hep-lat/0503037 [hep-lat].
[52] R. Y. Rubinstein and D. P. Kroese, Simulation and the Monte Carlo method,
vol. 707. John Wiley & Sons, 2011.
[53] G. C. Wick, “The Evaluation of the Collision Matrix, Phys. Rev. 80 (1950)
268–272.
[54] K. Jansen, “Lattice QCD: A Critical status report, PoS LATTICE2008 (2008)
010, arXiv:0810.5634 [hep-lat].
130
Bibliography
[55] PACS-CS Collaboration, S. Aoki, N. Ishii, K.-I. Ishikawa, N. Ishizuka,
T. Izubuchi, D. Kadoh, K. Kanaya, Y. Kuramashi, Y. Namekawa, O. H. Nguyen,
M. Okawa, K. Sasaki, Y. Taniguchi, A. Ukawa, N. Ukita, T. Yamazaki, and
T. Yoshié, “Lattice quantum chromodynamics at the physical point and
beyond, PTEP 2012 (2012) 01A102.
[56] T. Misumi, “New fermion discretizations and their applications, PoS
LATTICE2012 (2012) 005, arXiv:1211.6999 [hep-lat].
[57] T. Misumi, Research on novel lattice fermions toward efficient QCD
simulations. PhD thesis, Kyoto University, 2012.
[58] D. H. Adams, “On the fourth root prescription for dynamical staggered
fermions, Phys. Rev. D72 (2005) 114512, arXiv:hep-lat/0411030
[hep-lat].
[59] A. Hasenfratz and R. Hoffmann, “Validity of the rooted staggered
determinant in the continuum limit, Phys. Rev. D74 (2006) 014511,
arXiv:hep-lat/0604010 [hep-lat].
[60]
C. Bernard, M. Golterman, Y. Shamir, and S. R. Sharpe, “Comment on Chiral
anomalies and rooted staggered fermions’, Phys. Lett.
B649
(2007) 235–240,
arXiv:hep-lat/0603027 [hep-lat].
[61] C. Bernard, M. Golterman, Y. Shamir, and S. R. Sharpe, “’t Hooft vertices,
partial quenching, and rooted staggered QCD, Phys. Rev.
D77
(2008) 114504,
arXiv:0711.0696 [hep-lat].
[62]
C. Bernard, M. Golterman, Y. Shamir, and S. R. Sharpe, “Reply to: ‘Comment
on: t Hooft vertices, partial quenching, and rooted staggered QCD’, Phys.
Rev. D78 (2008) 078502, arXiv:0808.2056 [hep-lat].
[63]
D. H. Adams, “The Rooting issue for a lattice fermion formulation similar to
staggered fermions but without taste mixing, Phys. Rev. D77 (2008) 105024,
arXiv:0802.3029 [hep-lat].
[64] M. Creutz, “Chiral anomalies and rooted staggered fermions, Phys. Lett.
B649 (2007) 230–234, arXiv:hep-lat/0701018 [hep-lat].
[65]
M. Creutz, “Reply to: Comment on: ‘Chiral anomalies and rooted staggered
fermions’”, Phys. Lett.
B649
(2007) 241–242,
arXiv:0704.2016 [hep-lat]
.
[66] M. Creutz, “Comment on t Hooft vertices, partial quenching, and rooted
staggered QCD’, Phys. Rev. D78 (2008) 078501, arXiv:0805.1350
[hep-lat].
[67]
S. R. Sharpe, “Rooted staggered fermions: Good, bad or ugly?, PoS
LAT2006
(2006) 022, arXiv:hep-lat/0610094 [hep-lat].
131
Bibliography
[68] M. Golterman, “QCD with rooted staggered fermions, PoS
CONFINEMENT8 (2008) 014, arXiv:0812.3110 [hep-ph].
[69]
S. R. Sharpe, “Comments on non-degenerate staggered fermions, staggered
Wilson and Overlap fermions, and the application of chiral perturbation
theory to lattice fermions. Talk given at Yukawa Institute Workshop “New
Types of Fermions on the Lattice, Kyoto, Japan, 2012.
[70] P. de Forcrand, A. Kurkela, and M. Panero, “Numerical properties of
staggered overlap fermions, PoS LATTICE2010 (2010) 080,
arXiv:1102.1000 [hep-lat].
[71] P. de Forcrand, A. Kurkela, and M. Panero, “Numerical properties of
staggered quarks with a taste-dependent mass term, JHEP 04 (2012) 142,
arXiv:1202.1867 [hep-lat].
[72]
S. Aoki, “New Phase Structure for Lattice QCD with Wilson Fermions, Phys.
Rev. D30 (1984) 2653.
[73]
S. Aoki, “The U(1) Problem on a Lattice. 1. Strong Coupling Limit, Phys. Rev.
D33 (1986) 2399.
[74]
S. Aoki, “The U(1) Problem on a Lattice. 2. Strong Coupling Expansion, Phys.
Rev. D34 (1986) 3170.
[75] S. Aoki, “A Solution to the U(1) Problem on a Lattice, Phys. Rev. Lett. 57
(1986) 3136.
[76] S. Aoki, “U(1) Problem and Lattice QCD, Nucl. Phys. B314 (1989) 79–111.
[77] M. Creutz, “Wilson fermions at finite temperature, in RHIC Summer Study
96: Brookhaven Theory Workshop on Relativistic Heavy Ions Upton, New
York, July 8-19, 1996. 1996. arXiv:hep-lat/9608024 [hep-lat].
[78]
S. R. Sharpe and R. L. Singleton, Jr, “Spontaneous flavor and parity breaking
with Wilson fermions, Phys. Rev. D58 (1998) 074501,
arXiv:hep-lat/9804028 [hep-lat].
[79]
T. Misumi, M. Creutz, T. Kimura, T. Z. Nakano, and A. Ohnishi, “Aoki Phases
in Staggered-Wilson Fermions, PoS LATTICE2011 (2011) 108,
arXiv:1110.1231 [hep-lat].
[80] T. Misumi, T. Z. Nakano, T. Kimura, and A. Ohnishi, “Strong-coupling
Analysis of Parity Phase Structure in Staggered-Wilson Fermions, Phys. Rev.
D86 (2012) 034501, arXiv:1205.6545 [hep-lat].
[81] T. Z. Nakano, T. Misumi, T. Kimura, and A. Ohnishi, “Strong coupling
analysis of Aoki phase in Staggered-Wilson fermions, PoS LATTICE2012
(2012) 203, arXiv:1210.5954 [hep-lat].
132
Bibliography
[82]
S. Durr, “Taste-split staggered actions: eigenvalues, chiralities and Symanzik
improvement, Phys. Rev. D87 no. 11, (2013) 114501, arXiv:1302.0773
[hep-lat].
[83]
C. Lanczos, “An iteration method for the solution of the eigenvalue problem
of linear differential and integral operators, J. Res. Natl. Bur. Stand. B Math.
Sci. 45 (1950) 255–282.
[84] M. R. Hestenes and E. Stiefel, Methods of conjugate gradients for solving
linear systems, vol. 49. NBS, 1952.
[85]
J. R. Shewchuk, An introduction to the conjugate gradient method without
the agonizing pain, 1994.
[86] M. G. Alford, T. R. Klassen, and G. P. Lepage, “Improving lattice quark
actions, Nucl. Phys. B496 (1997) 377–407, arXiv:hep-lat/9611010
[hep-lat].
[87] SciDAC, LHPC, UKQCD Collaboration, R. G. Edwards and B. Joo, “The
Chroma software system for lattice QCD, Nucl. Phys. Proc. Suppl. 140
(2005) 832, arXiv:hep-lat/0409003 [hep-lat].
[88] D. H. Adams, “New staggered-type fermions on the lattice: a review. Talk
given at Yukawa Institute Workshop “New Types of Fermions on the Lattice,
Kyoto, Japan, 2012.
[89] M. H. Quenouille, “Problems in plane sampling, Ann. Math. Statist. 20
no. 3, (1949) 355–375.
[90] M. H. Quenouille, “Notes on bias in estimation, Biometrika 43 no. 3-4,
(1956) 353–360.
[91] J. W. Tukey, Bias and confidence in not-quite large samples, in Annals of
Mathematical statistics, vol. 29, pp. 614–614. 1958.
[92] J. Jeffers and J. Reinders, High Performance Parallelism Pearls Volume Two:
Multicore and Many-core Programming Approaches. Elsevier Science, 2015.
[93] P. A. Boyle, “The BAGEL assembler generation library, Comput. Phys.
Commun. 180 (2009) 2739–2748.
[94]
R. Brower, “Quarks, GPUs and Multigrid. Talk given at the summer institute
“Scientific Computing in the Americas: the challenge of massive parallelism,
Valparaiso, Chile, 2011.
[95] M. Wagner, “GPUs Immediately Relating Lattice QCD to Collider
Experiments. Talk given at the GPU Technology Conference 2013”, San José,
USA, 2013.
133
Bibliography
[96] HPQCD, UKQCD Collaboration, E. Follana, Q. Mason, C. Davies,
K. Hornbostel, G. P. Lepage, J. Shigemitsu, H. Trottier, and K. Wong, “Highly
improved staggered quarks on the lattice, with applications to charm
physics, Phys. Rev. D75 (2007) 054502, arXiv:hep-lat/0610092
[hep-lat].
[97] MILC Collaboration, K. Orginos, D. Toussaint, and R. L. Sugar, “Variants of
fattening and flavor symmetry restoration, Phys. Rev. D60 (1999) 054503,
arXiv:hep-lat/9903032 [hep-lat].
[98] M. F. L. Golterman, “Staggered Mesons, Nucl. Phys. B273 (1986) 663.
[99] W.-J. Lee and S. R. Sharpe, “Partial flavor symmetry restoration for chiral
staggered fermions, Phys. Rev. D60 (1999) 114503,
arXiv:hep-lat/9905023 [hep-lat].
[100] T. Bae, D. H. Adams, C. Jung, H.-J. Kim, 2, J. Kim, K. Kim, W. Lee, and S. R.
Sharpe, “Taste symmetry breaking with HYP-smeared staggered fermions,
Phys. Rev. D77 (2008) 094508, arXiv:0801.3000 [hep-lat].
[101] N. Ishizuka, M. Fukugita, H. Mino, M. Okawa, and A. Ukawa, “Operator
dependence of hadron masses for Kogut-Susskind quarks on the lattice,
Nucl. Phys. B411 (1994) 875–902.
[102] R. Narayanan and H. Neuberger, “Infinitely many regulator fields for chiral
fermions, Phys. Lett. B302 (1993) 62–69, arXiv:hep-lat/9212019
[hep-lat].
[103] R. Narayanan and H. Neuberger, “Chiral fermions on the lattice, Phys. Rev.
Lett. 71 (1993) 3251–3254, arXiv:hep-lat/9308011 [hep-lat].
[104] R. Narayanan and H. Neuberger, “Chiral determinant as an overlap of two
vacua, Nucl. Phys. B412 (1994) 574–606, arXiv:hep-lat/9307006
[hep-lat].
[105] R. Narayanan and H. Neuberger, “A construction of lattice chiral gauge
theories, Nucl. Phys. B443 (1995) 305–385, arXiv:hep-th/9411108
[hep-th].
[106] H. Neuberger, “Exactly massless quarks on the lattice, Phys. Lett. B417
(1998) 141–144, arXiv:hep-lat/9707022 [hep-lat].
[107] H. Neuberger, “A Practical Implementation of the Overlap Dirac Operator,
Phys. Rev. Lett. 81 (1998) 4060–4062, arXiv:hep-lat/9806025 [hep-lat].
[108] H. Neuberger, “More about exactly massless quarks on the lattice, Phys.
Lett. B427 (1998) 353–355, arXiv:hep-lat/9801031 [hep-lat].
134
Bibliography
[109] P. H. Ginsparg and K. G. Wilson, “A remnant of chiral symmetry on the
lattice, Phys. Rev. D25 (1982) 2649.
[110] P. Hasenfratz, V. Laliena, and F. Niedermayer, “The index theorem in QCD
with a finite cutoff, Phys. Lett. B427 (1998) 125–131,
arXiv:hep-lat/9801021 [hep-lat].
[111] M. Luscher, “Exact chiral symmetry on the lattice and the Ginsparg-Wilson
relation, Phys. Lett. B428 (1998) 342–345, arXiv:hep-lat/9802011
[hep-lat].
[112] D. H. Adams, “Index and overlap construction for staggered fermions, PoS
LATTICE2010 (2010) 073, arXiv:1103.6191 [hep-lat].
[113] P. Hasenfratz, “Prospects for perfect actions, Nucl. Phys. Proc. Suppl. 63
(1998) 53–58, arXiv:hep-lat/9709110 [hep-lat].
[114] I. Horvath, “Ginsparg-Wilson relation and ultralocality, Phys. Rev. Lett. 81
(1998) 4063–4066, arXiv:hep-lat/9808002 [hep-lat].
[115]
P. Hernandez, K. Jansen, and M. Luscher, “Locality properties of Neuberger’s
lattice Dirac operator, Nucl. Phys. B552 (1999) 363–378,
arXiv:hep-lat/9808010 [hep-lat].
[116]
J. van den Eshof, A. Frommer, T. Lippert, K. Schilling, and H. A. van der Vorst,
“A Comparative study of numerical methods for the overlap Dirac operator:
A Status report, Nucl. Phys. Proc. Suppl. 106 (2002) 1070–1072,
arXiv:hep-lat/0110198 [hep-lat].
[117]
J. van den Eshof, A. Frommer, T. Lippert, K. Schilling, and H. A. van der Vorst,
“Numerical methods for the QCD overlap operator. I. Sign function and error
bounds, Comput. Phys. Commun. 146 (2002) 203–224,
arXiv:hep-lat/0202025 [hep-lat].
[118] G. Arnold, N. Cundy, J. van den Eshof, A. Frommer, S. Krieg, T. Lippert, and
K. Schafer, “Numerical methods for the QCD overlap operator. 2. Optimal
Krylov subspace methods, arXiv:hep-lat/0311025 [hep-lat].
[119] N. Cundy, J. van den Eshof, A. Frommer, S. Krieg, T. Lippert, and K. Schafer,
“Numerical methods for the QCD overlap operator. 3. Nested iterations,
Comput. Phys. Commun. 165 (2005) 221–242, arXiv:hep-lat/0405003
[hep-lat].
[120] N. Cundy, S. Krieg, G. Arnold, A. Frommer, T. Lippert, and K. Schilling,
“Numerical methods for the QCD overlap operator IV: Hybrid Monte Carlo,
Comput. Phys. Commun. 180 (2009) 26–54, arXiv:hep-lat/0502007
[hep-lat].
135
Bibliography
[121] M. F. Atiyah and I. M. Singer, “The index of elliptic operators: I, Annals of
Mathematics 87 no. 3, (1968) 484–530.
[122]
M. F. Atiyah and I. M. Singer, “The index of elliptic operators: III, Annals of
Mathematics 87 no. 3, (1968) 546–604.
[123]
M. F. Atiyah and I. M. Singer, “The index of elliptic operators: IV, Annals of
Mathematics 93 no. 1, (1971) 119–138.
[124] M. F. Atiyah and I. M. Singer, “The index of elliptic operators: V, Annals of
Mathematics 93 no. 1, (1971) 139–149.
[125] J. E. Kiskis, “Fermions in a pseudoparticle field, Phys. Rev. D15 (1977) 2329.
[126] N. K. Nielsen and B. Schroer, “Axial anomaly and Atiyah-Singer theorem,
Nucl. Phys. B127 (1977) 493.
[127] M. M. Ansourian, “Index theory and the axial current anomaly in
two-dimensions, Phys. Lett. B70 (1977) 301.
[128] S. L. Adler, “Axial vector vertex in spinor electrodynamics, Phys. Rev. 177
(1969) 2426–2438.
[129] J. S. Bell and R. Jackiw, “A PCAC puzzle: π
0
γγ in the σ-model, Nuovo
Cim. A60 (1969) 47–61.
[130]
K. Fujikawa, “Path Integral Measure for Gauge Invariant Fermion Theories,
Phys. Rev. Lett. 42 (1979) 1195–1198.
[131] K. Fujikawa, “Path Integral for Gauge Theories with Fermions, Phys. Rev.
D21 (1980) 2848. Erratum: Phys. Rev. D22, 1499 (1980).
[132] K. Fujikawa, “Comment on Chiral and Conformal Anomalies, Phys. Rev.
Lett. 44 (1980) 1733.
[133] E. Witten, “Current Algebra Theorems for the U(1) Goldstone Boson, Nucl.
Phys. B156 (1979) 269–283.
[134] G. Veneziano, “U(1) Without Instantons, Nucl. Phys. B159 (1979) 213–224.
[135] G. Veneziano, “Goldstone Mechanism From Gluon Dynamics, Phys. Lett.
B95 (1980) 90–92.
[136] L. Giusti, G. C. Rossi, M. Testa, and G. Veneziano, “The U(A)(1) problem on
the lattice with Ginsparg-Wilson fermions, Nucl. Phys.
B628
(2002) 234–252,
arXiv:hep-lat/0108009 [hep-lat].
[137] E. Seiler, “Some more remarks on the Witten-Veneziano formula for the η
0
mass, Phys. Lett.
B525
(2002) 355–359,
arXiv:hep-th/0111125 [hep-th]
.
136
Bibliography
[138] L. Giusti, G. C. Rossi, and M. Testa, “Topological susceptibility in full QCD
with Ginsparg-Wilson fermions, Phys. Lett. B587 (2004) 157–166,
arXiv:hep-lat/0402027 [hep-lat].
[139] M. Luscher, “Topological effects in QCD and the problem of short distance
singularities, Phys. Lett. B593 (2004) 296–301, arXiv:hep-th/0404034
[hep-th].
[140] M. Aguado and E. Seiler, “The Clash of positivities in topological density
correlators, Phys. Rev. D72 (2005) 094502, arXiv:hep-lat/0503015
[hep-lat].
[141] Y. Kikukawa and A. Yamada, “Weak coupling expansion of massless QCD
with a Ginsparg-Wilson fermion and axial U(1) anomaly, Phys. Lett. B448
(1999) 265–274, arXiv:hep-lat/9806013 [hep-lat].
[142] K. Fujikawa, “A Continuum limit of the chiral Jacobian in lattice gauge
theory, Nucl. Phys. B546 (1999) 480–494, arXiv:hep-th/9811235
[hep-th].
[143]
H. Suzuki, “Simple evaluation of chiral Jacobian with overlap Dirac operator,
Prog. Theor. Phys. 102 (1999) 141–147, arXiv:hep-th/9812019 [hep-th].
[144]
D. H. Adams, “Axial anomaly and topological charge in lattice gauge theory
with overlap Dirac operator, Annals Phys. 296 (2002) 131–151,
arXiv:hep-lat/9812003 [hep-lat].
[145] D. H. Adams, “On the continuum limit of fermionic topological charge in
lattice gauge theory, J. Math. Phys. 42 (2001) 5522–5533,
arXiv:hep-lat/0009026 [hep-lat].
[146] A. Petrashyk, “Spectral flow of staggered Wilson Dirac operator in SU(2)
instanton backgrounds. Talk given at Yukawa Institute Workshop “New
Types of Fermions on the Lattice, Kyoto, Japan, 2012.
[147] D. H. Adams and W. Bietenholz, “Axial anomaly and index of the overlap
hypercube operator, Eur. Phys. J. C34 (2004) 245–253,
arXiv:hep-lat/0307022 [hep-lat].
[148] W. Bietenholz, “Solutions of the Ginsparg-Wilson relation and improved
domain wall fermions, Eur. Phys. J. C6 (1999) 537–547,
arXiv:hep-lat/9803023 [hep-lat].
[149] E. Anderson, Z. Bai, C. Bischof, S. Blackford, J. Demmel, J. Dongarra, J. D.
Croz, A. Greenbaum, S. Hammarling, A. McKenney, and D. Sorensen,
LAPACK Users Guide. Philadelphia, third ed., 1999.
137
Bibliography
[150] J. G. F. Francis, “The QR Transformation A Unitary Analogue to the LR
Transformation–Part 1, The Computer Journal 4 no. 3, (1961) 265–271.
[151] J. G. F. Francis, “The QR Transformation–Part 2, The Computer Journal 4
no. 4, (1962) 332–345.
[152] V. N. Kublanovskaya, “On some algorithms for the solution of the complete
eigenvalue problem, USSR Computational Mathematics and Mathematical
Physics 1 no. 3, (1962) 637–657.
[153] R. Lehoucq, D. Sorensen, and C. Yang, ARPACK Users Guide, 1998.
[154] R. B. Lehoucq and D. C. Sorensen, “Deflation Techniques for an Implicitly
Restarted Arnoldi Iteration, SIAM Journal on Matrix Analysis and
Applications 17 no. 4, (1996) 789–821.
[155] G. I. Egri, Z. Fodor, S. D. Katz, and K. K. Szabo, “Topology with dynamical
overlap fermions, JHEP 01 (2006) 049, arXiv:hep-lat/0510117
[hep-lat].
[156] JLQCD Collaboration, H. Fukaya, S. Hashimoto, K.-I. Ishikawa, T. Kaneko,
H. Matsufuru, T. Onogi, and N. Yamada, “Lattice gauge action suppressing
near-zero modes of H
W
, Phys. Rev. D74 (2006) 094505,
arXiv:hep-lat/0607020 [hep-lat].
[157] N. Cundy, S. Krieg, T. Lippert, and A. Schafer, “Topological tunneling with
dynamical overlap fermions, Comput. Phys. Commun. 180 (2009) 201–208,
arXiv:0803.0294 [hep-lat].
[158]
N. Cundy and W. Lee, “Modifying the molecular dynamics action to increase
topological tunnelling rate for dynamical overlap fermions,
arXiv:1110.1948 [hep-lat].
[159]
D. B. Kaplan, “A Method for simulating chiral fermions on the lattice, Phys.
Lett. B288 (1992) 342–347, arXiv:hep-lat/9206013 [hep-lat].
[160]
Y. Shamir, “Chiral fermions from lattice boundaries, Nucl. Phys.
B406
(1993)
90–106, arXiv:hep-lat/9303005 [hep-lat].
[161] V. Furman and Y. Shamir, “Axial symmetries in lattice QCD with Kaplan
fermions, Nucl. Phys. B439 (1995) 54–78, arXiv:hep-lat/9405004
[hep-lat].
[162] P. M. Vranas, “Chiral symmetry restoration in the Schwinger model with
domain wall fermions, Phys. Rev. D57 (1998) 1415–1432,
arXiv:hep-lat/9705023 [hep-lat].
138
Bibliography
[163] S. Aoki and Y. Taniguchi, “One loop calculation in lattice QCD with domain
wall quarks, Phys. Rev. D59 (1999) 054510, arXiv:hep-lat/9711004
[hep-lat].
[164] Y. Kikukawa, H. Neuberger, and A. Yamada, “Exponential suppression of
radiatively induced mass in the truncated overlap, Nucl. Phys. B526 (1998)
572–596, arXiv:hep-lat/9712022 [hep-lat].
[165]
P. Chen, N. H. Christ, G. T. Fleming, A. Kaehler, C. Malureanu, R. Mawhinney,
G. Siegert, C. zhong Sui, Y. Zhestkov, and P. M. Vranas, “Toward the chiral
limit of QCD: Quenched and dynamical domain wall fermions, in
High-energy physics. Proceedings, 29th International Conference, ICHEP’98,
Vancouver, Canada, July 23-29, 1998. Vol. 1, 2. 1998.
arXiv:hep-lat/9812011 [hep-lat].
[166]
T. Blum, “Domain wall fermions in vector gauge theories, Nucl. Phys. Proc.
Suppl. 73 (1999) 167–179, arXiv:hep-lat/9810017 [hep-lat].
[167]
G. T. Fleming, “How well do domain wall fermions realize chiral symmetry?,
Nucl. Phys. Proc. Suppl. 83 (2000) 363–365, arXiv:hep-lat/9909140
[hep-lat].
[168]
P. Hernandez, K. Jansen, and M. Luscher, “A Note on the practical feasibility
of domain wall fermions, in Workshop on Current Theoretical Problems in
Lattice Field Theory Ringberg, Germany, April 2–8. 2000.
arXiv:hep-lat/0007015 [hep-lat].
[169]
S. Aoki, T. Izubuchi, Y. Kuramashi, and Y. Taniguchi, “Domain wall fermions
in quenched lattice QCD, Phys. Rev. D62 (2000) 094502,
arXiv:hep-lat/0004003 [hep-lat].
[170] J. S. Schwinger, “Gauge Invariance and Mass. II, Phys. Rev. 128 (1962)
2425–2429.
[171]
A. M. Tsvelik, Quantum field theory in condensed matter physics. Cambridge
University Press, Cambridge, 2005.
[172] A. Borici, “Truncated overlap fermions, Nucl. Phys. Proc. Suppl. 83 (2000)
771–773, arXiv:hep-lat/9909057 [hep-lat].
[173]
R. C. Brower, H. Neff, and K. Orginos, “Möbius fermions: Improved domain
wall chiral fermions, Nucl. Phys. Proc. Suppl. 140 (2005) 686–688,
arXiv:hep-lat/0409118 [hep-lat].
[174] N. J. Higham, “The matrix sign decomposition and its relation to the polar
decomposition, Linear Algebra and its Applications 212 (1994) 3–20.
139
Bibliography
[175] T.-W. Chiu, “Optimal domain wall fermions, Phys. Rev. Lett. 90 (2003)
071601, arXiv:hep-lat/0209153 [hep-lat].
[176] T.-W. Chiu, “Locality of optimal lattice domain wall fermions, Phys. Lett.
B552 (2003) 97–100, arXiv:hep-lat/0211032 [hep-lat].
[177] E. I. Zolotarev, Application of elliptic functions to questions of functions
deviating least and most from zero, Zap. Imp. Akad. Nauk. St. Petersburg
30 no. 5, (1877) 1–59. reprinted in his Collected Works, Vol. II, Izdat. Akad.
Nauk SSSR, Moscow, 1932, pp. 1–59. In Russian.
[178]
N. I. Akhiezer, Elements of the Theory of Elliptic Functions, vol. 79. American
Mathematical Soc., Washington, DC, 1990.
[179] N. I. Achieser, Theory of approximation. Courier Corporation, 2013.
[180] T.-W. Chiu, T.-H. Hsieh, C.-H. Huang, and T.-R. Huang, “A Note on the
Zolotarev optimal rational approximation for the overlap Dirac operator,
Phys. Rev. D66 (2002) 114502, arXiv:hep-lat/0206007 [hep-lat].
[181]
M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions: with
Formulas, Graphs, and Mathematical Tables. Dover Books on Mathematics.
Dover Publications, 2012.
[182] TWQCD
Collaboration, Y.-C. Chen, T.-W. Chiu, T.-S. Guu, T.-H. Hsieh, C.-H.
Huang, and Y.-Y. Mao, “Lattice QCD with Optimal Domain-Wall Fermion:
Light Meson Spectroscopy, PoS LATTICE2010 (2010) 099,
arXiv:1101.0405 [hep-lat].
[183] TWQCD
Collaboration, T.-H. Hsieh, T.-W. Chiu, and Y.-Y. Mao, “Topological
Charge in Two Flavors QCD with Optimal Domain-Wall Fermion, PoS
LATTICE2010 (2010) 085, arXiv:1101.0402 [hep-lat].
[184] TWQCD Collaboration, T. W. Chiu, T. H. Hsieh, and Y. Y. Mao, “Topological
Susceptibility in Two Flavors Lattice QCD with the Optimal Domain-Wall
Fermion, Phys. Lett. B702 (2011) 131–134, arXiv:1105.4414 [hep-lat].
[185] TWQCD Collaboration, T.-W. Chiu, T.-H. Hsieh, and Y.-Y. Mao,
“Pseudoscalar Meson in Two Flavors QCD with the Optimal Domain-Wall
Fermion, Phys. Lett. B717 (2012) 420–424, arXiv:1109.3675 [hep-lat].
[186] TWQCD Collaboration, Y.-C. Chen and T.-W. Chiu, “Chiral Symmetry and
the Residual Mass in Lattice QCD with the Optimal Domain-Wall Fermion,
Phys. Rev. D86 (2012) 094508, arXiv:1205.6151 [hep-lat].
[187] TWQCD Collaboration, T.-W. Chiu and T.-H. Hsieh, “Lattice QCD with
optimal domain-wall fermion on the 20
3
× 40 lattice, PoS LATTICE2012
(2012) 205.
140
Bibliography
[188] T.-W. Chiu, “Domain-Wall Fermion with R
5
Symmetry, Phys. Lett. B744
(2015) 95–100, arXiv:1503.01750 [hep-lat].
[189] R. C. Brower, H. Neff, and K. Orginos, “Möbius fermions, Nucl. Phys. Proc.
Suppl. 153 (2006) 191–198, arXiv:hep-lat/0511031 [hep-lat].
[190] R. Brower, R. Babich, K. Orginos, C. Rebbi, D. Schaich, and P. Vranas,
“Moebius Algorithm for Domain Wall and GapDW Fermions, PoS
LATTICE2008 (2008) 034, arXiv:0906.2813 [hep-lat].
[191]
H. Neuberger, “Vectorlike gauge theories with almost massless fermions on
the lattice, Phys. Rev. D57 (1998) 5417–5433, arXiv:hep-lat/9710089
[hep-lat].
[192] Y. Kikukawa and T. Noguchi, “Low-energy effective action of domain wall
fermion and the Ginsparg-Wilson relation, Nuclear Physics B-Proceedings
Supplements 83 (2000) 630–632, arXiv:hep-lat/9902022 [hep-lat].
[193] Y. Kikukawa, “Locality bound for effective four-dimensional action of
domain wall fermion, Nucl. Phys. B584 (2000) 511–527,
arXiv:hep-lat/9912056 [hep-lat].
[194]
Y. Shamir, “Reducing chiral symmetry violations in lattice QCD with domain
wall fermions, Phys. Rev. D59 (1999) 054506, arXiv:hep-lat/9807012
[hep-lat].
[195] R. G. Edwards and U. M. Heller, “Domain wall fermions with exact chiral
symmetry, Phys. Rev. D63 (2001) 094505, arXiv:hep-lat/0005002
[hep-lat].
[196] P. Boyle, A. Juttner, M. K. Marinkovic, F. Sanfilippo, M. Spraggs, and J. T.
Tsang, “An exploratory study of heavy domain wall fermions on the lattice,
arXiv:1602.04118 [hep-lat].
[197] S. Durr, C. Hoelbling, and U. Wenger, “Filtered overlap: Speedup, locality,
kernel non-normality and Z
A
' 1, JHEP 09 (2005) 030,
arXiv:hep-lat/0506027 [hep-lat].
[198] M. Golterman and Y. Shamir, “Localization in lattice QCD, Phys. Rev. D68
(2003) 074501, arXiv:hep-lat/0306002 [hep-lat].
[199]
M. Golterman and Y. Shamir, “Localization in lattice QCD (with emphasis on
practical implications), Nucl. Phys. Proc. Suppl. 129 (2004) 149–155,
arXiv:hep-lat/0309027 [hep-lat].
[200] M. Golterman, Y. Shamir, and B. Svetitsky, “Mobility edge in lattice QCD,
Phys. Rev. D71 (2005) 071502, arXiv:hep-lat/0407021 [hep-lat].
141
Bibliography
[201] T. Blum, P. Chen, N. Christ, C. Cristian, C. Dawson, G. Fleming, A. Kaehler,
X. Liao, G. Liu, C. Malureanu, R. Mawhinney, S. Ohta, G. Siegert, A. Soni,
C. Sui, P. Vranas, M. Wingate, L. Wu, and Y. Zhestkov, “Quenched lattice QCD
with domain wall fermions and the chiral limit, Phys. Rev. D69 (2004)
074502, arXiv:hep-lat/0007038 [hep-lat].
[202]
T. Blum, N. Christ, C. Cristian, C. Dawson, G. Fleming, G. Liu, R. Mawhinney,
A. Soni, P. Vranas, M. Wingate, L. Wu, and Y. Zhestkov, “Nonperturbative
renormalization of domain wall fermions: Quark bilinears, Phys. Rev. D66
(2002) 014504, arXiv:hep-lat/0102005 [hep-lat].
[203]
V. Gadiyak, X.-D. Ji, and C.-W. Jung, “Domain wall induced quark masses in
topologically nontrivial background, Phys. Rev. D62 (2000) 074508,
arXiv:hep-lat/0002023 [hep-lat].
[204] C. Jung, R. G. Edwards, X.-D. Ji, and V. Gadiyak, “Residual chiral symmetry
breaking in domain wall fermions, Phys. Rev. D63 (2001) 054509,
arXiv:hep-lat/0007033 [hep-lat].
[205] C. Jung, R. G. Edwards, X.-D. Ji, and V. Gadiyak, “Residual chiral symmetry
breaking in domain wall fermions, Nucl. Phys. Proc. Suppl. 94 (2001)
748–751, arXiv:hep-lat/0010094 [hep-lat].
[206]
W. Kerler, “Dirac operator normality and chiral fermions, Chin. J. Phys.
38
(2000) 623–632, arXiv:hep-lat/9912022 [hep-lat].
[207] I. Hip, T. Lippert, H. Neff, K. Schilling, and W. Schroers, “Instanton
dominance of topological charge fluctuations in QCD?, Phys. Rev. D65
(2001) 014506, arXiv:hep-lat/0105001 [hep-lat].
[208]
I. Hip, T. Lippert, H. Neff, K. Schilling, and W. Schroers, “The Consequences
of non-normality, Nucl. Phys. Proc. Suppl. 106 (2002) 1004–1006,
arXiv:hep-lat/0110155 [hep-lat].
[209]
S. Durr and G. Koutsou, “Brillouin improvement for Wilson fermions, Phys.
Rev. D83 (2011) 114512, arXiv:1012.3615 [hep-lat].
[210] V. Azcoiti, G. D. Carlo, E. Follana, and A. Vaquero, “Topological Index
Theorem on the Lattice through the Spectral Flow of Staggered Fermions,
Phys. Lett. B744 (2015) 303–308, arXiv:1410.5733 [hep-lat].
[211]
J. Smit and J. C. Vink, “Remnants of the index theorem on the lattice, Nucl.
Phys. B286 (1987) 485.
[212]
S. Durr and C. Hoelbling, “Staggered versus overlap fermions: A study in the
Schwinger model with N
f
= 0, 1, 2, Phys. Rev. D69 (2004) 034503,
arXiv:hep-lat/0311002 [hep-lat].
142
Bibliography
[213] S. Durr and C. Hoelbling, “Scaling tests with dynamical overlap and rooted
staggered fermions, Phys. Rev. D71 (2005) 054501,
arXiv:hep-lat/0411022 [hep-lat].
[214] L. Giusti, C. Hoelbling, and C. Rebbi, “Schwinger model with the overlap
Dirac operator: Exact results versus a physics motivated approximation,
Phys. Rev. D64 (2001) 054501, arXiv:hep-lat/0101015 [hep-lat].
[215] S. Durr and C. Hoelbling, “Lattice fermions with complex mass, Phys. Rev.
D74 (2006) 014513, arXiv:hep-lat/0604005 [hep-lat].
[216] M. Falcioni, M. L. Paciello, G. Parisi, and B. Taglienti, “Again on SU(3)
glueball mass, Nucl. Phys. B251 (1985) 624–632.
[217] APE Collaboration, M. Albanese, F. Costantini, G. Fiorentini, F. Flore, M. P.
Lombardo, R. Tripiccione, P. Bacilieri, L. Fonti, P. Giacomelli, E. Remiddi,
M. Bernaschi, N. Cabibbo, E. Marinari, G. Parisi, G. Salina, S. Cabasino,
F. Marzano, P. Paolucci, S. Petrarca, F. Rapuano, P. Marchesini, and R. Rusack,
“Glueball Masses and String Tension in Lattice QCD, Phys. Lett.
B192
(1987)
163–169.
[218] S. Capitani, S. Durr, and C. Hoelbling, “Rationale for UV-filtered clover
fermions, JHEP 11 (2006) 028, arXiv:hep-lat/0607006 [hep-lat].
[219] S. Itoh, Y. Iwasaki, and T. Yoshie, “The U(1) problem and topological
excitations on a lattice, Phys. Rev. D36 (1987) 527.
[220] E. Follana, V. Azcoiti, G. D. Carlo, and A. Vaquero, “Spectral Flow and Index
Theorem for Staggered Fermions, PoS LATTICE2011 (2011) 100,
arXiv:1111.3502 [hep-lat].
[221] L. H. Karsten, “Lattice Fermions in Euclidean Space-time, Phys. Lett. B104
(1981) 315–319.
[222] F. Wilczek, “On Lattice Fermions, Phys. Rev. Lett. 59 (1987) 2397.
[223] S. Capitani, M. Creutz, J. Weber, and H. Wittig, “Renormalization of
minimally doubled fermions, JHEP 09 (2010) 027, arXiv:1006.2009
[hep-lat].
[224] J. H. Weber, S. Capitani, and H. Wittig, “Numerical studies of Minimally
Doubled Fermions, PoS LATTICE2013 (2014) 122, arXiv:1312.0488
[hep-lat].
[225]
A. Borici and P. de Forcrand, “Systematic errors of Lüscher’s fermion method
and its extensions, Nucl. Phys. B454 (1995) 645–662,
arXiv:hep-lat/9505021 [hep-lat].
143
Bibliography
[226] M. A. Clark and A. D. Kennedy, “Accelerating dynamical fermion
computations using the rational hybrid Monte Carlo (RHMC) algorithm with
multiple pseudofermion fields, Phys. Rev. Lett. 98 (2007) 051601,
arXiv:hep-lat/0608015 [hep-lat].
[227] M. A. Clark, “The Rational Hybrid Monte Carlo Algorithm, PoS LAT2006
(2006) 004, arXiv:hep-lat/0610048 [hep-lat].
144