You are on page 1of 112

Search for New Physics with Atoms and Molecules

M.S. Safronova1,2 , D. Budker3,4,5 , D. DeMille6 , Derek F. Jackson Kimball7 , A. Derevianko8 and C. W. Clark2
1
University of Delaware, Newark, Delaware, USA,
2
Joint Quantum Institute, National Institute of Standards and Technology and the University of Maryland,
College Park, Maryland, USA,
3
Helmholtz Institute, Johannes Gutenberg University, Mainz, Germany,
4
University of California, Berkeley, California, USA,
5
Nuclear Science Division, Lawrence Berkeley National Laboratory, Berkeley, California, USA
6
Yale University, New Haven, Connecticut, USA,
7
California State University, East Bay, Hayward, California, USA,
arXiv:1710.01833v1 [physics.atom-ph] 5 Oct 2017

8
University of Nevada, Reno, Nevada, USA

This article reviews recent developments in tests of fundamental physics using atoms
and molecules, including the subjects of parity violation, searches for permanent electric
dipole moments, tests of the CP T theorem and Lorentz symmetry, searches for spa-
tiotemporal variation of fundamental constants, tests of quantum electrodynamics, tests
of general relativity and the equivalence principle, searches for dark matter, dark energy
and extra forces, and tests of the spin-statistics theorem. Key results are presented in
the context of potential new physics and in the broader context of similar investigations
in other fields. Ongoing and future experiments of the next decade are discussed.

CONTENTS III. Precision tests of Quantum Electrodynamics 19


A. Introduction 19
I. Introduction 2 B. Anomalous magnetic moment of the electron 19
A. Recent advances in AMO physics 2 C. Quantum electrodynamics tests with polyelectrons 20
B. Problems with the Standard Model 3 1. Positronium 20
C. Search for new physics with precision measurements 3 2. Positronium anion, Ps− 21
D. Scope of this review 4 3. Diatomic positronium, Ps2 21
D. Proton radius puzzle 21
II. Search for variation of fundamental constants 4 E. Tests of QED in highly charged ions 21
A. Fundamental constants: an introduction 4 1. Energies 21
B. Units of measurement vs. fundamental constants 5 2. Hyperfine splittings 22
C. Theories with varying fundamental constants 6 3. QED tests for g factors 22
D. Tests of fundamental constant variations with atomic F. Conclusion 23
clocks 6
1. Dependence of hyperfine and electronic IV. Atomic parity violation 23
transitions on dimensionless constants 7 A. Introduction 23
2. Theoretical determination of the sensitivity of B. Nuclear-spin independent effects 24
atomic transitions to variations of α 7 1. Overview 24
3. Microwave vs. optical clock-comparison 2. Parity violation in cesium 25
experiments 8 3. Implications for particle physics and the dark
E. Current limits on α and µ variations from atomic sector 26
clocks and Dy spectroscopy 8 4. Isotopic chains and neutron skin 27
F. Prospects for the improvement of atomic clock C. Nuclear-spin-dependent effects and the nuclear
constraints on fundamental-constant variations 10 anapole moment 28
1. Improvements of current clocks 10 1. Overview 28
2. Prospects for optical clocks with highly charged 2. Nuclear anapole moments as a probe of hadronic
ions 11 parity violation 29
3. A candidate nuclear clock 12 D. New and ongoing APV experiments 30
G. Laboratory searches for variation of fundamental
constants with molecules 12 V. Time-reversal violation: electric dipole moments and
H. Limits on variation of α and µ from quasar related phenomena 31
absorption spectra 13 A. Introduction 31
1. Limits on variation of α from quasar absorption B. Observable effects in atoms and molecules 33
studies of atomic spectra 14 C. Underlying physical mechanisms for T,PV 33
2. Limits on variation of µ from quasar absorption 1. Semileptonic interactions 33
studies of molecular spectra 15 2. EDMs of constituent particles: Schiff’s theorem 34
I. Spatial variation of fundamental constants 16 3. Electron EDM 34
1. Search for coupling of fundamental constants to a 4. Hadronic T,PV: nuclear Schiff moment and
changing gravitational potential 17 related effects 35
2. Search for chameleons: testing the dependence of D. State-of-the-art experiments 36
fundamental constants on the mass density of the 1. General remarks 36
environment 18 2. Experiments on paramagnetic systems 36
2

3. Experiments on diamagnetic systems 37 6. Magnetometer and clock networks for detection of


4. Role of low-energy theory 38 transient dark matter signals 69
E. Impact on particle physics 38
F. Future directions 39 X. General relativity and gravitation 70
1. Paramagnetic systems 39 A. Tests of the Einstein equivalence principle 70
2. Diamagnetic systems 40 B. Determination of the Newtonian gravitational
constant 73
VI. Tests of the CP T theorem, matter-antimatter C. Detection of gravitational waves 73
comparisons 41 D. Gravity experiments with antimatter 75
E. Other AMO tests of gravity 75
VII. Review of laboratory searches for exotic spin-dependent
interactions 42 XI. Lorentz symmetry tests 76
A. Early work 42 A. Electron sector of the SME 77
1. Torsion in gravity 43
1. LLI tests with dysprosium 78
2. Axions and axion-like-particles (ALPs) 43
2. LLI test with calcium ion 78
3. Early experiments 44
3. Future prospects and other experiments 79
B. Theoretical motivation 44
B. Proton and neutron sectors of the SME 79
1. Axion-like-particles (ALPs) in string theory 44
2. The hierarchy problem 44 1. Cs clock experiment 79
3. Dark energy 45 2. Comagnetometer experiments 80
4. Unparticles 45 C. Quartz oscillators 81
5. Paraphotons, dark/hidden photons, and new Z 0 D. Photon sector of the SME 81
bosons 45
6. Conclusions 46 XII. Search for violations of quantum statistics, spin-statistics
C. Parametrization 46 theorem 81
1. Introduction 46
2. Moody-Wilczek-Dobrescu-Mocioiu (MWDM) XIII. Conclusion 83
formalism 46
3. MWDM formalism for Lorentz-invariant, Acknowledgements 83
single-boson exchange 47
4. Contact interactions 48 A. Notations, units, and abbreviations 83
5. Position representation and permutation 1. Atomic and molecular properties as encoded in
symmetry 48 spectroscopic notation 83
6. Quantum field theory details 49 2. Atomic symmetries 84
7. Connection between the MWDM formalism and 3. Molecular symmetries 85
various fundamental theories 49 4. Units 86
8. Relationship between coupling constants for 5. Symbols and abbreviations 86
atoms and elementary particles 50
9. Conclusions 51 References 86
D. Overview of experimental searches 51
E. Experimental constraints on monopole-dipole
interactions 52
1. Neutrons 53 I. INTRODUCTION
2. Electrons 54
3. Protons 55 A. Recent advances in AMO physics
4. Astrophysical constraints 56
F. Experimental constraints on dipole-dipole
The past two decades have been a transformational
interactions 56
1. Constraints on V3 (r) 56 era for atomic, molecular and optical (AMO) physics,
2. Constraints on V2 (r) 58 due to extraordinary accomplishments in the control of
3. Astrophysical constraints 59 matter and light. Experimental breakthroughs, includ-
G. Experimental constraints on other forms of ing laser cooling and trapping of atoms, attainment of
spin-dependent interactions 59
H. Emerging ideas 60 Bose-Einstein condensation, optical frequency combs and
quantum control - subject of Nobel Prizes in Physics
VIII. Searches for exotic spin-independent interactions 60 in 1997 (Chu, 1998; Cohen-Tannoudji, 1998; Phillips,
A. Introduction 60
1998), 2001 (Cornell and Wieman, 2002; Ketterle, 2002),
B. Motivation and Theoretical Landscape 61
C. Laboratory tests 62 2005 (Glauber, 2006; Hall, 2006; Hänsch, 2006), and
2012 (Haroche, 2013; Wineland, 2013), respectively -
IX. Searches for light dark matter 63 have led to widespread availability of ultracold (temper-
A. Introduction 63
B. Experimental Searches 66
ature T < 1 µK) ions, atoms and molecules, subject to
1. Microwave cavity axion experiments 66 precise interrogation and control. Revolutionary devel-
2. Spin-precession axion experiments 67 opments on several fronts have been made possible by
3. Radio axion searches 68 these advances, aided by improvements in precision time
4. Atomic clocks and accelerometers, and
spectroscopy 69
and frequency metrology, measurement techniques such
5. Exotic spin-dependent forces due to as atomic magnetometry and interferometry, and first-
axions/ALPs 69 principles atomic and molecular theory. These advances
3

brought forth a plethora of new AMO applications, in- C. Search for new physics with precision measurements
cluding novel tests of the fundamental laws of physics.
While one can search for new particles directly with
large-scale collider experiments at the TeV energy scale,
such as those carried out at the Large Hadron Collider
(LHC) at CERN, new physics may also be observed via
B. Problems with the Standard Model low-energy precision measurements. An early example of
the use of AMO physics in this paradigm, beginning in
the 1970s, was the deployment of highly sensitive laser-
The Standard Model (SM) of particle physics (Olive based techniques to observe parity violation in optical
et al., 2015) has been exceptionally successful in predict- transitions in atoms. This parity violation occurs due
ing and describing numerous phenomena and has been to exchange of Z bosons between electrons and nuclei in
extensively tested by a multitude of different approaches atoms, and quantitative measurements of the strength of
spanning most fields of physics. Despite its great suc- the effect can be used to test the predictions of theoret-
cess, the SM has major problems. Indeed, it is inconsis- ical models of the electroweak interaction (Khriplovich,
tent with the very existence of our Universe: the Stan- 1991). These investigations quickly led to the realiza-
dard Model cannot account for the observed imbalance tion that the accuracy of first-principles theory of atomic
of matter and antimatter (Dine and Kusenko, 2003). In structure needed radical improvement in order to inter-
addition, all attempts to combine gravity with the fun- pret the experimental results. This was particularly true
damental interactions described by the SM have been for heavy atoms like cesium (Cs, Z = 55), which re-
unsuccessful. quired the development of novel theoretical methodolo-
A long-standing mystery dating back to the 1930s gies enabled by modern computing architecture (Porsev
(Zwicky, 1933, 2009) is the apparent existence of “dark et al., 2009). Improved computational resources and de-
matter” that is observed only via its gravitational inter- velopment of high-precision methodologies have led to es-
actions. This is confirmed by numerous studies of astro- sential progress in related theoretical investigations, en-
nomical objects, which show that the particles of the SM abling improved analyses of precision experiments, de-
make up only ≈16% of the total matter present in our velopment of new experimental proposals, and improved
Universe. Decades of investigation have not identified theoretical predictions for yet unmeasured quantities. As
the nature of dark matter (Bertone, 2013). We do now a result, analyses of the Cs atomic parity violation (APV)
know what it is not—any of the particles of the SM. experiment (Wood et al., 1997) provided the most accu-
rate to-date tests of the low-energy electroweak sector
Studies of the Type I supernovae which were originally of the SM and constraints on a variety of scenarios for
aimed at measuring the deceleration rate of the Universe physics beyond the SM (Dzuba et al., 2012a; Porsev et al.,
arrived at a completely unexpected result: the expansion 2009). Combined with the results of high-energy collider
of the Universe is now accelerating (Perlmutter, 2012; experiments, Cs APV studies confirmed the energy de-
Riess, 2012; Schmidt, 2012). This seems to be possible pendence of the electroweak force over an energy range
only if our Universe contains a kind of “dark energy” spanning four orders of magnitude (Porsev et al., 2009).
which effectively acts as repulsive gravity. While we do Further details are given in Sec. IV.A.
not know what dark matter is, we know even less what
Similarly, for several decades AMO experiments have
such dark energy could be—while vacuum energy is a
been employed to search for violation of time-reversal (T )
handy potential candidate, the discrepancy between the
symmetry, as manifested by an electric dipole moment
sum of known contributions to vacuum energy in the uni-
(EDM) along the angular momentum axis of a quantized
verse and the cosmologically observed value is 55 orders
system. T -violation is required to generate a cosmolog-
of magnitude (Solà, 2013). According to the 2015 re-
ical matter-antimatter asymmetry, and sources beyond
sults of the Planck Mission study of cosmic microwave
those in the SM are required to explain the magnitude of
background radiation (Adam et al., 2016), our present
the observed imbalance (Dine and Kusenko, 2003). Ex-
Universe is 69% dark energy, 26% dark matter, and 5%
tensions to the SM frequently introduce new sources of
ordinary (Standard Model) matter.
T -violation that are associated with new particles (Barr,
In summary, we are at an extraordinary point in time 1993). In theories where these new particles have mass at
for physics discovery. We have found all of the particles the TeV scale—or, sometimes, well above it—EDMs are
of the SM and have tested it extensively, but we do not typically predicted with size near the limits set by current
know what makes 95% of the Universe, nor how ordinary AMO experiments (Engel et al., 2013; Pospelov and Ritz,
matter survived against annihilation with antimatter in 2005). Hence these EDM experiments probe commonly-
the aftermath of the Big Bang. This provides strong mo- predicted physics at similar or higher energy scales than
tivation to search for new particles (and/or the associated those accessible with the LHC. New experiments based
fields) beyond those described in the SM. on large enhancements of the observable EDM effects (in
4

experiments using polar molecules or deformed nuclei) A demonstration of the potential of quantum-
hold the promise to increase the energy reach for prob- information techniques in the search for physics beyond
ing new T -violating physics by an order of magnitude or the SM was provided by Pruttivarasin et al. (2015). Us-
even more in the near future. Further details are given ing a pair of trapped calcium (Ca, Z = 20) ions in a
in Sec. V. decoherence-free subspace, they improved by a factor of
AMO experiments also probe even higher energy 100 the bounds on a number of Lorentz-symmetry violat-
scales. A number of theories aiming to unify grav- ing parameters of the Standard Model Extension (SME)
ity with other fundamental interactions suggest viola- for electrons.
tions of cornerstones of modern physics such as Lorentz
symmetry and combined charge-conjugation (C), parity
(P ), and time reversal (CP T ) invariance (Colladay and D. Scope of this review
Kostelecký, 1998; Kostelecký and Russell, 2011) and im-
ply spatiotemporal variation of fundamental constants The examples above show the diversity of recent AMO
(Uzan, 2011). Whereas the energy scale of such physics searches for new physics. Here, we review this subject
is much higher than that attainable at present by par- as a whole rather than limit the treatment to a few spe-
ticle accelerators, Lorentz violation may nevertheless be cific topics, since this field is based on a commonality of
detectable via precision measurements at low energies. approaches that is likely to have even wider applicability
The unprecedented accuracy of AMO precision mea- in the future, given the growth that we have witnessed
surements coupled with accurate theory predictions facil- recently.
itated significant expansion of AMO fundamental physics Another active area of AMO physics is the simula-
studies. As a result, AMO physics now addresses ques- tion of condensed-matter systems using ultracold atoms
tions in fields from which it was once quite remote, such in optical potentials. This field has aspects of searches
as nuclear, particle and gravitational physics and cosmol- for new physics associated with novel quantum phases,
ogy. non-Abelian gauge potentials, atomtronics and the like.
For example, a number of AMO technologies such Our review will not deal with such topics since they
as high-precision magnetometery (Budker et al., 2014; are already addressed by other reviews (Bloch et al.,
Pustelny et al., 2013), atom interferometry (Hamilton 2012, 2008; Georgescu et al., 2014; Goldman et al., 2014;
et al., 2015), atomic clocks (Derevianko and Pospelov, Lewenstein et al., 2007; Stamper-Kurn and Ueda, 2013;
2014), and ultra high-intensity lasers (Di Piazza et al., Ueda, 2014; Windpassinger and Sengstock, 2013), and
2012) are aimed at the search for axions and other dark constitute a vast subject in their own right. We will
matter and dark energy candidates. The principles of a also exclude detailed consideration of quantum mechan-
new technique for detecting transient signals of exotic ori- ics tests with AMO systems which have been recently
gin using a global network of synchronized optical mag- reviewed as well (Aspelmeyer et al., 2014; Bassi et al.,
netometers were demonstrated by Pustelny et al. (2013). 2013; Hornberger et al., 2012).
The network may probe stable topological defects (e.g., Since the field of AMO tests of fundamental physics
domain walls) of axion-like fields. is a vast subject spanning decades of research, we limit
A recent Cs matter-wave interferometry experiment this review to recent developments and proposals. For
constrained a wide class of dynamical dark energy theo- each topic, we begin with an introduction to its specific
ries (Hamilton et al., 2015). The exceptional sensitivity relevance to physics beyond the SM. We present recent
of matter-wave interferometers operated with quantum key results in the context of potential new physics and
gases has generated new ideas for probing the funda- summarize ongoing and future experiments of the next
mental concepts of quantum mechanics, tests of general decade.
relativity, and gravitational wave detection (Biedermann
et al., 2015; Hogan and Kasevich, 2016; Müntinga et al., II. SEARCH FOR VARIATION OF FUNDAMENTAL
2013). The first quantum test of the universality of free CONSTANTS
fall with matter waves of two different atomic species was
reported by Schlippert et al. (2014). A. Fundamental constants: an introduction
The accuracy of atomic clocks has improved by a fac-
tor of 1000 in the past 10 years, to a fractional frequency First, we have to define what we mean by “fundamen-
uncertainty of two parts in 1018 (Nicholson et al., 2015) tal constants.” Opening a physics textbook on various
which corresponds to a temporal uncertainty of one sec- physics fields would produce various lists of measured
ond in the lifetime of our Universe. As a result, atomic quantities of specific importance to a given field. In this
clocks are now used to search for possible time varia- review, we follow the definition of Uzan (2015): a fun-
tions of the dimensionless fine-structure constant, α, and damental constant is any parameter not determined by
proton-electron mass ratio, mp /me (Godun et al., 2014; the theories in which it appears. This definition has the
Huntemann et al., 2014; Rosenband et al., 2008). following implications:
5

• the number of fundamental constants depends on a atomic, molecular, and optical (AMO) physics, is the de-
particular theory and termination of the fine-structure constant
1 e2
• the fundamental constants are not predicted by any α= , (1)
theory and thus their values must be determined 4π0 ~c
through measurements. which characterizes the strength of the electromagnetic
interaction, see Sec. III. Here, e is the elementary charge,
Present physics is described by general relativity (GR) ~ = h/2π is the reduced Planck constant, and 0 is the
and the Standard Model (SM) of particle physics that electric constant. A recent overview of the determina-
combines quantum chromodynamics (QCD) with the tions of fundamental constants from low-energy measure-
electroweak theory (Olive et al., 2015). The minimal ments is given by Karshenboim (2013).
SM has 19 parameters, with somewhat different sets of We note the values of the coupling constants of the
these parameters given in the literature (Hogan, 2000; SM depend on the energy at which they are measured
Scott, 2006; Uzan, 2013). Following the summary of (so-called “running” of the coupling constants discussed
Scott (2006), the list contains 6 quark masses, 3 lepton using the example of sin2 θW in Sec. IV.A). The fine-
masses, 3 quark mixing angles (θ12 , θ23 , θ13 ) and phase structure constant α is defined in the limit of zero mo-
δ, 3 electroweak parameters (fine-structure constant α, mentum transfer.
Fermi coupling constant GF , and mass of the Z boson
MZ ), Higgs mass, strong combined charge-conjugation
and parity (CP ) violating phase and the QCD coupling B. Units of measurement vs. fundamental constants
constant. The incorporation of the neutrino masses leads
to additional parameters. Experimental measurements can be reduced to com-
To reproduce known physics, the SM parameters must paring two physical systems, one of which defines the unit
be supplemented by the Newtonian constant of gravita- of measurement. For example, the International System
tion G of GR, the speed of light in vacuum c, and the of Units (SI) unit of time is defined as: “The second
Planck constant h. We note that this list of fundamen- is the duration of 9 192 631 770 periods of the radiation
tal constants lacks any description of either dark matter corresponding to the transition between the two hyper-
or dark energy and contains no cosmological information fine levels of the ground state of the cesium 133 atom”
about the Universe. The Standard Cosmological Model (BIPM, 2014). This definition refers to a Cs atom at rest
adds 12 more parameters, listed by Scott (2006) which in- at a temperature of 0 K. Therefore, absolute values of all
clude the Hubble constant, baryon, cold dark matter and other frequencies are determined relative to this Cs fre-
dark energy densities, and others. Further understanding quency and no absolute frequency measurement can be
of these phenomena may increase the required number of performed with smaller fractional frequency uncertainty
fundamental constants, while developing a unified theory than that of the best Cs frequency standard, which is
might reduce them. presently on the order of 10−16 (Heavner et al., 2014;
Measurements of fundamental constants and numerous Levi et al., 2014). Note that one can still make relative
other derived quantities, some of which can be predicted comparison of two frequencies to much better precision
from current theories with varying levels of accuracy, is a than the Cs standard provides (Ushijima et al., 2015). To
vast area of research. We refer the reader to the publica- make absolute frequency measurements accurate to, for
tions of Committee on Data for Science and Technology example, 10−18 of a second, we would need to change the
(CODATA), (Mohr et al., 2016) and Particle Data Group definition of the second from the Cs microwave frequency
(Olive et al., 2015) for measurement techniques, analysis transition to another physical system. Such system must
of data and current recommended values. The data are allow for the construction of the frequency standard with
continuously revised and improved, with critical assess- 10−18 uncertainty in a consistently reproducible way, ac-
ment of various types of experiments carried out prior to companied by a global technology infrastructure for fre-
new CODATA and PDG publications. quency comparison (Ludlow et al., 2015).
It should be kept in mind that there is no single experi- Changing the values of the constants in such a way
ment that determines the CODATA recommended value that all dimensionless combinations are unchanged will
of a given fundamental constant. There is a complex simply change the units. For example, in atomic units the
web of deep and sometimes subtle connections between values of e, the electron mass me , and the reduced Planck
fundamental constants - for example, between the fine constant ~ have the numerical value 1, and the electric
structure constant and the molar Planck constant NA h constant 0 has the numerical value 1/(4π). However,
(Mohr et al., 2016) - and the CODATA recommended the value of the dimensionless fine-structure constant α
values are determined by a least-squares adjustment that is still the same as in SI units as given by Eq. (26).
keeps inconsistencies within limits. Dimensionless fundamental constants play a special
An example of this interdependence, that is highlighted role in discussions of spatiotemporal variations of physi-
by Mohr et al. (2016) and is of particular relevance to cal laws. Their values are, by construction, independent
6

of the choice of units of measurement, which are arbi- chameleon field is expected to be more massive in high-
trary conventions that have changed in the past and may density regions on Earth than in low-density regions of
change in the future. the solar system (Khoury and Weltman, 2004a). Since
For example, it is difficult to see how one could measure the constants would be dependent on the local value of
unambiguously a time variation in the speed of light, c. the chameleon field, the values of the constants become
This may be viewed from the perspective of 1982. Then, dependent on their (mass density) environment.
the second and the meter were defined independently: While one can construct models in which only one or
the second as it is today, a defined multiple of the period a few constants vary, in most realistic current models, all
of the ground hyperfine transition of 133 Cs, and the meter constants vary if one does (Uzan, 2015). In unified the-
as a defined multiple of the wavelength of the 2p10 −5d5 ories of fundamental interactions the variations of fun-
spectral line of an isotope of krypton (Kr, Z = 36), 86 Kr. damental constants are correlated. However, including
On a simple observational basis, if a change in the 1982 such correlations in the analysis of experiments leads to
value of c was well established on the basis of multiple dependence of the results on the particular model.
independent observations, it seems impossible to disen- It has been pointed out that searching for variation
tangle that effect from changes in either, or both, of the of fundamental constants is a test of the local position
Cs frequency or the Kr wavelength. invariance hypothesis and thus of the equivalence prin-
The focus of modern studies of variation of fundamen- ciple [see Uzan (2011) and Uzan (2015) and references
tal constants is thus on dimensionless constants, and as therein].
concerns AMO physics, particularly on α and the pro- Searches for variation of fundamental constants are
ton/electron mass ratio, mp /me . conducted in a number of systems including atomic
clocks, astrophysical studies of quasar spectra and ob-
servation of the H I 21 cm line, the Oklo natural nu-
C. Theories with varying fundamental constants clear reactor, meteorite dating, stellar physics, cosmic
microwave background (CMB), and big bang nucleosyn-
While the 2014 CODATA value of fine-structure con- thesis (BBN). A detailed review of these topics is given
stant α has a remarkably small 2.3 × 10−10 uncertainty, by Uzan (2011). We limit our coverage to recent results,
it remains an open questions whether the value of α is ongoing experiments, and proposals relevant to AMO
variable across space and time. In the SM, all fundamen- physics.
tal constants are invariable. The dimensionless constants Laboratory tests for the variation of fundamental con-
become dynamical (i.e. varying) in a number of theories stants, such as carried out with atomic clocks, are only
beyond the SM and GR. Detailed review and references sensitive to present-day variation, while other searches
to theories with varying fundamental constants are given are probing whether α and other constants were different
by Uzan (2011) so we only give a brief summary here. in the past compared to what they are now, with differ-
Higher-dimensional theories, in particular string the- ent look-back times. We discuss this further in Sec. II.H.
ories, naturally lead to varying fundamental constants. The analysis of CMB and BBN in terms of constraining
String theories predict the existence of a scalar field, the variation of fundamental consistent is also dependent
the dilaton, that couples directly to matter (Taylor and on the cosmological model.
Veneziano, 1988). The 4-dimensional coupling constants In this section, we consider the “slow-drift” model of
are determined in terms of a string scale and various dy- variation of fundamental constants, as well as coupling
namical fields. As a result, the coupling constants nat- of fundamental constants to a changing gravitational po-
urally become varying, evaluated as the expectation val- tential, and testing for a dependence of fundamental con-
ues of these dynamical fields. The variation of the gauge stants on the mass density of the environment. Searches
couplings and of the gravitational constant may also arise for oscillatory and transient variation of fundamental
from the variation of the size of the extra dimensions. constants, and their relevance to the nature of dark mat-
Many other theories beyond the SM and GR have ter and dark energy, are discussed in Sec. IX.
been proposed in which fundamental constants become
dynamic fields. These include: discrete quantum grav-
ity (Gambini and Pullin, 2003); loop quantum gravity D. Tests of fundamental constant variations with atomic
(Taveras and Yunes, 2008); chameleon models (Khoury clocks
and Weltman, 2004a); dark energy models with a non-
minimal coupling of a quintessence field (Avelino et al., The most precise tests of modern-epoch variation
2006) and others. As a result, studies of the variation of of fundamental constants are carried out using atomic
fundamental constant may provide some information on clocks. From the standpoint of metrology and other pre-
potential origin of dark energy. Analysis of experiments cision experiments, testing the variation of fundamental
on the variation of fundamental constants also depends constants is necessary to ensure that the experiments are
on the nature of the particular model. For example, a reproducible at the level of their uncertainties. This be-
7

came particularly important due to exceptional improve- I = 7/2 and the ground state electronic configuration
ment of AMO precision metrology in recent years. If α or consists of a closed Xe-like core (xenon, Z = 54) with
µ = mp /me are space-time dependent, so are atomic and an unpaired single valence electron with j = 1/2. There-
molecular spectra. Therefore, the variation of the fun- fore, Cs [Xe]6s ground state splits into two hyperfine lev-
damental constants makes the clock tick rate dependent els, with F = 3 and F = 4, where the total angular
on location, time, or type of the clock - since frequencies momentum F = I + J. The frequency of electromagnetic
of Cs or Sr depend differently on fundamental constants. radiation associated with transitions between these levels
We have arrived at a level of precision such that new is conventionally expressed as
physics might show up unexpectedly as an irreducible
systematic error! An important question for AMO the- me
ory is predicting the best systems for dedicated exper- νhfs ' cR∞ Ahfs × gi × × α2 Fhfs (α), (4)
mp
iments where the variation of fundamental constants is
strongly enhanced. where R∞ is given by Eq. (3), Ahfs is the numerical quan-
We start with the discussion of the dependence of tity depending on a particular atom, and Fhfs (α) is a
atomic spectra on the dimensionless constants of interest. relativistic correction specific to each hyperfine transi-
The possibility of using atomic spectroscopy to detect tion. The dimensionless gi = µi /µN is the g-factor as-
variations in the fine-structure constant α is suggested sociated with the nuclear magnetic moment µi , where
in Dirac’s theory of the hydrogen atom. The energies of µN = e~/2mp is the nuclear magneton.
En,j of a Dirac electron bound to an infinite-mass point The Cs hyperfine F = 3 − F = 4 transition frequency
nucleus are given by νCs of ≈ 9 GHz defines the second, and all absolute
frequency measurements are actually measurements of
En,j = me c2 ν/νCs frequency ratios. Atomic clocks based on hyperfine
 −1/2 transitions are referred to in the literature as “microwave
clocks”, specifying the relevant region of the electromag-
2
 (Zα) 
netic spectrum (Ludlow et al., 2015).
× 1 +  , (2)
 
2
The transition frequency between electronic energy
q 
 1 2 2

j + 12 − (Zα)

n−j− 2 + levels in an atom can be expressed as

where Ze is the charge of the nucleus, with e the elemen- ν ' cR∞ AF (α), (5)
tary charge, n is the principal quantum number, and j
where A is the numerical factor depending on an
the electronic angular momentum in units of ~ (Greiner,
atom and F (α) depends upon the particular transition.
2000; Johnson, 2007). With reference to the discussion
Atomic clocks based on electronic transitions with fre-
of Sec. II.A, note that the Rydberg constant is given by
quencies in ∼ 0.4 − 1.1 × 1015 Hz region are referred to
in the literature as “optical clocks”.
1 α2
R∞ = · me c2 . (3)
hc 2
2. Theoretical determination of the sensitivity of atomic
The only fundamental constants present in Eq. (2) are transitions to variations of α
α2 and the rest mass energy of the electron, me c2 . Ex-
pansion of En,j in powers of α2 shows that for electronic The coefficient F (α) in Eq. (5) is obtained by calculat-
states with different values of the principal quantum ing the α-dependence of the energies of the two atomic
number n, the energy splitting scales with α2 , whereas levels involved in the transition. The dependence of elec-
the splitting scales with α4 for states with the same n but tronic energy level E on α is usually parameterized by
different j. Thus, ratios of the wavelengths of these two the coefficient q (Dzuba et al., 1999a,b)
types of transitions are sensitive to variations in α. The " 
2
#
dependence of the atomic spectra of more complicated α
E(α) = E0 + q −1 , (6)
atoms on fundamental constants is discussed below. α0

which can be determined rather accurately [generally to


1. Dependence of hyperfine and electronic transitions on 1-10%] from atomic-structure computations. Here, α0 is
dimensionless constants the current value of α (Mohr et al., 2016), the measure-
ment of which was discussed in Sec. II.A, and E0 is the
Interaction of atomic electrons with the magnetic and energy corresponding to this value of α0 . The coefficient
electric multipole fields of the nucleus leads to a splitting q depends weakly on electron correlations, so it can be
of atomic energy levels referred to as hyperfine structure. calculated more accurately than the actual energy of a
For example, the nuclear angular momentum of 133 Cs is level.
8

The coefficient q of an atomic state is computed by


TABLE I Sensitivity factors K to the variation of the fine-
varying the numerical value of α in the computation structure constant α for clock transitions (Dzuba and Flam-
of the respective energy level (Dzuba and Flambaum, baum, 2009a). K is defined by Eq. (9). All transitions except
2009a). Generally, three energy level calculations are Rb and Cs are optical frequency standards.
performed which differ only by the values of α. The first
Atom Transition K
calculation uses the current CODATA value of α2 (Mohr 87
Rb ground hyperfine 0.34
et al., 2016). Two other calculations are performed with 133
Cs ground hyperfine 0.83
α2 varied by a small but non-negligible amount, com- Al+ 3s2 1 S0 − 3s3p 3 P0 0.008
monly selected at δ = 0.01. Then, the value of q is de- Ca+ 4s 2 S1/2 − 3d 2 D5/2 0.15
rived from a numerical derivative Sr 5s2 1 S0 − 5s5p 3 P0 0.06
E(δ) − E(−δ) Sr+ 5s 2 S1/2 − 4d 2 D5/2 0.43
q= , (7) Yb 6s2 1 S0 − 6s6p 3 P0 0.31

Hg+ 6s 2 S1/2 − 5d 2 D5/2 -2.94
where E(±δ) are results of the energy calculations with Yb+ E2 4f 14 6s 2 S1/2 − 4f 14 5d 2 D5/2 1.03
α2 varied by ±1%, respectively. The additional calcula- Yb+ E3 4f 14 6s 2 S1/2 − 4f 13 6s2 2 F7/2 -5.95
tion (with the CODATA value of α) is used to verify that
the change in the energy is close to linear.
The parameter q links variation of the transition en- ν1 /ν2 of two clocks over time may set limits on varia-
ergy E, and hence the atomic frequency ν = E/h, to the tion of α, the proton-to-electron mass ratio µ = mp /me ,
variation of α and nuclear g factors, specifically gCs and gRb as these
δE 2q δα δα correspond to two microwave clocks with the smallest
= ≡K , (8)
E0 E0 α0 α0 uncertainties. We summarize the dependence of clock-
where frequency ratios on the dimensionless constants as fol-
lows:
2q
K= (9)
E0 • The ratio of two microwave clock frequencies de-
is a dimensionless sensitivity factor. pends on α and g-factors of the corresponding nu-
In α-variation tests with atomic clocks, the ratio of clei according to Eq. (4). For example, the ratio
two clock frequencies is monitored, and the sensitivity to of Cs to Rb (rubidium, Z = 37)clock frequencies is
the variation of α is then given by the difference in their proportional to
respective K values for each clock transition, i.e. ∆K = νCs gCs gCs 0.49
∝ × αKCs −KRb = α , (10)
|K2 − K1 |. The larger the value of K, the more sensitive νRb gRb gRb
is a particular atomic energy level to the variation of α.
where the K factors defined by Eq. (9) are given in
A note of caution has to be added here: while small E0
Table I.
may lead to large K following Eq. (9), it may also lead to
technical difficulties in measuring the relevant frequency • The ratio of frequencies of any two optical clocks
with the extremely high accuracy that is required for depends only upon α, according to Eq. (5).
tests of variation of fundamental constants. Small en-
ergy E0 corresponds to transitions in the infrared, with • The ratio of optical to microwave clock frequencies
wavelength that may exceed 3000 nm. Accurate theory depends on α, µ = mp /me ratio, and the g-factor
predictions are particulary difficult for such transitions, of the atomic nucleus of the microwave clock.
as small E0 is the result of strong cancellations of up- Reducing the potential variation of g-factors to more fun-
per and lower energies, leading to difficulties in locat- damental quantities, such as Xq = mq /ΛQCD , and cal-
ing weak clock transitions. Moreover, the clock instabil- culation of the corresponding dimensionless sensitivity
ity is inversely proportional to the transition frequency factors κCs and κRb , requires nuclear structure calcula-
so lower frequency leads to higher ultimate instability, tions which are dependent on a particular model (Dinh
which is particulary problematic with single ion clocks. et al., 2009; Flambaum and Tedesco, 2006; Jackson Kim-
Therefore, the actual transition frequency and other ex- ball, 2015). Here, mq is the average light-quark mass and
perimental considerations have to be taken into account ΛQCD is the QCD energy scale.
when designing dedicated experiments. This issue is fur-
ther discussed in Sec. II.F.
E. Current limits on α and µ variations from atomic clocks
and Dy spectroscopy
3. Microwave vs. optical clock-comparison experiments
At present, the best constraints on temporal variations
At the lowest level of the analysis that requires only of α and µ from comparisons of atomic transition fre-
atomic structure calculations, measuring the ratios R = quencies are due to combination of several experiments
9

optical clocks (Rosenband et al., 2008) and from the


measurement of Dy transition frequencies (Leefer et al.,
2013). The Al+ /Hg+ optical clock comparison (Rosen-
band et al., 2008) currently provides the most accurate
single test of only α-variation, setting the limit

α̇
= (−1.6 ± 2.3) × 10−17 yr−1 . (13)
α

The Dy limit on α-variation comes from spectroscopy


of radio-frequency transitions between nearly degenerate,
opposite-parity excited states rather than from an atomic
clock comparison. These states are sensitive to variation
FIG. 1 (color online). Constraints on temporal variations of of α due to large relativistic corrections of opposite sign
α and µ from comparisons of atomic transition frequencies. for the opposite-parity levels. The near degeneracy re-
Filled stripes mark the one standard deviation σ uncertainty
duces the relative precision needed to place strict con-
regions of individual measurements and the central blank re-
gion is bounded by the standard uncertainty ellipse resulting straints on α-variation. We note that filled stripes rep-
from the combination of all data. From Huntemann et al. resenting results of both Al+ /Hg+ and dysprosium (Dy,
(2014). Z = 66) experiments in Fig. 1 are vertical, since they are
sensitive only to variation of α and not mp /me .
We emphasize that Yb+ (ytterbium, Z = 70) has two
tracking ratios of different clock transitions (Godun et al., ultranarrow optical clock transitions at 467 and 436 nm:
2014; Huntemann et al., 2014). The analysis of current electric octupole (E3) 4f 14 6s 2 S1/2 − 4f 13 6s2 2 F7/2 and
α and µ clock constraints of Huntemann et al. (2014) electric quadrupole 4f 14 6s 2 S1/2 − 4f 14 5d 2 D5/2 . This is
is illustrated in Fig. 1. Filled stripes mark one-standard- the only case among the clocks presently under develop-
deviation uncertainty regions of individual measurements ment for which there is more than one clock transition.
and the central blank region is bounded by the standard
uncertainty ellipse resulting from the combination of all The frequency ratio of those two transitions in Yb+
data. The results of the experiments measuring the sta- was measured directly for the first time by Godun et al.
bility of the frequency ratios R = ν/νCs of optical Hg+ (2014), without reference to the Cs primary standard,
(Fortier et al., 2007), Yb+ quadrupole (E2) (Tamm et al., and using the same single ion of 171 Yb. This measure-
2014), Yb+ octupole (E3) (Huntemann et al., 2014), and ment is illustrated in Fig. 2. The E3/E2 frequency ra-
Sr (Le Targat et al., 2013) clocks to the Cs microwave tio was determined by stabilizing one laser to the E3
clock plotted in Fig. 1 were parameterized by transition and the other laser to the E2 transition and
measuring the ratio between the laser frequencies with
1 dR 1 dα 1 dµ 1 dXq an optical frequency comb. Both lasers were simultane-
= (K − KCs − 2) + − κCs , (11)
R dt α dt µ dt Xq dt ously stabilized to their respective transitions in the same
ion ensuring experimental simplicity and common-mode
where the coefficients K for the optical clocks and Cs rejection of certain systematic effects such as the gravita-
are listed in Table I. We note that the extra “2” in the tional redshift and relativistic time dilation. Such direct
parenthesis of the first term appears due to the presence measurements of the ratio of the two optical frequencies
of a factor of α2 in the hyperfine frequency expression are free from the additional uncertainties introduced by
given by Eq. (4). the primary Cs frequency standard.
The contribution due to the third term in Eq. (11) was Combining this measurement with constraints from
taken to be zero in the analysis of Godun et al. (2014). previous experiments, Godun et al. (2014) set the fol-
Huntemann et al. (2014) accounted for this term by using lowing limits to the present day variation of α and µ:
the result
1 dXq α̇
κCs = 0.14(9) × 10−16 /year (12) = (−0.7 ± 2.1) × 10−17 yr−1 (14)
Xq dt α
µ̇
inferred from the comparison of 87 Rb and 133 Cs clocks = (0.2 ± 1.1) × 10−16 yr−1 , (15)
µ
over 14 years reported by Guéna et al. (2012).
Figure 1 also includes constraints on temporal varia-
tion of α from comparisons of transition frequencies of which are similar to limits set by the analysis of Hunte-
Al+ (aluminum, Z = 13) and Hg+ (mercury, Z = 80) mann et al. (2014).
10

1. Improvements of current clocks


x2 LASER 2
436 nm

Femtosecond optical frequency comb


Figure 3 illustrates the evolution of fractional fre-
2D
3/2 quency uncertainties of atomic frequency standards
2F
7/2 based on microwave and optical transitions. All mi-
171Yb+ E2
crowave data in this figure come from Cs clocks. The
2S
ION 1/2 figure is adapted from Poli et al. (2013) with addition of
recent data up to 2016.
The present-day state-of-the-art Cs microwave clocks
2D
3/2
nE3 are approaching uncertainties of 10−16 (Heavner et al.,
2F
7/2 2014; Levi et al., 2014), which is near their practical
E3 nE𝟐
End-cap ion trap limitations. This is a remarkable achievement consid-
2S
1/2 ering that the Cs atomic clock transition has an intrinsic
467 nm quality factor Q, defined as the ratio of the absolute fre-
x2 LASER 1 quency of the transition to its natural the linewidth, of
Q ∼ 1010 . The Q factors of optical atomic clocks are
FIG. 2 Schematic experimental arrangement for measuring five orders of magnitude higher than those of microwave
the E2 and E3 clock frequencies of a single 171 Yb+ ion. clocks, giving optical clocks a tremendous advantage in
The E3/E2 frequency ratio was determined by stabilizing one terms of frequency stability (Ludlow et al., 2015; Poli
laser to the E3 transition and the other laser to the E2 tran- et al., 2013). Recent progress in the accuracy of the
sition and measuring the ratio between the laser frequencies
optical clocks has been extraordinary, with the world’s
with an optical frequency comb. (For experimental reasons,
the researchers used infrared lasers that have to be frequency best optical lattice atomic clocks approaching fractional
doubled to excite the E2 and E3 optical transitions. Adapted frequency uncertainties of 10−18 (Nicholson et al., 2015;
from Safronova (2014). Ushijima et al., 2015). The smallest uncertainty attained
to date is 2 × 10−18 in a strontium (Sr, Z = 38) optical
lattice clock (Nicholson et al., 2015). In 2016, a system-
F. Prospects for the improvement of atomic clock atic uncertainty of 3 × 10−18 was reported in a single-ion
constraints on fundamental-constant variations atomic clock based on the electric-octupole transition in
Yb+ (Huntemann et al., 2016). As a result, the most
The limits on the variation of the fundamental con- rapid improvement in this field is expected to come from
stants from comparison of two clock frequencies are de- optical to optical clock comparison, with optical to mi-
termined by (1) uncertainties of both clocks, (2) sensi- crowave comparison being limited by the ultimate accu-
tivity factors of each clock to the variation of different racy of microwave clocks.
constants, and (3) the time interval over which the ratios Results of experiments measuring the stability of two
are repeatedly measured. Therefore, strategies to im- optical clock-frequency ratios R = ν2 /ν1 are parame-
prove the limits set by atomic clocks on the variation of terised by a simpler version of Eq. (11):
fundamental constants may arise from the improvement
of any of the three factors: building clocks with lower Ṙ α̇
= (K2 − K1 ) , (16)
uncertainties, building conceptually different clocks with R α
higher sensitivities to variation of fundamental constants, where K1 and K2 are K sensitivity coefficients for clocks
and making measurements over longer time intervals. 1 and 2. Therefore, the sensitivity of optical clock fre-
For example, the Al+ /Hg+ clock constraint on α- quency ratios to α-variation is described by the difference
variation reported in 2008 (Rosenband et al., 2008) was in the corresponding K values, i.e. ∆K = |K2 −K1 |. The
obtained from repeated measurements during one year. K factors are small (0.008 − 1.0, see Table I) for most
Even with the same accuracy for both Al+ and Hg+ clocks currently in development: Mg, Al+ , Ca+ , Sr+ , Sr,
clocks, repeating the frequency-ratio measurements now Yb, Yb+ quadrupole transition, and Hg. The K fac-
would improve the 2008 limit (13) by almost of factor tors for Hg+ and Yb+ octupole clock transitions are −3
of 10, since almost a decade has passed since the first and −6, making them the best candidates for one mem-
measurements. For clock-ratio experiments that have al- ber of a clock-comparison pair, with the other member
ready accumulated more than a decade of data, such as taken from the previous group. Recently reported dras-
the Cs/Rb ratio (Guéna et al., 2012), only moderate im- tic reductions in the fractional frequency uncertainty of
provements can be achieved in the next decade without the Yb+ octupole clock (Huntemann et al., 2016) are ex-
the reduction of clock uncertainties. We start with a pected to lead to a more accurate test of α-variation, with
discussion of the prospects for further improvements in the second clock being, perhaps, Sr. Future prospects
searches for variation of fundamental constants with cur- for development of optical atomic clocks are discussed in
rent clocks and then explore new clock proposals. recent reviews (Ludlow et al., 2015; Poli et al., 2013),
11

clock with accuracy at the present state-of-the-art level,


we need, at the very least, a system with a transition in a
laser accessible range with very high Q, at least Q ∼ 1015 .
The high-Q requirement means that the upper state of
the transition is metastable, i.e. long-lived. There are a
number of other considerations to ensure small frequency
uncertainties by minimizing systematic effects. The sys-
tem also has to be amenable to cooling and trapping. If
we want to use our new clock to search for α-variation,
the clock transition has to be between states of differ-
ent electronic configurations, i.e. not between fine- or
hyperfine-structure levels since the K factors for such
states are similar.
These requirements were formulated in the criteria for
good clock transitions proposed by Dzuba et al. (2015a):
• The transition is in the near-optical region
FIG. 3 Evolution of fractional frequency uncertainties of (230 nm < λ < 2000 nm or 5000 cm−1 < ~ω <
atomic frequency standards based on microwave (Cs clocks) 43000 cm−1 ) as such transitions are accessible with
and optical transitions. Data points are from Huntemann available laser systems.
et al. (2016); Nicholson et al. (2015); Poli et al. (2013).
• The lifetime of the clock state is between 100 and
∼ 104 seconds as this enables high Q.
which envisage further decreases in atomic clock uncer-
tainties during the next decade. Comparison of different • There are other relatively strong optical transitions
clocks frequencies beyond 10−18 accuracy will become in the same atomic system with a lifetime of the
more challenging due to the sensitivity to the environ- upper level on the order of τ <∼ 1 ms, which may be
ment, including temperature and gravitational potential useful for laser cooling or optical pumping/probing.
(Ludlow et al., 2015). For example, a clock on the sur- • The clock transition is not sensitive to perturba-
face of the Earth that is higher by just 1 cm than another tions caused by blackbody radiation, gradients of
identical clock runs faster by δν/ν0 ∼ 10−18 (Ludlow external electric fields, etc.
et al., 2015). The blackbody radiation (BBR) shift has a
leading temperature dependence of T 4 , making clock fre- The first requirement seems to limit the potential sys-
quencies sensitive to the temperature fluctuations. The tems to neutral atoms or singly charged ions, all of which
BBR shift for a given temperature also varies signifi- have been considered as potential clock systems. Exam-
cantly, by orders of magnitude, between different clock ination of the NIST atomic spectra database (Kramida
transitions. The strategies for reducing BBR shifts in et al., 2016) establishes that the energies of the relevant
current clocks are discussed by Ludlow et al. (2015) and ion transitions involving the ground states tend to be
references therein. Comparisons of clocks based on two outside of the laser-accessible range with the degree of
transitions in a single ion, such as Yb+ quadrupole to ionization exceeding two. Remarkably, selected highly
octupole clock comparisons illustrated by Fig. 2 or with charged ions with degrees of ionization ranging from 9 to
two ions held in the same trap, may be used to reduce 18, actually have potential clock transitions in the optical
the environmental sensitivities of the clock ratios. range between different electronic configurations, as was
discovered by Berengut et al. (2010). This phenomenon
arises from the rearrangement of the order of electronic
2. Prospects for optical clocks with highly charged ions configurations: as more electrons are removed, the order
of levels becomes more hydrogenic, for example, restoring
Another pathway toward improved tests of α-variation the Coulomb ordering where the 4f shell becomes occu-
with atomic clocks is the development of frequency stan- pied prior to the 5s shell. For example, the ground state
dards based on new systems, which have higher K sen- of cadmium (Cd, Z = 48) is [Kr]4d10 5s2 . Proceeding
sitivities, while still enabling highly accurate measure- along the Cd isoelectronic sequence, the ground state re-
ments of the frequency ratios. Put simply, it is much mains in this configuration up to Nd12+ , but the ground-
easier to measure large effects, so the search for high- state configuration of Sm14+ (samarium, Z = 62) be-
sensitivity systems is a major ongoing effort of AMO comes [Kr]4d10 4f 2 (Safronova et al., 2014a). As a result,
theory. different electronic configurations move close together for
This brings us to a question: what are the require- two or three ions in an isoelectronic sequence when the
ments for such new systems? If we would like to build a order of levels is rearranged. An example is the Nd12+
12

(neodymium, Z = 60), Pm13+ (promethium, Z = 61), 3. A candidate nuclear clock


Sm14+ (samarium, Z = 62) part of the Cd-like sequence.
This provides an unexpected gift of optical transitions With atoms and ions of the periodic table now con-
for metrology applications. Extensive theoretical effort sidered, we turn our attention to nuclei. Can we build
during the past five years resulted in the identification clocks based on transitions between different states of a
of many such candidate systems in highly charged ions nucleus? A great attraction of such an idea is the sup-
(HCIs), predictions of their properties, and assessments pression of the field-induced frequency shifts since the
of their potential for tests of α-variation (Berengut et al., nucleus is highly isolated from the environment due to
2011a,c, 2012a,b; Derevianko et al., 2012; Dzuba et al., the electron cloud and interacts only via the relatively
2012b,b, 2013; Dzuba et al., 2015a; Kozlov et al., 2013; small nuclear moments (Yamaguchi et al., 2015). There
Yudin et al., 2014). The most accurate calculations were is a vast catalog of nuclear energy levels (Firestone and
done using a state-of-the-art hybrid approach that com- Shirley, 1998), but their transition frequencies are higher
bines coupled-cluster and configuration interaction meth- by factors of 104 − 106 than those accessible by modern
ods (Dzuba et al., 2015b; Safronova et al., 2014a,b,c). laser technologies. Only one sufficiently long-lived nu-
Proposals for α-variation searches in HCIs were reviewed clear transition, between the ground state of 229 Th (tho-
by Berengut et al. (2013a) and Ong et al. (2014). rium, Z = 90) and a low-lying isomer (i.e. long-lived ex-
cited nuclear state), has a suitable wavelength, predicted
to be 160(10) nm (Beck et al., 2007; Beck et al., 2009).
A particular attraction of the HCIs for constructing
This transition was proposed for application in a nuclear
highly accurate clocks is the suppression of the clock-
clock (Campbell et al., 2012; Peik and Tamm, 2003), but
frequency shifts due to external electric fields which
a decade of searches did not result in its detection (Jeet
can lead to systematic errors, due to the contraction
et al., 2015; Peik and Okhapkin, 2015; Yamaguchi et al.,
of the electron cloud with increasing ionization stage.
2015). Finally, in 2016, the existence of the isomer was
Stronger relativistic effects resulting from localization of
confirmed (von der Wense et al., 2016), although there
the electron cloud also provide enhanced sensitivity to
remains a significant uncertainty in its energy, motivating
α-variation. Assessments of systematic effects in optical
continued searches.
clocks based on HCIs concluded that an uncertainty of
Flambaum (2006) estimated that the relative effects of
10−19 is achievable (Derevianko et al., 2012; Dzuba et al.,
the variation of α and mq /ΛQCD in this 229 Th nuclear
2012b, 2013; Dzuba et al., 2015a).
transition are enhanced by 5 − 6 orders of magnitude
using the Walecka model of nuclear forces and other as-
Up to this point we have not discussed the techni- sumptions. Other nuclear calculations predicted no en-
cal feasibility of using HCIs to build clocks. Until very hancement (Hayes et al., 2008), but their uncertainly was
recently, the realm of HCI research had little overlap also very large, with a 4 × 103 enhancement factor still
with field of ultracold precision metrology. In 2015, a being within their uncertainty limit, [see Berengut et al.
breakthrough experiment (Schmöger et al., 2015) demon- (2009) for a discussion]. With nuclear calculations cur-
strated sympathetic cooling of Ar13+ (argon, Z = 18) rently being unable to determine the sensitivity factor, an
ions with laser-cooled Be+ (beryllium, Z = 4) ions in a alternative method for extracting sensitivity to α varia-
cryogenic Paul trap. This result removes a major obsta- tion using laboratory measurements of the change in nu-
cle for HCI investigations with high-precision laser spec- clear mean-square charge radius and electric-quadrupole
troscopy, paving the way toward future experiments with moment between the isomer and the ground-state nucleus
cold and precisely controlled HCIs. Experimental work was proposed by Berengut et al. (2009).
toward this goal is underway, starting with the identifica-
tion of the HCI spectra of interest to α-variation studies
(Windberger et al., 2015). Optical transitions in HCIs G. Laboratory searches for variation of fundamental
and their applications will be reviewed in detail in a sep- constants with molecules
arate Reviews of Modern Physics article.
Molecular spectroscopy provides further possibilities
for testing the stability of fundamental constants owing
Hyperfine transitions of hydrogen-like HCIs, such as to rich spectra with many different transition types. The
207
Pb81+ (lead, Z = 82), have also been proposed for proton-to-electron mass ratio µ defines the scales of elec-
tests of fundamental constant variation (Schiller, 2007). tronic, vibrational, and rotational intervals in molecular
Due to high degree of ionization, the ground-state hy- spectra:
perfine transition wavelength is in the infrared, with a
Q-factor of about 1014 . The importance of such HCI Eel : Evib : Erot ∼ 1 : µ−1/2 : µ−1 . (17)
transitions is their sensitivity to the variation of µ and
g-factors, and Q factors that are much larger than those Purely vibrational and rotational transitions will have
of Cs and Rb hyperfine transitions. the Kµ = −1/2 and Kµ = −1 sensitivity factors to vari-
13

ation of µ, respectively. Moreover, molecules have fine the prospects for µ-variation measurements. Coherent
and hyperfine structures, Λ-doubling, hindered rotation, control of molecular quantum states, which is a pre-
etc., which adds a variety of dependences on the funda- requisite for a “molecular lattice clock”, was achieved
mental constants (Chin et al., 2009). for Sr2 (McGuyer et al., 2015). Tests of µ-variation
The first experimental comparison of a molecular clock might be made using vibrational transitions in diatomic
to an atomic clock (Shelkovnikov et al., 2008) was ob- alkali-alkaline-earth molecules and alkaline-earth hydride
tained by comparing the frequency of a rovibrational molecular ions (Kajita et al., 2014).
transition in SF6 with the hyperfine transition of the Cs The leading systematic effects for realization of optical
clock. The measured rovibrational transition frequency clocks with rovibrational levels of the molecular ions H2+
in SF6 depends only on µ (Rydberg constant cancels out and HD+ were assessed by Schiller et al. (2014) and Karr
when comparing to any other transition) : (2014), who also discussed their potential sensitivity to
µ-variation. The principal issues limiting the accuracy of
1/2
such clocks involved effects due to light shifts. Ramsey-

me
νSF6 = A R∞ , type spectroscopy in a beam of metastable CO molecules
mp
was reported by (de Nijs et al., 2014) for further tests of
where A is a numerical factor. The resulting constraint variation of µ.
on the fractional temporal variation of the proton-to- Santamaria et al. (2014) discussed the design of an
electron mass ratio was reported as µ̇/µ = (−3.8 ± 5.6) × experiment aimed to constrain the fractional tempo-
10−14 yr−1 . While this limit is less stringent than that set ral variation of µ at a level of 10−15 /yr using spectro-
by optical clocks (Godun et al., 2014; Huntemann et al., scopic frequency measurement on a beam of cold CF3 H
2014), this study offers a clean separation of µ-variation molecules. Progress toward precision spectroscopic mea-
from α-variation. surement with ultra-cold deuterated ammonia, ND3 , for
Proposals for future tests of variation of fundamental future laboratory tests for variation of µ was reported in
constants with ultracold molecules were reviewed by Chin a paper by Quintero-Pérez et al. (2014) and references
et al. (2009) and we provide only a brief summary here. therein. Prospects for high-resolution microwave spec-
These proposals are based on the enhanced sensitivities troscopy of methanol, CH3 OH, and CD3 OH molecules
to α, µ, and mq /ΛQCD for accidentally closely spaced in a Stark-deflected molecular beam were discussed by
levels. The relative effect of α-variation in microwave (Jansen et al., 2013), but the precision must be signif-
transitions between very close and narrow rotational- icantly enhanced for laboratory tests. A current goal
hyperfine levels may be enhanced by 2-3 orders of magni- of methanol studies is to improve precision to reference
tude in diatomic molecules with unpaired electrons like the astrophysical searches of µ-variation described in
LaS, LaO, LuS, LuO, YbF, and similar molecular ions Sec. II.H.2.
due to accidental degeneracies of hyperfine and rota- An alternative proposal to test variation of fundamen-
tional structures (Flambaum, 2006). Degeneracies be- tal constants with atoms and molecules involves precise
tween the fine and vibrational structures within the elec- measurements of the scattering lengths in Bose-Einstein
tronic ground states of diatomic molecules, such as Cl+ 2, condensate and Feshbach molecular experiments (Chin
CuS, IrC, SiBr, and HfF+ , lead to enhanced sensitivi- and Flambaum, 2006; Gacesa and Côté, 2014). A mea-
ties to the variation of both α and µ (Beloy et al., 2010; surement of the scattering length accurate to 10−6 per-
Flambaum and Kozlov, 2007). Strong enhancements of formed near narrow Feshbach resonances in two consec-
α- and µ-variation effects in dihalogens and hydrogen utive years was estimated to probe the variation of µ at
halides, HBr+ , HI+ , Br+ + + + +
2 , I2 , IBr , ICl , and IF , were the level of 10−15 − 10−18 yr−1 depending on the choice
reported by Pašteka et al. (2015). The calculation of of atomic species (Chin et al., 2009).
Flambaum et al. (2013) demonstrated enhanced sensitiv- Recent advances in cooling and control of molecules
ity to the variation α and mq /ΛQCD in opposite-parity (Zeppenfeld et al., 2017) promise future progress in lab-
closely spaced levels of the 207 Pb19 F molecule due to a oratory tests of variation of fundamental constants with
near cancellation of the omega-type doubling and mag- molecules.
netic hyperfine-interaction-energy shifts.
Experiments with cold diatomic molecules Cs2 (De-
Mille et al., 2008b) and Sr2 (Zelevinsky et al., 2008) have H. Limits on variation of α and µ from quasar absorption
also been proposed. DeMille et al. (2008b) predicted that spectra
the splitting between pairs of Cs2 nearly-degenerate vi-
brational levels, where each state is associated with a dif- The discussion of Secs. II.D - II.G concerns with a ques-
ferent electronic potential, could be measured precisely tion: Do fundamental constants vary now? The depen-
enough to sense a fractional change of δµ/µ < −17
∼ 10 . dence of atomic and molecular spectra on fundamental
3 +
Detailed spectroscopy of the Cs2 a Σu state was per- constants may also be used to probe for their variation
formed by Sainis et al. (2012), who further discussed in a distant past, as far back as ∼ 10 billion years ago,
14

the scale given by the age of the Universe. The basic from these sources, since the atoms are relatively light,
idea is the same: to compare the spectra from two dif- Z ≤ 30, which generally leads to smaller values of K.
ferent times, but to increase the time separation δt from For example, the maximum ∆K difference for any two
δt = 1−15 years of the laboratory tests to δt = 3−13 bil- lines of Fe+ , is ∆K = 0.11, with an estimated 30% un-
lions of years. In practice, we need a particularly bright, certainty (Porsev et al., 2007). Another difficulty of the
distant astrophysical light source, such as a quasar, to many-multiplet method is ensuring that one compares
serve as a backlight of high-redshift gas clouds in which transition lines from the same object, i.e. at the same
atomic or molecular absorption spectra can be observed. redshift z. The advantage of SIDAM is using lines of
Emission spectra are also used in some studies. The sen- the same element, eliminating or simplifying this issue.
sitivities of those spectra to the variations of α and µ Another significant systematic arises from the assump-
are defined and calculated in the same way as for the tion of the isotopic-abundance ratios for each atom or
terrestrial experiments. ion used for the analysis in the distant past, in particu-
Due to the expansion of the Universe, all wavelengths lar, the 25,26 Mg/24 Mg (magnesium, Z = 12) ratios, and
of light λ from the Universe’s past are redshifted. A their possible deviations from the terrestrial values. This
cosmological redshift z is defined as the ratio issue, discussed by Kozlov et al. (2004), is further com-
plicated by the lack of isotope-shift measurements for a
λlab − λ number for transitions that are used in the quasar ab-
z= , (18)
λ sorption studies (Berengut et al., 2011b).
where λ is the wavelength of the absorbed/emitted light A large-scale many-multiplet analysis of the Keck
and λlab is the wavelength of the light which is observed telescope high-resolution Echelle-spectrometer (HIRES)
on Earth. A redshift of z = 1 means that a 500 nm ab- data from 143 absorption systems at 0.2 < z < 4.2, indi-
sorption wavelength is observed on Earth as 1000 nm cated a variation of α (Murphy et al., 2004):
instead. This corresponds to a “look back” time of ∼ 8 ∆α αobs − αlab
billion years (Pilipenko, 2013). = = (−0.57 ± 0.11) × 10−5 , (19)
α αlab
To separate the redshift, one needs to compare “an-
cient” and present wavelengths of at least two spectral where αobs corresponds to a value of α in the distant
lines that have different sensitivities to the constants of past, between 2 and 12.4 gigayears here, and the αlab is
interest. the current terrestrial value.
Uzan (2011) provided a detailed review of atomic and However, the analysis of data from 23 absorption sys-
molecular quasar absorption studies, so we will provide tems taken by the Very Large Telescope (VLT) ultravi-
only key points and more recent results here. olet and visual Echelle spectrograph (UVES) yielded a
null result
1. Limits on variation of α from quasar absorption studies of ∆α
atomic spectra = (−0.06 ± 0.06) × 10−5 , (20)
α

Quasar absorption studies of α-variation use alkali- for 0.4 < z < 2.3. (Chand et al., 2004; Srianand et al.,
doublet (Murphy et al., 2001b), many-multiplet (Webb 2004). This analysis was disputed by Murphy et al.
et al., 1999), and single-ion differential α-measurement (2007, 2008b,c), who obtained different results from an
(Levshakov et al., 2006) methods. The alkali-doublet analysis of the same data; this was followed by the reply
method uses the ns − np1/2 , ns − np3/2 fine-structure of Srianand et al. (2007).
intervals of alkali-metal atoms as a probe of α-variation. An intriguing solution to this discrepancy was sug-
The many-multiplet method is a generalization of this gested by Webb et al. (2011): since Keck and VLT data
approach which uses many atomic transitions with dif- come from different hemispheres, both results can be
ferent dependences on α, and yields more accurate re- made consistent by introducing a dipole spatial varia-
sults than the alkali-doublet method. The single-ion dif- tion of α; this topic is discusses further in Sec. II.I.
ferential α-measurement (SIDAM) method uses different The VLT data were reanalysed in the more recent work
transitions of one ionic species in an individual exposure, by Wilczynska et al. (2015). Considering both statisti-
in an attempt to reduce some of the systematics of the cal and systematic error contributions, Wilczynska et al.
many-multiplet method. It is mainly used with Fe+ (iron, (2015) obtained δα/α = (0.22 ± 0.23)10−5 , consistent
Z = 26) which has several transitions with both positive with the dipole spatial variation limits introduced by
and negative K, allowing one to compare lines within Webb et al. (2011).
a single spectrum. Most of the quasar-absorption stud- An impact of instrumental systematic errors on α-
ies with atoms are based on strong UV lines redshifted variation results obtained from atomic quasar-absorption
into the visible spectrum range. Unfortunately, these data was recently studied by Whitmore and Murphy
transitions depend weakly on α for most atoms visible (2015) using 20 years of archival spectra from VLT and
15

Keck spectrographs. Whitmore and Murphy (2015) con- Werner band lines are strong dipole-allowed absorption
cluded that systematic errors in their wavelength scales lines of the H2 molecule with λ = 910 − 1140 Å. All
were substantial and capable of significantly weakening of these transitions have weak dependence on µ, with a
the evidence for variations of α from quasar absorption maximum sensitivity coefficient ∆Kµ ∼ 0.06 (Kozlov and
lines. However, they still can not entirely explain the Levshakov, 2013).
Keck/HIRES result (19). Improved limits on the variation of µ are obtained
To summarize, atomic quasar-absorption data remains by going from optical to microwave frequencies, where
a subject of open controversy which requires further a number of molecular transitions are available with val-
study and future deployment of high-resolution ultra- ues of Kµ greater by factors of 100-1000. The dependence
stable spectrographs like ESPRESSO (for the VLT) and of microwave and submillimeter molecular transitions on
ELT-HIRES (Martins, 2015) for improved astrophysical fundamental constants was reviewed by Kozlov and Lev-
measurements. Laser frequency-comb techniques for pre- shakov (2013). The following molecules were considered:
cise astronomical spectroscopy were described by Murphy CH, OH, NH+ , C3 H, H3 O+ , NH3 (ammonia), H2 O2
et al. (2012). (hydrogen peroxide), CH3 OH (methanol), and CH3 NH2
(methylamine). Nine diatomic and 16 polyatomic molec-
ular candidates for µ-variation studies were reviewed by
2. Limits on variation of µ from quasar absorption studies of Jansen et al. (2014).
molecular spectra
In 2011, a number of polyatomic molecules, including
methanol and methylamine were observed for the first
Molecular spectra provide clean constraints on µ-
time at high redshift, z = 0.89, corresponding to look-
variation since rotational and vibration transitions have
back time of 7.5 × 109 years. The Kµ coefficient for
different µ-dependences given by Eq. (17). There are two
ammonia is -4.5 (Kozlov and Levshakov, 2013), which
main considerations when selecting molecules for astro-
represents a two orders of magnitude enhancement in
physical studies of µ-variation:
comparison with H2 . However, all of ammonia lines ex-
• How sensitive are the molecular transitions to vari- hibit the same sensitivity, so comparison with other sys-
ation of µ? This is quantified with a dimensionless tems is required. Two absorption systems are known
sensitivity factor Kµ , analogous to the K factor for with NH3 lines, at z = 0.69 and z = 0.89. Studies
sensitivity to α-variation. of µ-variation in these systems resulted in a 2σ limit
|∆µ/µ| < 1.8 × 10−6 (Murphy et al., 2008a) and 3σ limit
• How abundant is this molecule in the Universe? A of |∆µ/µ| < 1.4×10−6 (Henkel et al., 2009), respectively.
high sensitivity factor would be good for laboratory A joint three-component fit to the NH3 , CS, and H2 CO
tests of Sec. II.G, but useless for astrophysical stud- lines yielded |∆µ/µ| < 3.6 × 10−6 , for z = 0.69 (Kanekar,
ies if it is impossible to observe the corresponding 2011).
transitions. The sensitivity coefficients in methanol transitions
It is particularly advantageous if a molecule has several range from 17 to -43, potentially allowing for the max-
transitions with different Kµ , preferably of opposite sign. imum enhancement of |∆Kµ | ∼ 60 (Kozlov and Lev-
Then, transitions in the same molecule can be used for shakov, 2013)1 . In 2013, Bagdonaite et al. (2013b) set
the astrophysical search for µ-variation, reducing impor- the most stringent limits of past variation of µ, |∆µ/µ| <
tant systematic effects. 1 × 10−7 at (1σ), using four methanol lines at z = 0.89.
Until recently, all of the most accurate astrophysical An extended study of µ-variation based on 17 measure-
limits on the variation of µ came from H2 studies, recently ments of ten different absorption lines of methanol was
reviewed by Ubachs et al. (2016), since H2 is the most carried out by Bagdonaite et al. (2013a), allowing for a
abundant molecule observed, with 56 absorption systems quantitative analysis of previously unaddressed underly-
known at the present time. A combined H2 result from 10 ing systematic effects yielding ∆µ/µ = (−1.0 ± 0.8stat ±
systems with 2.0 < z < 4.2 sets the limit on the variation 1.0sys )×10−7 . Assuming a linear variation of µ with time,
of µ at this limit translates into µ̇/µ < 2 × 10−17 yr−1 which is
more constraining than the atomic clock limit (Godun

∆µ µobs − µlab et al., 2014; Huntemann et al., 2014) associated with the
= ≤ 5 × 10−6 (3σ), (21)
µ µlab same linear model of fundamental constant variation. We
note that there is no theoretical basis to assume the linear
where µobs corresponds to the value of µ in the distant
past, from 10 to 12.4 gigayears in this study, and µlab is
the current terrestrial value (Ubachs et al., 2016). These
molecular-hydrogen studies use the UV transitions in Ly- 1 We caution the reader that here µ = mp /me but µ = me /mp is
man and Werner bands that are redshifted into the visible frequently used in the literature, leading to opposite signs of the
spectrum. The B 1 Σ+ 1 + 1 +
u −X Σg Lyman and C Πu −X Σg
1 + Kµ coefficients in different sources.
16

variation of fundamental constants. We make such com- unless under extreme densities, such as in the vicinity of
parison only as an illustration of the accuracies reached a neutron star (Uzan, 2011). Therefore, the Webb et al.
by the astrophysical and laboratory studies. (2011) hypothesis of a dipole spatial variation of α intro-
In 2015, one of the four methanol lines observed at duced to explain the discrepancy between Keck and VLT
z = 0.89 and used in the analysis of this absorption data discussed in Sec. II.H.1 was quite extraordinary.
system, was noted to have a different line profile: the The spatial variation idea arises from the geograph-
line full widths at half-maximum was larger, at 4.3σ sig- ical positions of Keck and VLT telescopes in northern
nificance, suggesting that the sightline in this transition (Hawaii) and southern (Chile) hemispheres, respectively,
traces different absorbing gas from that detected in the separated by 45◦ in latitude. These two telescopes, on
other three lines (Kanekar et al., 2015). Therefore, it average, observe different directions in the sky and Keck
was recommended to exclude this line from the analysis, and VLT α-variation results can be made consistent by
resulting in a 2σ constraint of |∆µ/µ| < 4 × 10−7 . introducing a spatial variation of α. The result of Webb
Using combinations of atomic and molecular lines et al. (2011) would mean that α was larger in the past
allows one to probe variation of various combina- in one direction and smaller in the past in the opposite
tions of fundamental constants. A comparison of the direction according to
atomic hydrogen 21 cm hyperfine ground-state transi-
tion with atomic UV spectral lines (Rahmani et al., 2012; ∆α α(r) − α0
= = (1.10 ± 0.25) × 10−6 r cosψ Gly−1 ,
Tzanavaris et al., 2007, 2005) or OH molecular transi- α α0
tions (Chengalur and Kanekar, 2003) constrains combi- (22)
nations of α, µ, and the proton g-factor. where (α(r) − α0 )/α0 is a variation of α at a particular
Comparing the 21 cm line to molecular rotational tran- place r in the Universe relative to α0 on Earth at r = 0.
sitions in CO, HCO+ , and HCN eliminates the depen- The function r cosψ describes the geometry of the spatial
dence on µ, which is 1/µ for both types of transition variation, where ψ is the angle between the direction of
(Murphy et al., 2001a). the measurement and the axis of the dipolar variation.
The combination F = α2 µ was probed with a C+ and The distance function r is the light-travel distance r = ct
CO transitions (Levshakov et al., 2012, 2008), thus elim- measured in giga-lightyears. The Keck/VLT data were
inating the dependence on gp and yielding a constraint further analyzed in terms of spatial variation of α by
∆F/F < 2 × 10−5 at z = 5.2. Berengut et al. (2011d, 2012c) and King et al. (2012).
In summary, currently the best astrophysical con- A subsequent analysis of the Keck and VLT system-
straint on the µ-variation for high redshifts, up to z = atic instrumental errors by Whitmore and Murphy (2015)
4.2, (12.4 Gyr), come from the H2 data (Ubachs et al., weakened but not completely eliminated such a sce-
2016), while the strictest constraints for lower redshifts, nario. The extraordinary claim of the spatial variation
z = 0.89, are obtained from the methanol data (Bagdon- of α will require future extraordinary evidence obtained
aite et al., 2013a; Kanekar et al., 2015). Further improve- with next-generation ultra-stable high-resolution spec-
ment may come from observation of ammonia, methanol trographs and a higher level of control of systematic er-
and other more complicated molecules with high sensitiv- rors.
ities to µ-variation at higher redshifts, increased sensitiv- Nevertheless, the subject of the spatial variation of
ity and spectral resolution of astronomical observations fundamental constants is an interesting subject and
and increased precision of the laboratory measurements various scenarios for new physics could exist that may
for the most sensitive molecular transitions (Kozlov and be tested with astrophysics and laboratory studies.
Levshakov, 2013). Regardless of validity of the Webb et al. (2011) result,
we invite the reader to use it as an example to consider
the following question: What type of new physics can
I. Spatial variation of fundamental constants induce a spatial cosmological variation of fundamental
constants and how can we test for it? Berengut and
As discussed in Sec. II.C, if the fundamental constants Flambaum (2012), Berengut et al. (2012c), and Olive
depend on some dynamical scalar field φ they become et al. (2012, 2011) considered three scenarios, described
dynamical. A coupling of such scalar field φ to elec- below.
tromagnetic fields induces a coupling to matter which
may depend on the local matter density. Such density- I. Fundamental constants may fluctuate on a cosmo-
dependent couplings may lead to a spatial variation of logical scale involving regions not in causal contact
fundamental constants: fundamental constants will be due to super-Hubble quantum fluctuations of a light
different in the regions of high density of matter in com- field during inflation; further constraints from CMB are
parison to regions of low density. However, such spa- described by Sigurdson et al. (2003).
tial variation at the cosmological scales is expected to be
much smaller in most theories than a temporal variation II. A background value of φ depends on position and
17

time, i.e. there is a non-zero spatial gradient of the field domain walls are discussed in Sec. IX.
∇φ 6= 0. It was pointed out by Olive et al. (2011) that the Analysis of H2 molecular spectra in terms of spatial
generalization of the Copernican principle that assumes a dependence of µ on cosmological scales is presented by
homogeneous Universe at large scales is not fully satisfied Ubachs et al. (2016). Spatial variation of fundamental
in such a model and its theoretical foundation is unclear. constants may also manifest itself at local scales (Milky
Such a model will result in a dipole variation of the fun- Way and the Solar system). Two types of tests are being
damental constants in the general form of Eq. (22) with pursued with atoms and molecules described below.
the axis of the dipole being in the direction of the gra-
dient ∇φ. The spatial variation of fundamental constant
X is described by 1. Search for coupling of fundamental constants to a changing
gravitational potential
δX
= kX δφ, (23)
X First, a spatial change in fundamental constants may
be induced by light scalar fields that change linearly with
where kX is a dimensionless factor quantifying the spa-
changes in the local gravitational potential. This may
tial variation of the fundamental constant X. Assuming
be tested by searching for a dependence of fundamental
that the same field is responsible for the variation of all
constants on a varying gravitational potential.
fundamental constants, the direction of the dipole is the
Variations in fundamental constant X with the change
same for all fundamental constants.
in the gravitational potential are modeled as
Berengut and Flambaum (2012) proposed that such
a dipole variation can be tested using atomic clocks, ∆X ∆U (t)
quasar atomic and molecular spectra, the Oklo natural = kX , (24)
X c2
nuclear reactor, meteorite dating, and cosmic microwave
background. The Earth is moving along with the Sun where ∆U (t) is the variation in the gravitational poten-
with respect to the rest frame of the CMB and this tial. The goal of experiments is to measure or set con-
motion has a component along the direction of the φ straints on the quantities kX which quantify the cou-
gradient. This model results in a small spatial variation plings of various fundamental constants to the changing
as well as annual modulation of fundamental constants gravitational potential. Due to the eccentricity of the
with Earth motion around the Sun. The result of Webb Earth’s orbit around the Sun, the gravitational potential
et al. (2011) roughly translates into a α̇/α ∼ 10−19 y−1 has a seasonal 3% variation and a corresponding modula-
variation with a ∆α/α ∼ 10−20 annual modulation. tion of the constants may be studied with atomic clocks
Therefore, atomic clocks with high sensitivities to and other precisions instruments.
α-variation described in Sec. II.F.2 are particulary The idea for such experiment is illustrated in Fig. 4.
desirable for such tests. Present CMB constraint on the Blatt et al. (2008) searched for such change in fundamen-
dipolar modulation of α (corresponding to a gradient tal constants by monitoring the ratio of Sr and Cs clock
across the observable Universe) from 2015 Planck data frequencies. They combined their result with Hg+ /Cs
is (−2.7 ± 3.7) × 10−2 at the 68% confidence level (Adam (Fortier et al., 2007) and H-maser/Cs (Ashby et al., 2007)
et al., 2016). clock experiments to set constraints on the couplings of
fundamental constants α, 1/µ = me /mp (designated by
III. Olive et al. (2011) theorized that such spatial vari- the subscript µ in this section), and Xq = mq /ΛQCD (des-
ations of α may be a signature of a domain wall pro- ignated by the subscript q) to the changing gravitational
duced in the spontaneous symmetry breaking in the early potential defined by Eq. (24):
Universe, involving a scalar field coupled to electromag-
netism. In this scenario, there is no spatial gradient but kα = (2.5 ± 3.1) × 10−6 ,
a discontinuity in the values of fundamental constants kµ = (−1.3 ± 1.7) × 10−5 ,
at the domain wall (or walls) in our Hubble volume. kq = (−1.9 ± 2.7) × 10−5 ,
The fundamental constants on either side of the wall dif-
fer, and the quasar absorption spectra may not be actu- We note that decoupling of kq is not straightforward
ally testing deviation of α from the current Earth value, and is dependent on the nuclear model (Flambaum and
but probe locations of the domain walls in our Hubble Tedesco, 2006; Jackson Kimball, 2015). Only the depen-
volume. Attempts to fit Keck/VLT quasar absorption dence of the H-maser frequency on the light quark mass
data into the one or two-wall models faced difficulties was taken into account, but not of the Cs clock.
(Berengut et al., 2012c; Olive et al., 2012). A limit on the kα was obtained with a measurement of
Atomic clocks are only sensitive to such a scenario of frequencies in atomic dysprosium, which is only sensitive
spatial α-variation if the Earth actually passes thorough to α-variation (Ferrell et al., 2007).
a domain wall at the present time. Precision magnetome- Guéna et al. (2012) reported a limit on a fractional
tery and atomic clock experiments aimed at detection of variation of the Rb/Cs frequency ratio with gravitational
18

2009). The use of optical clocks based on the Yb+ oc-


tupole transition (Huntemann et al., 2016) as well as
new clock schemes with high sensitivity to α-variation
described in Sec. II.F.2 may significantly improve the
constraints.
Berengut et al. (2013b) proposed a new test of α-
variation in a strong gravitational field using metal lines
in the spectra of white dwarf stars. A goal of such studies
is to probe α-variation with gravitational potential me-
diated by a light scalar field at a much stronger, by five
orders of magnitude, gravitational potential than probed
by clock experiments. Laboratory measurements of rel-
FIG. 4 (color online). Earth orbiting the Sun (mass m ) in evant metal lines, such as the Fe4+ and Ni4+ (nickel,
gravitational potential; the orbit eccentricity is exaggerated Z = 28) spectra, are needed to improve on the results of
to show geometry. Picture credit: Jun Ye’s group and Greg Berengut et al. (2013b). Limits on a gravitational field
Kuebler, JILA. dependence of µ from H2 spectra in white dwarfs was
reported in Bagdonaite et al. (2014).

potential at the level of (0.11 ± 1.04) × 10−6 . A global


fit to available clock data yielded constraints similar to 2. Search for chameleons: testing the dependence of
the analysis of Blatt et al. (2008). Tobar et al. (2013) fundamental constants on the mass density of the environment
constrained a fractional variation of the Cs/H and Rb/H
frequency ratios with gravitational potential at the level Second, a spatial variation of fundamental constants
of 3.6(4.8) × 10−6 and 6.3(10) × 10−6 , respectively, over may result from a shift in the expectation value of
8 years of measurements. φ between dense and rarefied environments (Olive and
Peil et al. (2013) reported limit on fractional variation Pospelov, 2008), when coupling of field to matter depends
of the Rb/Cs, Rb/H, and Cs/H frequency ratios with the on its density, via, for example chameleon mechanism.
gravitational potential at the level of (−1.6 ± 1.3) × 10−6 , Such tests also probe local position invariance.
(−2.7±4.9)×10−6 , and (−0.7±1.1)×10−6 , respectively, In chameleon models, the scalar fields which are dark
over 1.5 years of measurements. Peil et al. (2013) per- energy candidates are ultra-light in a cosmic vacuum but
formed a global fit of constraints which included clock possess an effectively large mass when coupled to normal
data from Ashby et al. (2007); Blatt et al. (2008); Fortier matter (Joyce et al., 2015; Olive and Pospelov, 2008)
et al. (2007) and Guéna et al. (2012) which gave improved as discussed in Sec. IX, hence the “chameleon” name.
values: Chameleon dark matter models and relevant experimen-
tal tests have been recently reviewed by Joyce et al.
kα = (1.7 ± 7.5) × 10−7 , (2015). The chameleon mechanism can potentially sig-
kµ = (−2.5 ± 5.4) × 10−6 , nificantly affect quasar absorption spectra used to search
kq = (3.8 ± 4.9) × 10−6 . for variation of fundamental constants as well as compar-
ison of laboratory and astrophysics limits.
In 2013, a new Dy frequency measurement set an im- Here, we describe testing a class of chameleon mod-
proved limit on the kα (Leefer et al., 2013) els with scalar-field couplings to matter that are much
stronger than gravitational (Olive and Pospelov, 2008).
kα = (−5.5 ± 5.2) × 10−7 . In such a scenario, fundamental constants depend on the
local matter density ρ and one expects δα/α 6= 0 and
Clock experiments intended for the ACES space mis- δµ/µ 6= 0 for all interstellar clouds, when compared to
sion on the International Space Station (ISS) (Caccia- terrestrial laboratory values. This change is due to differ-
puoti et al., 2007) could improve upon the precision of ences in densities of the interstellar clouds and Earth en-
absolute redshift measurements. However, this does not vironments, ρ⊕ /ρcloud > 1010 (Levshakov et al., 2011b).
help differential measurements since the annual modula- The large matter density on Earth results in a screen-
tion of the gravitational potential due to the Sun is the ing of the cosmological chameleon field for terrestrial fre-
same for clock on Earth and ISS, and the orbit around quency measurements.
the Earth is circular. Molecular studies with CO, CH, ammonia (NH3 ) and
Further improvements may come from further optical- methanol CH3 OH have provided the most accurate lim-
clock tests and proposed space missions that would put its on matter-density couplings of fundamental constants
clocks in a highly eccentric earth orbit (Schiller et al., because of high sensitivity of some molecular absorption
2009) or a solar-system-escape trajectory (Wolf et al., spectra in our galaxy to µ-variation. Variation of the
19

quantity F = α2 µ with matter density was probed us- dated to the extent that the deduced values of α from
ing a combination of C+ and CO transitions (Levshakov different methods, one of which incorporates QED calcu-
et al., 2010c), yielding a constraint |∆F/F | < 3.7 × 10−7 . lations, agree with each other, as described in Sec. III.B.
The best-quality radio astronomical data for methanol The comparisons are affected by the uncertainties in the
lines were used to constrain the variability of µ̄ = 1/µ in values of fundamental constants [such as masses, Ry-
the Milky Way at the level of |∆µ̄/µ̄| < 2.8 × 10−8 (Lev- dberg constant, etc.] and by the uncertainties in the
shakov et al., 2011a). This result can be further improved strong-force [hadronic] contributions beyond QED.
with better laboratory spectroscopy of the CH3 OH mi- In general, one distinguishes between free-particle
crowave lines. properties, such as the anomalous magnetic moment of
In 2010, Levshakov et al. (2010a,b) reported a surpris- the electron, and bound-state QED [Lamb shift being the
ing non-zero ∆µ̄/µ̄ = (26 ± 1stat ± 3sys ) × 10−9 result for most prominent example].
µ-variation using the ammonia method. This approach Bound-state QED can be tested in a number of sys-
involves observations of the NH3 inversion lines comple- tems and we highlight the advantages of various ap-
mented by rotational lines of other molecules in the Milky proaches below. Such tests are expected to be more ac-
Way and comparing these frequencies with terrestrial val- curate in light systems such as H, D, 3 He+ , He, positro-
ues. In 2010-2013, Levshakov et al. (2013) carried out ad- nium (Ps) and muonium (Mu), where the contribution
ditional observations in the Milky Way to test for hidden of inter-electron interaction is either absent or can be
errors and found a systematic error in the radial veloci- evaluated to high accuracy. QED tests with these sys-
ties of earlier studies. A revised value of ∆µ̄/µ̄ < 2×10−8 tems were reviewed by Karshenboim (2005). The rela-
at the 3σ confidence level obtained using the ammonia tive importance of inter-electron interactions is also re-
method was reported in Levshakov et al. (2013), resolv- duced in highly charged ions [HCIs]. However, in HCIs
ing the discrepancy. the nuclear structure uncertainty is the limiting factor
A spectroscopic method for pulsed beams of cold and QED calculation in heavy ions require a develop-
molecules was developed by Truppe et al. (2013) and ap- ment of non-perturbative methods. Interesting interme-
plied to measure the frequencies of microwave transitions diate cases are few-electron systems, where the electron-
in CH. Comparing new CH values and OH laboratory electron correlations must be taken into account on par
results (Hudson et al., 2006) with those measured from with the nominally field-theoretic (QED) contributions.
Milky Way sources of CH and OH, Truppe et al. (2013) High precision QED atomic calculations for Li and Be+
constrained the variation of α and µ between the high- were carried out by Yan et al. (2008, 2009) and result-
and low-density environments of the Earth and the inter- ing energies were found to be in good agreement with
stellar medium at the levels of ∆α/α = (0.3 ± 1.1) × 10−7 experiment, with the exception of the Be+ ionization
and ∆µ̄/µ̄ = (−0.7±2.2)×10−7 . Sensitivities for relevant potential. QED corrections to the 2p fine structure in
transitions were calculated by (Kozlov, 2009). Li were calculated by Puchalski and Pachucki (2014),
yielding the splitting value with 6 × 10−6 uncertainty, in
agrement with recent high-precision experiment (Brown
III. PRECISION TESTS OF QUANTUM et al., 2013a,b). Precision test of many-body QED was
ELECTRODYNAMICS reported by Nörtershäuser et al. (2015) using the Be+ 2p
fine-structure doublets measured in short-lived isotopes.
A. Introduction Simple molecules, H2 , HD, and D2 , and H2+ , HD+ molec-
ular ions (Biesheuvel et al., 2016; Dickenson et al., 2013;
In this section, we give an overview of low-energy preci- Salumbides et al., 2011) offer additional QED tests.
sion tests of Quantum Electrodynamics (QED) with tools Below we highlight a few recent examples of precision
of atomic physics. Historically, QED is the first relativis- QED tests and review the recent progress in QED tests
tic quantum field theory, which laid the foundation of the with HCIs.
modern formalism of the Standard Model. It is arguably
the most stringently tested sector of the Standard Model.
We focus on recent results and existing inconsistencies.
B. Anomalous magnetic moment of the electron
The reader is referred to textbooks [e.g., Bjorken and
Drell (1964); Peskin and Schroeder (1995)] for a general
At present, the most accurate contributions to the de-
introduction to QED and recent reviews by Beiersdorfer
termination of α come from comparison of theory and
(2010); Drake and Yan (2008); Eides et al. (2001); and
experiment for the electron magnetic-moment anomaly
Karshenboim (2005) for a more detailed exposure.
ae (Mohr et al., 2016, 2012). This quantity is defined as
Precision tests of QED are carried out by comparing
follows. The magnetic moment of the electron is
experimental results with theoretical predictions. For ex-
ample, QED predictions depend on the value the electro- e
magnetic fine-structure constant α. QED is then vali- µe = ge s, (25)
2me
20

where ge is the (dimensionless) electron g-factor, me is its example: it has the same discrete spectrum as the hy-
mass, and s is its spin. The magnetic-moment anomaly drogen atom in nonrelativistic quantum mechanics, up
ae is defined by to a multiplicative factor of (mp + me ) /2mp . Wheeler
(1946) used a simple variational calculation to show that
|ge | = 2(1 + ae ). that Ps− should also be stable, and Hylleraas and Ore
(1947) determined that Ps2 should be stable. These three
The solution of the Dirac equation for a free electron
species were subsequently found experimentally. Reviews
gives ge = −2 and thus ae = 0. In the Standard Model,
of developments in this field up to 2012 were given by
ae 6= 0: it is given by
Karshenboim (2005) and Namba (2012), and of more re-
ae(th) = ae (QED) + ae (weak) + ae (hadronic), cent work by Nagashima (2014) and Mills (2014).
As purely leptonic systems, polyelectrons offer a
where the three terms account respectively for the purely testbed for precision comparison of QED theory with ex-
quantum electrodynamic, predominantly electroweak, periment, particularly for bound-state systems. We re-
and predominantly hadronic contributions [using the no- view recent results and future prospects below. There
tation of Mohr et al. (2016)]. The ae (QED) contribution is no experimental evidence for more complex polyelec-
depends on α and can be expressed as a powers series trons. Frolov and Wardlaw (2008) suggest that Ps− 2 and
of α/π whose coefficients are calculated from QED. The Ps3 should be stable, but Varga (2014) and Bubin et al.
dependence of ae on α for the other two contributions is (2013), respectively, consider these two species to be un-
negligible. stable.
The most accurate measurement of ae was carried out
with a single electron that was suspended for months at a
time in a cylindrical Penning trap (Hanneke et al., 2008). 1. Positronium
The ratio of electron spin-flip frequency to the cyclotron
frequency in the trap determines ae . The resulting value Positronium (Ps), the atom consisting of one electron
of α extracted by combining the 2008 measurement (Han- and one positron, was first identified in the laboratory
neke et al., 2008) and theoretical result for ae is by Deutsch (1951). It is a system in which bound-state
QED has been studied with precision. Most recently, the
1/α = 137.035999084(51), (26) structure of the lowest triplet term of Ps was measured
which has a relative uncertainty of 3.7 × 10−10 . The by optical spectroscopy (Cassidy et al., 2012c), and the
theoretical uncertainty contribution is 2.8 × 10−10 . transition between the triplet and singlet terms of the Ps
An alternative determination of α, accurate to 6.6 × ground state has been observed directly (Miyazaki et al.,
10−10 , is obtained from a completely different approach 2015; Yamazaki et al., 2012). This energy splitting is a
and is based on the precise atom-interferometric mea- benchmark for first-principles QED calculations of two-
surement of the ratio of Planck’s constant to the mass of particle systems. It has been calculated by QED theory
the 87 Rb atom (Bouchendira et al., 2011). up to order α6 to an accuracy of 1 ppm. The result dif-
The agreement of these two determinations of α, from fers by 4σ from the experimental determination, which

the measurements of ae and of h/m 87 Rb , validates the presently has an uncertainty of around 3 ppm (Cassidy
theoretical QED calculation of ae (Aoyama et al., 2008). et al., 2012c). Improvements in precision are required
This, in turn, provides the most accurate test of quantum to resolve this discrepancy. [A more recent experiment
electrodynamics and the SM to date. We emphasize that does claim to have a result closer to theory (Ishida, 2015;
ae is calculated in terms of the fundamental constant α, Ishida et al., 2014)]. There are suggestions about beyond
but α is a SM parameter as it cannot be calculated from SM physics mechanisms to which positronium might be
the first principles. particularly susceptible (Lamm, 2017), and there are sub-
stantial efforts to extend QED theory to order α7 in or-
der to sharpen the comparison with experiment(Adkins
C. Quantum electrodynamics tests with polyelectrons et al., 2015; Eides and Shelyuto, 2017).
Another noteworthy recent development is the advent
In 1946, Wheeler denoted as “polyelectrons” all bound of Ps Rydberg spectroscopy, in which states have been
complexes consisting of only electrons and positrons resolved with principal quantum numbers n as large as
(Wheeler, 1946). Although all such complexes are likely 50 (Alonso et al., 2015; Cassidy et al., 2012a; Jones et al.,
unstable with respect to electron-positron annihilation 2014; Wall et al., 2015). Such states may be of funda-
into gamma rays, there are some that are stable with mental interest for testing QED, as some QED correc-
respect to dissociation into simpler complexes, and thus tions appear at lower orders of α than they do for the
may live sufficiently long to have physical and even chem- ground state (Lamm, 2017). Ps Rydberg states can also
ical significance. Positronium (Ps), the “atom” consist- have longer lifetimes than the n = 1 ground state, since
ing of one electron and one positron, is the simplest the electron-positron annihilation rate is proportional to
21

the probability density at the point of contact, which D. Proton radius puzzle
scales as n−3 . This could be advantageous for precision
spectroscopic study, or for use of Rydberg Ps states as The proton radius puzzle, as it is known colloquially,
reservoirs for positrons used to produce the antihydro- has perplexed the physics community for over half a
gen required for the studies described in Sec. VI. If Ps decade (Carlson, 2015; Hill, 2017; Jentschura, 2011; Pohl,
could be placed in highly “circular” Rydberg states, it 2016; Pohl et al., 2013a). The highly-precise r.m.s. charge
could be a candidate for studies of the Einstein equiv- radius rp = 0.84087(39) fm extracted from the 2S − 2P
alence principle for a mixed matter-antimatter system, Lamb shift in muonic hydrogen (Antognini et al., 2013;
either via free–fall measurements or gravitational quan- Pohl et al., 2010) is in significant disagreement with the
tum state spectroscopy (Dufour et al., 2015). result rp = 0.8758(77) fm deduced from spectroscopy
with ordinary hydrogen (Mohr et al., 2012). The lat-
ter value is also supported by electron scattering experi-
ments, further exacerbating the problem (Arrington and
2. Positronium anion, Ps−
Sick, 2015; Bernauer et al., 2010; Mohr et al., 2012). This
outstanding discrepancy has prompted speculations that
In the laboratory, the positron that ends up in Ps is
the discrepancy may hint at physics beyond the Stan-
typically born at an energy 0.5 − 2 MeV by beta decay of
dard Model [see e.g., Dahia and Lemos (2016b); Liu
a radionuclide such as 22 Na. The positron is moderated
et al. (2016); Onofrio (2013)]. Resolution may be more
down to ∼ 10 meV energies by passage through mat-
mundane, such as missing systematic corrections both
ter, at which stage it can be controlled by conventional
in theory and experiment or incorrect value of the Ry-
electron-optical techniques for use in scattering exper-
dberg constant (Pohl et al., 2013b). In fact, Czarnecki
iments, or produce Ps by electron capture form solids
and Szafron (2016) have recently pointed out that light-
(Charlton and Humberston, 2000; Mills, 2014). It was
by-light scattering diagrams have been erroneously ne-
first obtained in the laboratory by Mills in 1980 (Mills,
glected in the computations of the Lamb shift; these
1981). An experimental breakthrough in 2008 made it
authors estimate that such contributions decreases the
possible to generate the positronium anion, Ps− , with
theoretical prediction for the 1S − 2S splitting in hy-
efficiencies above 1%, i.e. for every 100 positrons enter-
drogen by an amount 28 times larger than the exper-
ing the moderator, one Ps− emerges (Nagashima, 2014).
imental error (Parthey et al., 2011). Interestingly, the
This development transformed the study of Ps− , for ex-
same muonic-hydrogen collaboration (CREMA) has re-
ample, making possible the observation of a shape res-
ported (Pohl et al., 2016) the value of deuteron radius
onance in its photodetachment (Michishio et al., 2016).
that shows a similar discrepancy with results of deu-
It also provides a means for producing energy-tunable
terium spectroscopy. Ongoing spectroscopic experiments
beams of Ps, by applying standard acceleration proce-
on hydrogen and simple hydrogenic ions such as He+
dures to Ps− and then neutralizing it by photodetach-
aim to unravel this mystery [see, for example,Yost et al.
ment. There is a considerable body of theory on the
(2016)].
structure of Ps− , including treatment of QED corrections
(Drake et al., 2002; Frolov, 2005). In time, the postron-
ium anion may become a benchmark for testing QED in
E. Tests of QED in highly charged ions
three-particle systems.
QED tests in highly charged ions were recently re-
viewed by Beiersdorfer (2010) and Volotka et al. (2013),
3. Diatomic positronium, Ps2 and we focus on key results and new developments. The
spectroscopies properties involved in the HCI QED tests
The Ps2 molecule was also observed, both in its ground are atomic transition energies, hyperfine splittings and
state (Cassidy and Mills, 2007) and in an L = 1 bound g factors. A particular attraction of the HCIs tests is a
excited state that was predicted by Varga et al. (1998) possibility to test two-loop QED contribution and many-
and Usukura et al. (1998) and subsequently observed by body QED corrections.
Cassidy et al. (2012b). The wavelength of the ground
- excited state transition in Ps2 was predicted to be
250.9179(11) nm. The observed wavelength reported by 1. Energies
Cassidy et al. (2012b) is 250.979(6) nm. At present, the
reason for the difference is not understood in detail. The The ground-state Lamb shift in H-like uranium (U,
Ps2 is thought to be located in a porous silica matrix, Z = 92) was measured by Gumberidze et al. (2005) with
which has been found to produce shifts in Ps transition 1% uncertainty, 460.2 ± 4.6 eV. The theory prediction is
wavelengths that are comparable to the difference be- 464.26(50) eV, with QED contributing 265.2 eV [of which
tween the theoretical and observed values for Ps2 . -1.26(33) eV is due to 2nd order QED] and finite nuclear
22

size contributing 198.81 eV (Yerokhin et al., 2003). Com- Shabaev (2015). This resolved previously-reported dis-
bining theory and experiment provided a test of QED at agreements between the numerical all-order and ana-
the 2% level. lytical Zα-expansion approaches, which were caused by
The 2S − 2P1/2 transition energy in Li-like U89+ , unusually large higher-order terms omitted in the Zα-
280.645(15) eV, was measured with much higher, expansion calculations. This work eliminated the second-
0.005% precision, in agreement with theoretical value largest theoretical uncertainty in the 1S and 2S Lamb
of 280.71(10) eV (Kozhedub et al., 2008). Li-like ura- shift of H.
nium study tested second-order (in α) QED effects to 6%
(Volotka et al., 2013). Theoretical accuracy of HCI QED
tests is limited by the nuclear polarization correction. 2. Hyperfine splittings
The experimental accuracy is much higher for lighter
ions. The 1s2p 1 P1 − 1s2 1 S0 resonance line in He- A number of measurements of the hyperfine splitting in
like Ar16+ was measured with a relative uncertainty of H-like HCIs with about 1 × 10−4 uncertainty [for exam-
2 × 10−6 for a test of two-electron and two-photon QED ple, measurements in 203 Tl80+ and 205 Tl80+ (thallium,
radiative corrections (Bruhns et al., 2007). The exper- Z = 81) by Beiersdorfer et al. (2001)] motivated cor-
imental value was in perfect agreement with theoretical responding theoretical efforts. Since the theoretical un-
prediction (Artemyev et al., 2005), as well as with a later certainty is dominated by the correction due to nuclear
1.5 ppm measurement of 3139.581(5) eV (Kubiček et al., magnetization distribution (the Bohr-Weisskopf effect),
2012). However, a measurement of the He-like Ti (tita- Shabaev et al. (2001) proposed to consider a specific dif-
nium, Z = 22) resonant line, 4749.85(7) eV, by Chantler ference of the ground-state hyperfine splitting in Li-like
et al. (2012) differed by 3σ from the theoretical predic- ion, ∆E(2s), and H-like ion, ∆E(1s), for the same nu-
tion (Artemyev et al., 2005). Chantler et al. (2012) noted cleus:
that there appeared to be an evidence for a Z-dependent
∆0 E = ∆E(2s) − ξ∆E(1s). (27)
divergence between experiment and calculation in He-
like isoelectronic sequence with Z > 20. This analysis The parameter ξ introduced to cancel the Bohr-
was disputed by Epp (2013), in particular the omission Weisskopf effect can be calculated with high precision.
of the Kubiček et al. (2012) value from the fit. This is- In 2012, the theoretical accuracy of the specific difference
sue was addressed by Chantler et al. (2013) and further between the hyperfine splitting values of H- and Li-like
discussed by Gillaspy (2014), indicating need for further Bi (bismuth, Z = 83) ions was significantly improved [to
experimental and theoretical work. Measurement of the a relative uncertainty of ∼ 10−4 ] due to a new evalu-
resonant line in He-like Fe24+ (Kubiček et al., 2014) was ation (Volotka et al., 2012) of the two-photon exchange
found to be in agreement with theory (Artemyev et al., corrections to the hyperfine structure in Li-like ion. Mea-
2005) and inconsistent with a claim of systematic diver- surements of the H-like and Li-like Bi hyperfine splittings
gence between theory and experiment (Chantler et al., at the 10−6 level will allow probing the many-body QED
2012) at a 3σ level. The other Fe24+ spectroscopy data effects at a few percent level (Volotka et al., 2013).
(Rudolph et al., 2013) are also consistent with QED the- Hyperfine splitting of the 2s and 2p1/2 levels in Li-like
ory values. and Be-like ions of 141 Pr were measured by Beiersdorfer
The energy of the 1s2s 3 S1 − 1s2 1 S0 magnetic dipole et al. (2014) using high-resolution spectroscopy of the
transition in helium-like Argon was measured to 2.5 ppm 2s − 2p1/2 transition in the extreme ultraviolet region
accuracy by Amaro et al. (2012), differing by 1.6 σ (EUV). This work demonstrated that EUV spectroscopy
from the theoretical prediction of Artemyev et al. (2005). can be used to measure the hyperfine structure in high-Z
Even higher precision of 0.6 ppm was achieved for ions with a few valence electrons at the meV level.
1s2 2s2 2p 2 P3/2 - 2 P1/2 transition in boron-like Ar13+
ions, 441.25568(26) nm (Mäckel et al., 2011). The the-
ory value (Artemyev et al., 2007) is in agrement with 3. QED tests for g factors
the experiment, but two orders of magnitude less accu-
rate. Since nonrelativistic energies of p1/2 and p3/2 states The most stringent bound-state QED test of the
are the same, this transition energy is determined by the ground state g factor for a three-electron systems was car-
relativistic and QED effects, making it an excellent can- ried out for Li-like 28 Si11+ (silicon, Z = 14) by Volotka
didate for precision QED tests. Experimental accuracy et al. (2014); Wagner et al. (2013). The g-factor measure-
can be significantly increased by recent demonstration of ment carried out in a Penning trap, 2.000 889 889 9(21)
sympathetic cooling of HCIs (Schmöger et al., 2015), and was found to be in excellent agreement with the theoret-
theory accuracy urgently needs to improve. ical value 2.000 889 909(51) (Wagner et al., 2013). The
A high-precision nonperturbative (in Zα) calculation theory precision was further improved by Volotka et al.
of the nuclear-recoil effect on the Lamb shift of light (2014) due to rigorous QED evaluation of the two-photon
hydrogenic atoms was carried out by Yerokhin and exchange corrections to the g factor, yielding 2.000 889
23

892(8). A comparison of this new theoretical value with observing PNC in atoms. He concluded that the effect
the experimental result (Wagner et al., 2013) provides was too small to be of experimental significance. How-
tests of relativistic interelectronic interaction at the 10−5 ever, Bouchiat and Bouchiat (1974) realized that PNC is
level of precision, the one-electron bound-state QED in amplified in heavy atoms. They showed that the relevant
the presence of a magnetic field at the 0.7% level, and PNC amplitude scales steeply with the nuclear charge Z,
the screened bound-state QED at the 3% level. roughly as Z 3 . PNC amplitudes in heavy atoms, such as
For heavy ions, a specific difference scheme similar to Cs, are enhanced by over those in hydrogen by a factor of
Eq. (27) can be employed to largely cancel the nuclear 105 − 106 . This crucial observation enabled experiments
effects in the g-factor HCI calculations (Shabaev et al., on APV. In atomic physics, the first P -violating signal
2002) for further QED tests. A possibility for a deter- was observed in 1978 by Barkov and Zolotorev (1978) in
mination of α in bound-electron g factor experiments via Bi. In the same year, parity violation was reported in
the study of a specific difference of the g factors of B-like inelastic electron-deuterium scattering (Prescott et al.,
and H-like ions, for the same isotope with zero nuclear 1978). Direct observations of the charged W± boson and
spin in the Pb region, was discussed by Shabaev et al. neutral Z boson (responsible for APV) were not made
(2006). until 1983 when they were observed at CERN’s proton-
In 2014, Volotka and Plunien (2014) performed a sys- antiproton collider (Arnison et al., 1983a,b).
tematic study of the nuclear polarization effects in one- Over the following decades, AMO experiments were re-
electron and few-electron heavy ions, which included the fined, with PNC effects observed in several atoms. The
calculation of the nuclear polarization corrections to the most accurate measurement to date was performed in
binding energies, the hyperfine splitting, and the bound- Cs (Wood et al., 1997) and the most recent reported
electron g factor in the zeroth and first orders in 1/Z. measurement is in ytterbium (Yb, Z = 70) (Tsigutkin
Strong cancellation of nuclear polarization effects deter- et al., 2009). New experiments on atomic and molecu-
mining the ultimate accuracy of the QED tests was ob- lar PNC are underway or in planning stages as described
served in all cases for the specific differences described in Sec. IV.D. There are a number of extensive review
above. articles of APV (Bouchiat and Bouchiat, 1997; Budker,
1999; Derevianko and Porsev, 2007; Ginges and Flam-
baum, 2004; Roberts et al., 2015) as well as a mono-
F. Conclusion
graph (Khriplovich, 1991). Basics of electroweak theory
can be found in a number of textbooks, e.g., Commins
Finally, we would like to emphasize that the detailed
and Bucksbaum (1983) which has a discussion on APV.
understanding of atomic structure and, in particular,
QED contributions, is crucial for a number of precision The parity transformation P , or spatial inversion, is
tests of physics beyond the SM. QED contributions are equivalent to mirror reflection and rotation by 180◦ . The
needed for determining fundamental constants. While eigenvalues of P are ± 1, referred to as even and odd,
the stark discrepancies in proton and deuteron radii respectively. Under this transformation, all position vec-
determinations from various methods have spurred re- tors r change sign: r → −r, while spin and orbital angu-
examinations of both theory and experiment, the puz- lar momenta remain unaffected. Electric and magnetic
zle still remains unresolved. A number of technologi- fields are transformed as E → −E and B → B.
cal advances, such as high-precision Ramsey-comb spec- The QED Lagrangian governing AMO physics com-
troscopy at deep ultraviolet wavelengths (Altmann et al., mutes with P , leading to distinct selection rules in atomic
2016) and HCI trapping (Schmöger et al., 2015) are an- physics. For the electronic configuration n1 `1 . . . nNe `Ne
ticipated to extend the QED tests to new regimes. of Ne -electron
PNe
atom, the parity eigenvalue is given by
Π = (−1) i=1 `i . The parity of a given configuration
is determined by the parity of the open electron shell,
IV. ATOMIC PARITY VIOLATION
e.g. the parity of the [Hg]6p3 4 S◦3/2 ground state of Bi
A. Introduction
is odd. The conventional spectroscopic notation of elec-
tronic terms arising from a given electronic configuration
In this section, we give an overview of parity violation includes the label “◦” for odd parity states.
in atomic and molecular physics. This field is generally Laporte (1924) discovered parity conservation in atoms
referred to as parity non-conservation (PNC) or atomic from analysis of the iron spectrum, and formulated a rule:
parity violation (APV) in the literature. The field of electric dipole transitions between states of like parity are
parity violation started with the seminal paper by Lee strictly forbidden. To see this, consider the electric-dipole
(E1)
and Yang (1956) and discovery of PNC in nuclear β- (E1) transition amplitude Tf i = hΨf | D |Ψi i, where
decay (Wu et al., 1957), followed by the Nobel Prize in the atomic states Ψi,f are parity eigenstates and D is the
physics awarded to Yang and Lee in 1957. Soon after this electric dipole moment operator. Inserting the identity
discovery, Zeldovich (1959) contemplated a possibility of 1 = P †P ,
24

Ae Ve the electron vector currents to quark axial-vector cur-


e e e e e e e Ae e
rents. These interactions and constants could be fur-
ther combined into couplings to protons and neutrons of
Z0 Z0 Z0 atomic nuclei (Marciano and Sanda, 1978), e.g.,
(1)
N VN N N AN N N N N
VN N
Cp(1) = 2Cu(1) + Cd ,
(1)
(a) (b) (c) (d) Cn(1) = Cu(1) + 2Cd ,

FIG. 5 (Color online) Major diagrams contributing to the reflecting the quark composition of nucleons. Explicitly
parity violation in atoms. N and e− label nucleons and in terms of the Weinberg angle θW :
atomic electrons. Ae,N and Ve,N denote axial-vector and vec-
tor currents. (a) Z-boson exchange between electron axial- 1
Cp(1) = 1 − 4 sin2 θW ,

vector and nucleon vector currents (An Ve ); (b) Z-boson ex- 2
change between nucleon axial-vector and electron vector cur- 1
rents (Vn Ae ); (c) Electromagnetic interaction of atomic elec- Cn(1) =− ,
2
trons with the nuclear anapole moment (shown as a blob); (d)
Combined effect of the An Ve diagram (a) and hyperfine inter-
Cp(2) = −Cn(2) = gA Cp(1) ,
action. The vertical line separates nuclear spin-independent
where gA ≈ 1.26 is the scale factor accounting for the
(a) and spin-dependent (b)–(d) diagrams.
partially conserved axial vector current and sin2 θW =
0.23126(5) (Olive et al., 2015). Since sin2 θW ≈ 1/4, the
(1)
Cn contribution dominates HPV except for the 1 H atom.
The main diagrams contributing to PNC processes in
(E1) atoms are shown in Fig. 5. The HPV terms discussed
Tf i = hΨf | P † P D |Ψi i
(E1)
above are illustrated by diagrams (a) and (b). In addi-
= hP Ψf | P (D |Ψi i) = −Πf Πi Tf i . tion, there is also a contribution from the nuclear anapole
(E1) moment (c) and a combined effect of Z-boson exchange
Now if the two states have the same parities, Tf i = and hyperfine interaction (d). The effective weak Hamil-
(E1) (E1)
−Tf i and thereby Tf i
= 0. If parity is not con- tonian arising from diagram (a) does not depend on the
served, the eigenstates of the full atomic Hamiltonian no nuclear spin, while that from the set of diagrams (b)–(d)
longer possess a well defined parity. In other words, PNC does. We will consider the former in Sec. IV.B and the
leads to [usually small] mixing of opposite-parity states latter in Sec. IV.C.
(E1)
leading to non-vanishing values of Tf i . The theory and
experiments described below show how Laporte’s rule is
violated in atoms and molecules. B. Nuclear-spin independent effects
Microscopically, APV is caused by the weak interaction
1. Overview
mediated by the exchange of a Z boson. Since the range
of this interaction is ∼ ~/(mZ c) ∼ 2 × 10−3 fm [mZ ≈
The dominant contribution to parity violation in atoms
91 GeV/c2 is the mass of the Z boson], it is essentially
arises from the electron axial-vector – nucleon-vector
a contact interaction on the scale of atomic distances.
term in HPV , Fig. 5(a). If we treat the nucleon mo-
The relevant contact contribution to the SM Hamiltonian
tion non-relativistically, average over the nucleon distri-
density reads (Marciano, 1995)
bution, and neglect the difference between proton and
GF X  (1)  neutron distributions, we reduce the corresponding part
HPV = √ Cq ēγµ γ5 e q̄γ µ q + Cq(2) ēγµ e q̄γ µ γ5 q ,
2 q of HPV to an effective weak Hamiltonian in the electron
sector
(28)
where the Fermi constant GF
HW = QW √ γ5 ρ (r) , (29)
GF ≈ 1.17 × 10−5 (~c)3 GeV−2 = 2.22 × 10−14 a.u. 8

determines the overall strength of the weak interaction, where ρ (r) is the nuclear density and QW is a nuclear
the summation is over quark flavors, q = {u, d, s, ...}, e weak charge. The non-relativistic limit of the operator
and q are field operators for electrons and quarks respec- γ5 ρ (r) is
tively, γµ are Dirac matrices, and γ5 is the Dirac matrix
1
associated with pseudoscalars. [2ρ(r)(σ · p) − i(σ · ∇ρ)] ,
The coupling of the electron axial-vector currents to 2c
the quark vector currents is parametrized by the con- where p is the linear momentum operator and σ are elec-
(1) (2)
stants Cq ; the constants Cq describe the coupling of tron Pauli matrices.
25

The nuclear weak charge QW entering the effective experiments. Additional interference techniques are de-
weak Hamiltonian is scribed in Sec. IV.D.
QW ≡ 2Z Cp(1) + 2N Cn(1) ,
where Z and N are the numbers of protons and neu- 2. Parity violation in cesium
trons in the nucleus. Electrons predominantly couple
to neutrons and QW ≈ −N . This is a “tree-level” [or The measurement of APV in 133 Cs (Wood et al.,
the lowest-order] value of QW ; more accurate values in- 1997) is the most accurate to date, and supplemented
clude SM radiative corrections (Marciano, 1995), which with sophisticated atomic theory, it probes the SM low-
are typically a few percent and can be computed to high energy electroweak sector with exquisite precision. An
accuracy. A major theme in APV is a comparison of the alkali-metal atom with 55 electrons, Cs has a single va-
extracted QW with its SM calculated value: a difference lence electron outside a tightly-bound Xe-like core: its
between the two values can indicate physics beyond the ground electronic level is designated [Xe]6s 2 S1/2 , some-
SM. times called 6S1/2 . We focus on the optical transition
HW is a pseudo-scalar operator with its matrix ele- between a 6S1/2 ground state and an excited state of the
ment accumulated inside the nucleus. Its largest matrix same parity, 7S1/2 . This transition

is E1-forbidden
due
element is between s1/2 and p1/2 atomic orbitals. Since to the parity selection rule: 6S1/2 D 7S1/2 = 0. The
parity is no longer a good quantum number, yet the total weak interaction leads to an admixture of states of oppo-
angular momentum J is conserved, atomic states of nom- site parity: P1/2 states mix with the S1/2 states, leading
inal parity acquire admixtures of state of opposite parity to a small E1 transition amplitude EP V ∼ 10−11 a.u..2
with the same J. The relative size of the admixture is The Stark interference technique mentioned in
governed by the ratio of the matrix element of HW [typ- Sec. IV.B.1 is used to amplify the parity-violating signal.
ically ∝ Z 3 ] to the energy splitting between the nearby Application of an external electric field E induces an ad-
states of opposite parity [typically ∼ 1 a.u.]. ditional admixture of P states. This provides a strong E1
Since GF ∼ 10−14 a.u., matrix elements of HW are ex- pathway with a transition amplitude βE, where β is the
ceptionally small [∼ 10−11 a.u. for Cs] compared to the vector transition polarizability. The optical excitation
typical 1 a.u. transition amplitudes in atomic physics. To rate for the 6S1/2 −7S1/2 transition is proportional to the
amplify the PNC signal, all experiments rely on an in- square of the transition amplitude, β 2 E 2 +(βEEP V +c.c),
terference technique, where the HW -induced amplitude where the term quadratic in EP V is negligible. By chang-
TPV is amplified by beating it against an allowed am- ing direction of the electric field, the excitation rate can
plitude T0 . Indeed, if the total transition amplitude is be modulated, and the PNC amplitude EP V can be iso-
Ttot = T0 + TPV , then the transition probability (ampli- lated.
tude squared) acquires an interference term T0∗ TPV + c.c. The nuclear spin of 133 Cs is I = 7/2, so each of the
and the experiments extract TPV by measuring this in- S1/2 electronic states is split into F = 3 and F = 4
terference term. hyperfine components. Measuring the transition ampli-
The first APV signal was observed by the Novosibirsk tudes between the different hyperfine states enables one
group in 1978 using the “optical rotation” technique in Bi to separate nuclear spin-dependent and spin-independent
Barkov and Zolotorev (1978). This technique is based on effects. Multiple reversals of the electric field, magnetic
the interference between the APV and the allowed mag- substates and laser polarization are used to further iso-
netic dipole (M1) transition amplitudes. Parity violation late the APV effect. The measured quantity is the ra-
leads to optical activity, i.e., atoms interacting differently tio RStark = Im(EPV )/β for F = 3 → F = 4 and
with left- and right-circularly polarized light. Thereby F = 4 → F = 3 transitions between hyperfine states.
the polarization vector of linearly polarized light is ro- A first measurement of RStark , accurate to 10%, was
tated as the light passes through an atomic vapor. The performed by the Paris group (Bouchiat et al., 1984),
measured quantity, the rotation angle, is proportional to who ultimately reached an accuracy of 2.6% (Lintz et al.,
the ratio of APV and M1 amplitudes. APV was mea- 2007). A series of measurements by the JILA group cul-
sured in optical-rotation experiments with 209 Bi, 209 Pb, minated in a determination of RStark with an accuracy of
and 205 Tl (Edwards et al., 1995; Macpherson et al., 1991; 0.35% (Wood et al., 1997). The JILA measurements also
Meekhof et al., 1993; Phipp et al., 1996; Vetter et al., resolved the difference between RStark (6SF =3 → 7SF =4 )
1995; Warrington et al., 1993). and RStark (6SF =4 → 7SF =3 ), providing the first signa-
An alternative to the optical rotation scheme is the
Stark interference technique (Bouchiat and Bouchiat,
1975), which we illustrate below using a 133 Cs experi-
ment (Wood et al., 1997) as an example. This technique 2 It is conventional to define E

P V parity violating amplitude as the
was used in Cs (Lintz et al., 2007; Wood et al., 1997), Tl transition matrix element Ψf D |Ψi i between the states with
(Conti et al., 1979) and Yb (Tsigutkin et al., 2009) APV the with maximum values of the magnetic quantum numbers m.
26

ture of a nuclear anapole moment. This is discussed fur- overall 0.27% theoretical uncertainty in the structure fac-
ther in Sec. IV.C. tor kPV (Porsev et al., 2009; Porsev et al., 2010). The
The nuclear-spin-independent parity-violating ampli- final value of the extracted QW was in essential agree-
tude is extracted from the measured RStark and β (Ben- ment with the SM value. Recent reevaluation of some
nett and Wieman, 1999): sub-leading correlation contributions to kPV by Dzuba
et al. (2012a) raised the theoretical uncertainty back to
Im(EPV ) = −0.8374(31)exp (21)th × 10−11 a.u. 0.5%, slightly shifting the value of kPV from that of Por-
sev et al. (2009); Porsev et al. (2010), but maintaining
Extraction of the weak charge QW requires calculations
agreement with the SM. One of us, A.D., thinks that
of an atomic structure factor kPV , defined as part of the correction of Dzuba et al. (2012a) may have
EPV = kPV QW . (30) come from the use of many-body intermediate states by
Dzuba et al. (2012a) that is inconsistent with the one
Reaching theoretical accuracy in kPV equal to or bet- employed by Porsev et al. (2009), as the summation over
ter than the experimental accuracy of 0.35% has been a intermediate states while evaluating kPV must be carried
challenging task. In fact, theoretical calculations of kPV out over a complete set and thereby the results of Dzuba
and extraction of the weak charge from the Cs APV ex- et al. (2012a) require revision. This matter remains unre-
periment (Wood et al., 1997) has been a subject of a solved at present but new methods are being developed
controversy and lively activity over the past 15 years. At to address it. The ever-increasing power of computa-
the time of the 1997 APV measurement, the accuracy of tion is anticipated to bring further improvements in the
the theoretical calculations (Blundell et al., 1990; Dzuba atomic-structure analysis.
et al., 1989) was estimated to be 1%. New atomic life-
time and polarizability data reported by 1999 improved
the agreement of theory and experiment and the theoret- 3. Implications for particle physics and the dark sector
ical uncertainty was reduced to 0.4% (Bennett and Wie-
man, 1999). The resulting value of QW differed by 2.5σ Atomic parity violation yields the most accurate up-
from the prediction of the SM (Bennett and Wieman, to-date probe of the low-energy electroweak sector of the
1999). That discrepancy prompted substantial interest SM, playing a unique role complementary to that of high-
in the particle physics community (Casalbuoni et al., energy physics experiments. Figure 6 illustrates the en-
1999; Ramsey-Musolf, 1999; Rosner, 2000, 2002). At ergy dependence (or “running”) of the electroweak in-
the same time, the reduced theoretical uncertainty raised teraction and places APV in the context of other pre-
the question of whether some “small” sub-1% atomic- cision electroweak measurements. The solid curve is
structure effects could be the reason for the discrepancy. the SM prediction for the dependence of sin2 θW on the
Several groups contributed to understanding such small four-momentum transfer Q. At low Q, it describes the
corrections (Derevianko, 2000; Derevianko and Porsev, evolution primarily through quark loops with small lep-
2002; Dzuba et al., 2001b; Johnson et al., 2001; Ko- tonic corrections; the minimum at 100 GeV/c occurs
zlov et al., 2001; Kuchiev and Flambaum, 2002; Milstein when the W+ W− loop starts contributing substantially
et al., 2002, 2003; Sapirstein et al., 2003; Shabaev et al., at Q ∼ 2mW , mW being the mass of W bosons. The
2005; Sushkov, 2001) [reviewed by Derevianko and Por- Cs APV result is placed at Q = 2.4 MeV/c (Bouchiat
sev (2007)]. The dominant corrections were found to be and Piketty, 1983), which is roughly ~/(a0 /Z), where a0
due to the Breit interaction, radiative QED effects, and is the Bohr radius. This relates the momentum to the
the neutron skin correction, which is the difference be- radius of the innermost electron shell of the Cs atom.
tween the well-known proton nuclear distribution and the Together with the results of high-energy collider exper-
relatively poorly known neutron distribution that dom- iments, APV demonstrates the validity of the running
inates HW . This issue is described in Sec. IV.B.4. In of the electroweak force over an energy range spanning
2005, these corrections essentially reconciled APV in Cs five orders of magnitude. An alternative and more de-
with the SM, with theoretical uncertainty standing at tailed plot in a different renormalization scheme can be
0.5%, still larger than the experimental error bar. found in (Olive et al., 2015); this Particle Data Group re-
With the small corrections sorted out, major theo- view also provides further discussion of relevant particle
retical effort turned to more accurate calculation of the physics experiments.
dominant many-body Coulomb correlation contribution The transitions measured in APV studies are typically
to the structure factor kPV (Porsev et al., 2009; Porsev on a 1 eV energy scale, yet the exquisite accuracy of
et al., 2010). State-of-the-art calculations built upon the those measurements and calculations probes minute con-
ab initio relativistic coupled-cluster scheme of Blundell tributions of the sea of virtual particles at a much higher
et al. (1990) and included a large class of higher-order mass scale, including candidates beyond the SM. For ex-
many-body effects. All relevant atomic properties were ample, APV is uniquely sensitive to extra Z [Z 0 ] bosons
reproduced at a level better than 0.3%, leading to an predicted in grand unified theories, technicolor models,
27

0.242 coupled particles and their associated dark forces (Rou-


Dark boson 50 MeV
ven et al., 2013). One such example is the dark photon
0.240
E158 Qweak (Holdom, 1986), discussed in Sec. IX, that is hypoth-
DIS
)(Q 2 )

0.238 esized to be a massive particle which couples to elec-


0.236
tromagnetic currents just like the photon does. In ad-
W

dition, dark Z bosons have been proposed (Davoudiasl


sin 2 (

0.234 et al., 2014) that couple to the weak neutral currents, i.e.,
APV(Cs) Standard LEP
0.232 their interactions are parity violating. In a sense, dark
Model
photons are massive photons, while dark Z bosons are
0.230 SLAC
light versions of Z bosons. In Fig. 6, the yellow-colored
10 3
0.01 0.1 1 10 100 10 3 area represents the limits on dark Z bosons in the model
momentum transfer Q, GeV/c by Davoudiasl et al. (2014); the unique sensitivity and
discovery potential of APV are apparent. We also point
FIG. 6 (Color online) Running of the weak mixing angle
the reader to Bouchiat and Piketty (1983), who consid-
with momentum transfer Q. The solid red curve is the SM
prediction. The Cs APV result is supplemented with data ered APV mediated by a light gauge boson.
from particle physics experiments: E158, Møller [electron- Another novel possibility for probing the dark sector
electron] scattering; Q-weak, PNC electron-proton scattering, with APV experiments is associated with the search for
ν-DIS, deep inelastic scattering; LEP and SLAC results. Area axions and axion-like particles (Roberts et al., 2014a,b;
colored in yellow comes from one of the “new physics” sce- Stadnik and Flambaum, 2014c). The axion [see also
narios (Davoudiasl et al., 2014): a dark boson of mass 50 Secs.VII and IX] is a pseudo-scalar particle introduced
MeV. The colored area is limited by constraints on model pa-
by Peccei and Quinn (1977b) to solve the strong CP
rameters that would explain the discrepancy (Bennett et al.,
2006) between the muon’s experimental anomalous magnetic problem, which is the “unnatural” smallness of the θ̄QCD
moment and the SM prediction. Adopted from (Davoudiasl parameter in the QCD Lagrangian that quantifies the
et al., 2014). amount of CP -violation (Weinberg, 1976), see Sec. V for
more detail. Axions are also viable dark matter candi-
dates (Bertone, 2013). The relevant pseudo-scalar cou-
pling is
SUSY, and string theories (Langacker, 2009). Limits on
their masses set by APV are at the TeV scale (Porsev L0 = iζ1 me φ ēγ5 e + ζ2 (∂µ φ) ēγ µ γ5 e,
et al., 2009), and these were only recently improved upon
by direct searches at the Large Hadron Collider (Olive where ζ1 , ζ2 are the coupling strengths. The spin-0
et al., 2015). Such Z 0 bosons can also mediate new spin- bosonic field
dependent interactions, see Sec. VII.B.5.
φ(r, t) = φ0 cos(ωφ t − k · r + ...) (31)
Low-energy precision measurements are also uniquely
sensitive to possible “dark forces” which are motivated has an amplitude φ0 related to the DM energy den-
by the intriguing possibility of a “dark sector” extension sity and it oscillates at the particle Compton frequency.
to the SM (Andreas, 2012). Dark matter may be a small ωφ = mφ c2 /~ for a particle of mass mφ and ... in
part of the dark sector, or indeed many dark sectors could Eq. (31) stands for an unknown phase. The k-vectors
exist, each with their own forces and constituent parti- follow the virial distribution of DM velocities. This in-
cles. Dark matter may be accompanied by heretofore teraction induces small oscillations in the APV ampli-
unknown gauge bosons (dark force carriers) which can tude at the Compton frequency. A power spectral den-
couple dark matter particles and ordinary particles with sity of the measured time series of APV amplitude would
exceptionally weak couplings. Modern colliders can be exhibit a characteristic peak at the Compton frequency
blind to such new forces, even though the mass of the with a characteristic strongly asymmetric profile derived
dark force carriers can be quite small. This is because by Derevianko (2016). Such proposals are complemen-
the cross-sections of relevant processes for ordinary mat- tary to searches for axion-induced P-conserving M1 tran-
ter are so small that the dark force events are statisti- sitions (Sikivie, 2014).
cally insignificant and are discarded in high-energy ex-
periments.
Light-mass, weakly-coupled dark-sector particles that 4. Isotopic chains and neutron skin
interact with ordinary matter have been proposed as
explanations of astronomical anomalies (Arkani-Hamed All APV studies to date have been conducted with a
et al., 2009b; Fayet, 2004) as well as discrepancies be- single isotope and required the theoretical calculation of
tween the calculated and measured magnetic moment a kPV factor in Eq. (30). Considering challenges faced by
of the muon (Fayet, 2007; Pospelov, 2009). There are such calculations, an alternative approach was proposed
several proposed inroads into the detection of weakly- by Dzuba et al. (1986). The idea was to form a ratio R of
28

the PNC amplitudes for two isotopes of the same element. While the single isotope measurements are sensitive to
Since the factor kPV remains the same, it cancels out in new physics associated with electron-neutron couplings,
the ratio. However, Fortson et al. (1990) pointed out a the isotopic ratios predominantly probe electron-proton
conceptual limitation to this approach – an enhanced sen- (e-p) couplings (Ramsey-Musolf, 1999). Bounds on the
sitivity of possible constraints on “new physics” to uncer- e-p new physics can also be directly established from
tainties in the neutron distributions. This problem is usu- PNC electron scattering off protons in the Q-weak ex-
ally referred to as that of “neutron skin.” The neutron periment at JLAB (Androic et al., 2013). While it was
skin is defined as the difference between the root-mean- previously argued (Derevianko and Porsev, 2002; Fortson
square radii Rn and Rp of neutron and proton distribu- et al., 1990; Ramsey-Musolf, 1999) that APV ratios, due
tions. While nuclear charge densities (i.e., proton dis- to neutron skin uncertainties, are not competitive to such
tributions) have been accurately measured with electron direct experiments, Brown et al. (2009) showed that the
scattering, and the mean-square charge radii are well de- induced neutron skin uncertainties for isotopes are highly
termined from isotope-shift measurements, neutron dis- correlated and tend to strongly cancel while forming R.
tributions, while expected to largely follow the proton This observation makes APV isotopic ratio experiments
distributions, are poorly known (Brown et al., 2009). a competitive tool in probing new physics e-p couplings,
Even in the interpretation of the most accurate to date provided the experiments can reach the required level of
single-isotope measurement in Cs (Wood et al., 1997), accuracy.
neutron skin was a point of concern, as the induced uncer-
tainty was comparable to the experimental uncertainty in
the APV amplitude (Pollock and Welliver, 1999; Vrete- C. Nuclear-spin-dependent effects and the nuclear anapole
moment
nar et al., 2000). The question was addressed in Dere-
vianko and Porsev (2002), where empirical antiprotonic- 1. Overview
atom data fit for the neutron skin was used (Trzcinska
et al., 2001), and the associated uncertainty in the neu- The three nuclear-spin-dependent diagrams, Fig. 5(b-
tron skin contribution to kPV was substantially reduced. d), can be reduced to the effective interaction in the elec-
An analysis for multiple isotopes (Brown et al., 2009) tron sector
shows that in Fr and Ra+ , the present uncertainty in
neutron skin would limit extraction of weak charge to GF
HNSD = √ (ηaxial + ηNAM + ηhf ) (α · I) ρ (r) , (32)
0.2% accuracy. 2
The question of determining neutron skin is of interest where α is the velocity operator (αi = γ0 γ i ) for atomic
in its own right, for example, as it relates to the equa- electrons, ρ is the nuclear density, and I is the nuclear
tion of state for neutron stars. The 208 Pb Radius Exper- spin. This contribution is only present for I 6= 0 isotopes
iment (PREX) at Jefferson Lab (JLAB) (Abrahamyan and open-shell atoms. The dimensionless parameters η
et al., 2012) uses PNC asymmetry in elastic scattering primarily come from nuclear physics. In the ideal nuclear
of electrons from 208 Pb with the goal of measuring Rn shell-model limit these coefficients are associated with
to 1% accuracy. Brown et al. (2009) examined the ques- the properties of the valence nucleon N : N = p or N =
tion of whether neutron skin can be probed with APV. n depending on a specific nucleus. The non-relativistic
The neutron skin correction is about 0.2% for Cs APV reduction of the operator (α · I) ρ (r) in HNSD reads
and 0.6% for Fr and Ra+ . Yb, francium (Fr, Z=87),
and radium (Ra, Z = 88) have a number of isotopes 1
[2ρ(r)(I · p) − i(I · ∇ρ) + σ · (∇ρ × I), ] .
available for APV experiment and highly accurate mea- 2c
surements of APV in two isotopes may, in principle, be The coefficient ηaxial is associated with the Z exchange
used to extract the neutron skin data. In isotopic chain interaction from nucleon axial-vector (An Ve ) currents,
experiments the largest effect is attained for a pair of iso- Fig. 5(b), and its nuclear shell-model value is (Flambaum
topes comprised of the lightest (neutron-depleted) and and Khriplovich, 1980)
the heaviest (neutron-rich) isotopes of the chain. For Yb
the accuracy in the ratio determination should be smaller (2) κN− 1/2
ηaxial = −CN , (33)
than 0.2% [0.3% for Fr and Ra+ ] just to detect the effects I(I + 1)
of having different rms neutron radii for two isotopes. (2)
This may prove more challenging than the single isotope where the weak-interaction constants Cn,p were intro-
approach, unless common systematics in measuring APV duced in Sec. IV.A and
amplitudes in individual isotopes cancel out in the ratio. κN = (I + 1/2)(−1)I+`N +1/2
APV measurements on Yb (Leefer et al., 2014) also bene-
fit from a 100-fold enhancement in EPV compared to Cs; is the relativistic angular quantum number for the un-
the enhancement is due to the presence of closely spaced paired nucleon in a state with orbital angular momen-
opposite parity levels (DeMille, 1995). tum `N . Notice that this contribution is substantially
29

suppressed compared to the Vn Ae diagram 5(a) because that are typically studied with deep inelastic scatter-
(2) ing (PVDIS-Collaboration, 2014). The boundary be-
|CN /Cn(1) | = gA (1 − 4 sin2 θW ) ≈ 0.1 tween the axial- and anapole-dominated regimes depends
and only the unpaired nucleon contributes to Fig. 5(b) on quantum numbers of the valence and type of the va-
(2)
whereas all nucleons coherently contribute to Fig. 5(a). lence nucleon (DeMille et al., 2008a). Values of Cn,p can
The ηNAM coefficient parameterizes the nuclear set constraints on exotic new physics such as leptopho-
anapole moment (NAM) contribution to atomic parity bic Z 0 bosons (Buckley and Ramsey-Musolf, 2012), while
violation. It is illustrated in Fig. 5(c) and discussed NAMs probe hadronic PNC.
in Sec. IV.C.2. Parity violation in the nucleus leads
to toroidal currents that in turn generate a parity-odd,
time-reversal-even (P-odd, T-even) moment, known as 2. Nuclear anapole moments as a probe of hadronic parity
violation
the nuclear anapole moment, that couples electromag-
netically to atomic electrons. The nuclear shell model
The traditional multipolar expansion of electromag-
expression for the anapole moment (Flambaum et al.,
netic potentials generated by a finite distribution of cur-
1984),
rents and charges leads to the identification of mag-
κN netic (MJ) and electric (EJ) multipolar moments (Jack-
ηNAM = 1.15 × 10−3 µN gN A2/3 , (34)
I(I + 1) son, 1999). Non-vanishing nuclear multipolar moments
depends on the atomic number A, the magnetic moment (charge E0, magnetic-dipole M1, electric-quadrupole E2,
µN of the unpaired nucleon expressed in units of the . . . ) respect parity and time reversal, i.e. they are P-even
nuclear magneton, and the weak coupling constant gN . and T-even, and describe multipolar fields outside the fi-
Their values are µp ≈ 2.8, µn ≈ −1.9, gp ≈ 5, and nite distribution. Weak interactions inside the nucleus
gn ≈ −1. lead to additional P-odd moments (Gray et al., 2010);
The combined action of the hyperfine interaction and the leading moment is referred to as the anapole mo-
the spin-independent Z-exchange interaction from nu- ment. Zel’dovich and Vaks were the first to point out
cleon vector (Vn Ae ) currents leads to the third nuclear- the possibility of such a moment (Zel’dovich, 1958).
spin dependent parity violating effect, Fig. 5(d). This The anapole moment a of a current density distribu-
contribution is quantified by a parameter ηhf . An an- tion j(r) is defined as
alytical approximation for ηhf was derived by Flam- Z
baum and Khriplovich (1985b) and values of ηhf were a = −π d3 r r2 j(r), (35)
determined for various cases of experimental interest by
Bouchiat and Piketty (1991) and Johnson et al. (2003). with magnetic vector potential A = aδ(r), leading to
Johnson et al. (2003) also tabulated the values of ηhf the electromagnetic coupling of electrons to the nuclear
for microwave transitions between ground-state hyper- anapole moment, (α · A). A classical analog of the
fine levels in atoms of potential experimental interest. anapole moment is a Tokamak-like configuration shown
Recently, Flambaum (2016) pointed out a novel nu- in Fig. 7. The inner and outer parts of the toroidal cur-
clear spin-dependent effect: the quadrupole moment of rents are weighted differently by r2 in Eq. (35), leading
the neutron distribution leads to a tensor weak interac- to a nonvanishing value of the anapole moment. Mi-
tion that mixes opposite parity states in atoms with total croscopically, a nuclear anapole moment can be related
angular momentum difference ≤ 2. This effect should be to a chiral distribution of nuclear magnetization caused
carefully investigated in future work to see if it influences by parity-violating nuclear forces (Bouchiat and Piketty,
determination of the anapole moments from APV mea- 1991). Due to the Wigner-Eckart theorem, the NAM
surements. The effect is of interest on its own as a probe (just as the nuclear magnetic moment) is proportional to
of the neutron distributions in nuclei (Flambaum et al., the nuclear spin I so that
2017). The atom or molecule should contain a nucleus
with I > 1/2, and there is an enhancement for heavy and GF
deformed nuclei. a= √ ηNAM I,
|e| 2
An outstanding question is the relative importance
of the nuclear spin-dependent contributions. The ηhf defining the constant ηNAM in Eq. (32). Atomic electrons
coefficient can be carefully evaluated and it is usually interact with NAM only inside the nucleus, as is appar-
suppressed compared to ηNAM and ηaxial . Generically, ent from the classical analog, since the magnetic field is
because of the A2/3 scaling, the anapole contribution entirely confined inside the “doughnut”. Another impor-
dominates for heavier nuclei. For lighter nuclei, the tant observation is that the NAM is proportional to the
axial contribution is more important and APV experi- area of the toroidal winding, i.e., ∝ (nuclear radius)2 ∝
(2)
ments can be a sensitive probe of Cn,p electroweak pa- A2/3 , where A is the atomic number, illustrating the
rameters, providing a window on the An Ve interactions trend in Eq. (34).
30

FIG. 7 (Color online) The toroidal component of current den-


sity j produces anapole moment a, with magnetic field B that
is entirely confined inside the “doughnut”. The azimuthal
component of current density generates magnetic dipole mo-
ment aligned with a, with its associated conventional dipolar
magnetic field not shown.

FIG. 8 (Color online) Constraints on combinations of par-


Microscopically, the nuclear anapole arises due to ity violating meson couplings (×107 ) derived from Cs anapole
nucleon-nucleon interaction, mediated by meson ex- moment (yellow band) and nuclear experiments. Bands have
change, where one of the nucleon-meson vertexes is a width of one standard deviation. Best value predicted by
strong and another is weak and P-violating. Thus, the DDH analysis is also shown. This figure combines Cs
determination of anapole moments from atomic parity NAM band from (Haxton and Wieman, 2001) with more re-
violation provides an important window into hadronic cent nuclear-physics constraints figure from (Haxton and Hol-
PNC (Haxton and Wieman, 2001). The innards of stein, 2013).
the anapole bubble in Fig. 5(c) are shown in Fig. 7
of the review by Haxton and Wieman (2001). The
nuclear-physics approach is to characterize weak meson- area colored red lies at the intersection of nuclear ex-
nucleon couplings in terms of parameters of Desplan- perimental bands. There is some tension with the Cs
ques, Donoghue and Holstein (DDH) (Desplanques et al., anapole moment result, although the Cs result is consis-
1980), who deduced SM estimates of their values. These tent with “reasonable ranges” of the DDH parameters.
six hadronic PNC parameters are fπ , hρ0,1,2 , hω 0,1
, where Haxton and Wieman (2001) point out that additional
the subscript (π, ρ, ω) indicates meson type and the su- APV experiments with unpaired-neutron nuclei would
perscript stands for isoscalar (0), isovector (1), or isoten- produce a band perpendicular to the Cs band (the 133 Cs
sor (2). We refer the reader to Haxton and Wieman anapole moment is primarily due to a valence proton).
(2001) for a detailed review of nuclear structure cal- This provides strong motivation for the ongoing exper-
culations of NAMs within the DDH parameterization. iments to measure nuclear-spin-dependent APV effects
The effective field theory parameterizations of hadronic in nuclei with unpaired neutrons such as 171 Yb (Leefer
PNC, an alternative to DDH, are also discussed (Ramsey- et al., 2014), 212 Fr (Aubin et al., 2013), and 137 Ba (De-
Musolf and Page, 2006), although NAM analysis in this Mille et al., 2008a).
framework remains to be carried out. It should be
pointed out that a more recent review (Haxton and Hol-
stein, 2013) omits the Cs result. These authors explain D. New and ongoing APV experiments
the omission by the fact that the accuracy of the con-
straints on the nucleon-nucleon PNC interaction derived We limit our discussion to APV experiments that are
from the NAM experiments is somewhat difficult to as- now being actively pursued. We refer the reader to the
sess due to complex nuclear polarizability issues. earlier reviews (Bouchiat and Bouchiat, 1997; Budker,
The derived bounds (Haxton and Wieman, 2001; Hax- 1999; Ginges and Flambaum, 2004; Roberts et al., 2015)
ton and Holstein, 2013) on PNC meson couplings are for a discussion of various proposals.
shown in Fig. 8. The 133 Cs APV result is shown in addi- Experimental efforts to improve the accuracy of PNC
tion to constraints from scattering of polarized protons on measurements in Cs are underway at Purdue Univer-
unpolarized proton and 4 He targets and emission of cir- sity (Antypas and Elliott, 2014). This group is exploring
cularly polarized photons from 18 F and 19 F nuclei. The a new two-pathway coherent control technique. Here, two
31

optical excitations, starting from the same initial state fine transitions. DeMille et al. (2008a) outline a Stark-
(6S1/2 ) and leading to the same final (7S1/2 ) state, are interference technique to measure spin-dependent APV
driven by two different mutually-coherent fields. One of effects to determine the mixing between opposite-parity
the lasers is resonant with the 6S1/2 − 7S1/2 transition rotational/hyperfine levels of ground-state molecules. By
and the other operates at half the resonant frequency using a magnetic field to tune these levels to near-
driving an allowed two-photon E1 amplitude. The ab- degeneracy, the usual PNC-induced mixing is dramati-
sorption rate contains an interference term between the cally amplified (Kozlov et al., 1991). This method can in
two-photon amplitude and a sum of Stark-induced and principle give a large enhancement in sensitivity relative
PNC amplitudes, and it depends on the relative phase of to traditional experiments with atoms. The technique is
the applied laser fields. By experimentally varying the applicable to nuclei over a wide range of atomic num-
relative phase one would observe oscillating modulation bers in diatomic species that are theoretically tractable.
(2)
of the transition rate. As a demonstration, the Purdue Both NAMs and Cn,p electroweak parameters, discussed
group has measured several atomic properties of Cs (An- in Sec. IV.C, can be probed. Such experiments are un-
typas and Elliott, 2011; Antypas and Elliott, 2013a,b). derway at Yale.
Francium and the Ra+ ion have an electronic atomic While PNC interactions do not normally cause first-
structure similar to Cs, but larger nuclear charge Z and order energy shifts because they mix states of opposite
thereby larger PNC amplitude due to the Z 3 enhance- parity, such energy shifts do occur in chiral systems.
ment. Both atoms are amenable to the application of the This fact has been recognized since 1970s (Letokhov,
same theoretical techniques as Cs (Dzuba et al., 2001a, 1975), and searches for minute PNC energy shifts be-
1995; Pal et al., 2009; Safronova and Johnson, 2000; Sa- tween states of chiral enantiomers (molecules that are
hoo, 2010; Wansbeek et al., 2008) and potentially offer mirror images of one another) via high-resolution spec-
improved probes of the low-energy electroweak sector. troscopy have been ongoing ever since then [see, for ex-
The experimental challenge with these systems lies in ample, (Tokunaga et al., 2013) and references therein].
their radioactivity which requires special experimental So far there have been no conclusive observations of a
facilities. A Fr experiment is in preparatory stages at parity violating effect in chiral molecules. Eills et al.
the TRIUMF facility in Vancouver (Aubin et al., 2013), (2017) proposed a new experiment to search for PNC in
while Ra+ ion is investigated in Groningen (Gomez et al., chemical shifts of chiral molecules using nuclear magnetic
2006; Nuñez Portela et al., 2013). Ra+ is an ion and re- resonance (NMR) spectroscopy. A proof-of-principle ex-
quires application of novel experimental techniques (Fort- periment with 13 C-containing molecules was presented,
son, 1993). with molecules containing heavier nuclei with enhanced
Since the accuracy of atomic calculations for multiva- PNC effects to be used next. Precision measurements
lent systems is unlikely to approach that achieved for of this kind may be useful for studying nuclear PNC and
atoms with a single valence electron [Cs, Fr, Ra+ ], the testing exotic physics models that predict the presence of
strategy for ongoing experiments in Yb is to pursue iso- parity-violating cosmic fields (Roberts et al., 2014a,b).
topic ratios, as discussed in Sec. IV.B.4. One of the most
immediate goals of Yb APV experiments (Leefer et al.,
2014) is verification of the isotopic dependence of the V. TIME-REVERSAL VIOLATION: ELECTRIC DIPOLE
weak charge, with the Yb experiment (recently moved MOMENTS AND RELATED PHENOMENA
from Berkeley to Mainz) currently taking data. Experi-
ments with Dy, where there are nearly-degenerate states A. Introduction
of opposite parity, have not yet detected APV (Nguyen
et al., 1997); however, this is expected in the new gener- In this section, we review phenomena related to simul-
ation of the apparatus (Leefer et al., 2014). taneous time-reversal- (T -) and parity- (P -) violation in
While Cs is the only experiment to date that has mea- atomic and molecular physics. As we will describe, re-
sured NAM (Wood et al., 1997), there are several propos- cent searches for T -, P -violating (T,PV) effects in these
als on NAM detection in atomic and molecular exper- systems are probing energy scales well above 1 TeV in
iments. Bouchiat (2007) discusses a NAM-induced lin- particle theory models widely considered as natural ex-
ear dc Stark shift of the individual substates of an alkali tensions to the SM. Clear prospects for future improve-
atom in its ground state, dressed by a circularly polarized ments make it likely that work in this area will remain
laser field. Choi and Elliott (2016) propose an application at the forefront of particle physics for some time.
of the two-pathway coherent control technique for direct A relevant example of a T,PV effect is when a particle
measurement of the anapole moment using the ground- has an electric dipole moment (EDM), d, along its spin
state hyperfine splitting of Cs. Measurements in a chain s, i.e., d = d s/s (Fig. 9). The idea that elementary
of Fr isotopes (Aubin et al., 2013; Gomez et al., 2007) particles might possess a permanent electric dipole mo-
are being actively explored, with future plans for APV ment (EDM) in addition to a magnetic dipole moment
measurements using both 7S1/2 −8S1/2 and 7S1/2 hyper- was proposed by Purcell and Ramsey (1950). This leads
32

The time-reversal operator T is anti-unitary: it can be


represented as the product of a unitary operator and the
complex conjugation operator (Sakurai and Napolitano,
2011). Hence while quantum wave equations with real-
valued potentials are T -invariant (i.e., if some wavefunc-
tion Ψ(t) is a solution, then Ψ∗ (−t) is also a solution),
T -violating effects arise for complex-valued potentials.
Hence, T - (and CP -) violation is associated with the ir-
reducible presence of complex numbers in the underlying
theory. The strength of CPV interactions is proportional
to sin φCP , where φCP is the phase of such a complex
number (Fortson et al., 2003). It is known that CPV oc-
curs in nature, from observations of CPV in decays of K 0
and B 0 mesons (Patrignani, 2016). These observations
are all consistent with a single source of CPV in the SM:
FIG. 9 Basic concept of EDM measurements. When a par-
a complex phase in the Cabibbo – Kobayashi – Maskawa
ticle has an EDM d along its spin axis s, an electric field E
causes s to precess about E. (CKM) matrix that describes the mixing between quark
flavors to form mass eigenstates. The measured value of
this phase is large: δCKM ∼ 1 rad. However, the linkage
to an interaction with an electric field E described by the with flavor mixing causes the observable effects of CPV
Hamiltonian to be systematically small in the SM (Bernreuther and
Suzuki, 1991; Khriplovich and Zhitnitsky, 1982). In par-
HEDM = −d · E ∝ d s · E. (36) ticular, EDMs within the SM are exceptionally small, de-
spite the large value of δCKM . By contrast, theories that
HEDM is odd under both T and P : s changes sign under
extend the SM naturally can, and frequently do, include
T but E does not, while under P , E changes sign but
new CPV phases that contribute to EDMs and related
the axial vector s does not. Most T,PV effects in atomic
phenomena in a more direct way, with no obvious mech-
and molecular systems result in an EDM or some closely
anisms for suppression (Barr, 1993). This makes EDMs
related quantity (for example, an interaction between a
a nearly background-free signal for detecting new physics
spin and the internuclear axis in a molecule).3
associated with CPV (Engel et al., 2013; Pospelov and
In relativistic quantum field theories, the combined
Ritz, 2005).
symmetry CP T (where C is charge conjugation) is al-
ways conserved (Streater and Wightman, 2000). More-
over, CP T conservation has been experimentally con-
firmed to extraordinary precision (see Sec. V). Hence, in
typical theoretical extensions to the SM, it is assumed
that T -violation is equivalent to CP -violation (CPV),
and for the remainder of this section we do so as well. Moreover, there must be new sources of CPV in na-
Based on very general considerations in quantum field ture. This conclusion arises from the observation of a
theory, at low energies the largest effects of CPV are baryon asymmetry—a cosmological imbalance between
expected to appear as T,PV interactions rather than T - matter and antimatter (Dine and Kusenko, 2003). Since
violating but P -conserving (TV) signals (Khriplovich and matter-antimatter annihilations in the aftermath of the
Lamoureaux, 1997). In fact, limits on T,PV effects in Big Bang produce photons, this asymmetry is typically
combination with established principles of field theory parameterized by the cosmological baryon-to-photon ra-
rule out TV effects (Conti and Khriplovich, 1992) far tio η ∼ 10−10 . Sakharov showed that, among other con-
below the level of any conceived experiment to detect ditions, CPV is necessary to generate this asymmetry
them, see Hopkinson and Baird (2002); Kozlov and Por- (Sakharov, 1967). While the SM in principle incorpo-
sev (1989). Hence for the remainder of this section we rates all the Sakharov conditions, the size of CPV effects
use the terms CPV and T,PV interchangeably. in the SM is far too small to account for the observed
value of η (Gavela et al., 1994). By contrast, theoret-
ical models containing new particles with masses near
the electroweak scale MZ ∼ 100 GeV, together with new
3 In this section, expressions related to electomagnetism will use CPV phases, could explain the experimentally observed
the cgs system of units so that electric and magnetic fields, as
well as electric and magnetic dipole moments, have the same
value of η. In these scenarios—known as electroweak
dimensions. For many expressions we use mixed units that are baryogenesis— CPV (or, equivalently, T,PV) signals typ-
standard for the field, such as electric fields in units V/cm and ically are predicted to appear at a level near current ex-
electric dipole moments in units e · cm. perimental sensitivities (Engel et al., 2013).
33

B. Observable effects in atoms and molecules Hence, in this weakly-polarized regime it is sensible to
say that HTP induces a dipole moment (of opposite sign
Atomic and molecular experiments searching for T,PV in the upper/lower states of the system), and the en-
can be broadly classified into two categories, based ergy shifts of interest are proportional to the strength of
on whether the system is paramagnetic—unpaired elec- the applied field, E. Next, consider the strong-field case,
±
tron spins—or diamagnetic—closed electron shells, but where 1 − P  1. In this regime, ∆ETP ≈ ∓δTP sgn(m).
nonzero nuclear spin (Barr, 1993; Khriplovich and Lam- Here the T,PV energy shifts are independent of E, and it
oureaux, 1997). Paramagnetic systems are most sensi- is no longer sensible to speak of a T,PV dipole moment
tive to effects that depend explicitly on electron spin: of the system. The shifts are also maximal in this regime.
the electron EDM (eEDM) and one type of semileptonic It is infeasible to reach the strong-field regime in the
(electron-nucleus) interaction. Diamagnetic systems are ground states of atoms: for typical splittings between
most sensitive to effects that depend explicitly on nu- opposite-parity levels ∆ ∼ Eh (the atomic unit of en-
clear spin: purely hadronic T,PV interactions, EDMs of ergy, e2 /a0 ) and dipole matrix elements D ∼ ea0 , the re-
nucleons, and other types of semileptonic interactions. quired field strength E > 2
∼ Eat , where Eat = e/(4π0 a0 ) ∼
To understand how T,PV effects are related to EDMs 9
5 × 10 V/cm is the atomic unit of electric field, is
in atoms and molecules, consider a toy system consisting far too large to apply in the lab. However, in polar
of two states with opposite parity eigenvalues Π = ±1, molecules there are levels of opposite parity with much
each with angular momentum j, split in energy by ∆, smaller energy splittings but similar dipole matrix ele-
and in particular a pair of substates |j, m, Πi with the ments, making it far easier to polarize these systems.
same projection m. These states can be mixed by a T - Such pairs of levels—associated with rotational structure
,P -odd Hamiltonian HTP ; since this is a rank-0 tensor, (where ∆ ∼ [me /mp ]Eh ) or Ω-doublet structure (where
its matrix elements δTP are independent of m. The levels ∆ ∼ [me /mp ]n Eh , with n = 1 or 2 depending on the
can also be mixed by the Stark Hamiltonian type of electronic state)—make it routine to reach the
regime of nearly full polarization in these systems (San-
HSt = −D · E ẑ, (37) dars, 1967; Sushkov and Flambaum, 1978). The increase
in observable T,PV energy shifts, relative to the case of
with electric dipole matrix element
atoms in lab-scale E-fields, is typically 3-5 orders of mag-
hj, m, +|Dz |j, m, −i ≡ Dsgn(m), nitude. Hence, experiments with molecules play an im-
portant role in this field (Kozlov and Labzowsky, 1995).
where the m-dependence follows from the Wigner-Eckart
theorem. This system is described by the Hamiltonian
  C. Underlying physical mechanisms for T,PV
−∆/2 −DEsgn(m) + δTP
Htoy = 1. Semileptonic interactions
−DEsgn(m) + δTP ∆/2
(38)
In addition to the usual Stark shifts there are T,PV shifts, Semileptonic interactions (SLIs) arise in several
given to 1st order in δTP by particle-theory models. They can be described as a 4-
fermion interaction, related to the exchange of a heavy
±
∆ETP = ∓PδTP sgn(m). (39) force-carrying boson between electrons and the nucleus.
Effects due to exchange of lighter force-carriers are dis-
Here, the dimensionless quantity cussed in Sec. VII.C. A few distinct forms of interac-
DE tion give nonzero effects (Khriplovich and Lamoureaux,
P≡p (40) 1997). The first is the coupling of a scalar current from
(∆/2)2 + D2 E 2 nucleons n in the nucleus to a pseudoscalar electron cur-
(with values 0 ≤ P < 1) quantifies the polarization of rent, described by the relativistic Lagrangian density
the system. The superscript ± refers to the upper/lower
state of the system. X
LSP ∝ ψ̄e iγ 5 ψe ψ̄n ψn . (41)
This has simple behavior in the limiting cases where n
DE  ∆ or  ∆. Consider the weak-field limit, where
P  1; then This yields a relativistic Hamiltonian for the interaction
of a single electron with a pointlike nucleus,
±
∆ETP ≈ ∓(2DE/∆)δTP sgn(m).
rel GF 1
HSP = i√ QSP δ 3 (r)γ 0 γ 5 . (42)
This is exactly the shift that would be found for a system 2 2me c
with permanent dipole moment d± = ±dj/j, where
Here QSP is the effective charge of the nucleus for the
d = 2DδTP /∆. scalar-pseudoscalar interaction, analogous to the weak
34

charge QW for the P V weak interaction. This is fre- hHEDM i = −d · hE tot i = 0. The physical meaning of
quently written in the form QSP = ACS , where A is the this result, known as Schiff’s theorem, is that other parts
mass number and CS is the average effective charge per of the system rearrange so as to completely screen the
nucleon. In the nonrelativistic (n.r.) limit, this Hamilto- external E-field felt by the charged particle; otherwise,
nian takes the form it would undergo a net acceleration. Mechanisms for
evading Schiff’s theorem are thus central to experiments
nr GF 1
HSP = i√ ACS {{s · p, δ 3 (r)}, (43) searching for constituent particle EDMs in atoms and
2 2me c~ molecules.
where {} denotes the anticommutator. This has the same
form as the P -odd (but T -even) Hamiltonian arising from
3. Electron EDM
Z 0 -boson exchange, aside from the factor of i. Due to the
contact nature of the interaction, H SP mixes only s1/2 First, we consider the eEDM in a paramagnetic atom.
and p1/2 orbitals in atoms, with typical matrix element Remarkably, the relativistic motion of the bound electron
δSP = hs1/2 |HSP |p1/2 i can lead to energy shifts orders of magnitude larger than
the shift for a free electron, ∆ETP = −de · E. This
~
∼ CS AZ 2 GF ∼ 10−16 × CS AZ 2 Eh . enhancement, first recognized by Sandars (1965), makes
me ca40 atomic and molecular experiments particularly sensitive
The explicit dependence of HSPnr
on s shows that, to lowest to the eEDM. We discuss the underlying mechanism here.
order, the effect of HSP is nonzero only for paramagnetic The relativistic Lagrangian density associated with the
systems; in diamagnetic systems, hyperfine-induced mix- interaction between the eEDM, de , and an electromag-
ing leads to a nonzero effect at higher order (Flambaum netic field, described by the field tensor F µν , is
and Khriplovich, 1985a). de
Other forms of SLIs lead to Lagrangian densities with LeEDM = −i Ψσ µν γ 5 ΨFµν , (44)
2
the form of a pseudoscalar nucleon-scalar electron cur-
rent or a tensor-tensor interaction, which give rise to where Ψ is the Dirac bispinor for the electron and σ µν =
i µ ν ν µ
Hamiltonians that depend on the nuclear spin I in the 2 (γ γ − γ γ ). This yields the single-electron rela-
rel
system. However, the effects of these interactions are tivistic Hamiltonian HeEDM :
usually strongly suppressed, either in the underlying par- rel
HeEDM = −de γ 0 Σ · E, (45)
ticle theory models or at the atomic/nuclear level. We
refer the reader to Khriplovich and Lamoureaux (1997) where Σ is a Dirac spin operator. From Schiff’s theorem,
for more details. on application of an external field E the n.r. version of
this Hamiltonian (still expressed in terms of bispinors),
−de Σ·E tot , will yield a vanishing energy shift. Hence, we
2. EDMs of constituent particles: Schiff’s theorem
may subtract this term away to find an effective Hamil-
tonian that will account for any observable energy shift
We next turn to the question of how EDMs of con-
due to the eEDM:
stituent particles in an atom or molecule—electrons or
rel,eff
nuclei—can lead to observable T,PV. The answer is sub- HeEDM = −de (γ 0 − 1)Σ · E tot . (46)
tle. Schiff (1963) showed that under reasonable first-
order assumptions—i.e., non-relativistic point particles In the n.r. limit, this takes the form
moving in a purely electrostatic potential—there is no de
nr,eff
(s · p)(s · E tot )(s · p) .
 
energy shift when an E-field is applied to a neutral sys- HeEDM =4 (47)
me c2 ~3
2
tem built from such consitutents. The proof is simple.
The total electric field E tot experienced by the particle nr,eff
The matrix elements of HeEDM between atomic s and p
of interest—which is the sum of an externally applied orbitals are
field E and the internal field E int due to other particles
in the system—can be expressed as E tot = −∇Φ, where δeEDM ∼ de (Z 3 α2 )Eat
Φ is an electrostatic potential. The Hamiltonian for the
particle of charge q and mass m, neglecting the EDM, is (Khriplovich and Lamoureaux, 1997; Sandars, 1966). On
application of a polarizing external field E, this gives rise
H0 = p2 /(2m) + qΦ. to energy shifts
±
Since p = −i~∇, E tot ∝ [p, H0 ]. Thus, for any eigen- ∆EeEDM = ∓PδeEDM .
state |ψi of H0 , the expectation value of the total E- For a fully polarized system, we can write
field vanishes: hE tot i = 0. Hence, the energy shift
±
due to the constituent particle’s EDM d also vanishes: ∆EeEDM = −de E eff ,
35

where distribution that gives a constant electric field E S k I


within the volume of the nucleus (Flambaum and Gin-
E eff ∼ ±(Z 3 α2 )Ermat . ges, 2002); it has dimensions of [charge·volume].
This effective electric field can be orders of magnitude This yields a term HS in the n.r. atomic/molecular
larger than the applied field E: for Z ≈ 90, typically Hamiltonian that, for a spherical nucleus of radius RN ,
E eff ∼ 100 GV/cm. For a weakly-polarized system, has the form
±
∆EeEDM = ∓2DEδeEDM /∆ can be written as 15e 5
±
HS = − (S/RN )r · I/I (r < RN ). (48)
∆EeEDM = ∓de Fe (Z)E, 4π0

where the quantity Fe (Z) is referred to as the eEDM This interaction gives first-order effects in both diamag-
enhancement factor for atoms: it describes the factor by netic and paramagnetic systems. The associated T,PV
which, in the limit of weak polarization, ∆EeEDM exceeds atomic/molecular matrix elements have typical size δS ≡
the shift for a free electron. With D ∼ ea0 and ∆ ∼ Eh , hp|HS |si ∼ Z 2 S/(ea30 )Eh (Khriplovich and Lamoureaux,
Fe (Z) ∼ 2Z 3 α2 , with the typical values Fe ≈ 100 − 600 1997).
for Z ≈ 55 − 80. A nuclear SMt can be induced by a variety of mi-
The evasion of Schiff’s theorem here is remarkable, croscopic physics effects. An example is when the nu-
since even in the relativistic case the expectation value cleus contains a valence nucleon n with dipole moment
of E tot vanishes. The nonzero effect can be understood dn . In a nuclear shell model where n moves around a
heuristically as arising from the relativistic length con- uniform spherical core of radius RN = R0 A1/3 (where
traction of the eEDM, acting in concert with the spa- A is the nuclear mass number and R0 ≈ 1.2 fm is
tial variation of the Coulomb field E int (Commins et al., the characteristic nuclear size), the SMt has magnitude
2007). Since neither the length-contracted dipole mo- S ∼ 0.1dn A2/3 R02 . In the weak polarization limit, HS
ment drele nor the electric field E tot = E + E int are induces an atomic/molecular EDM da . Since S ∝ dn ,
constants over the atomic volume, it makes sense that the quantity Fn = da /dn is analogous to the eEDM en-
tot tot hancement factor Fe for the eEDM. However, here there
h−drel
e ·E 6 0 = hdrel
i= e i · hE i.
is instead a suppression: Fn ∼ Z 2 A2/3 R02 /a20 , with typi-
cal numerical value Fn ≈ 10−3 for Z = 80 (Khriplovich
4. Hadronic T,PV: nuclear Schiff moment and related effects and Lamoureaux, 1997).
In most theoretical models, T -,P -odd intranuclear in-
Much like how T -,P -odd SLIs and/or the eEDM can teractions, rather than the nucleon EDM, give dominant
induce an atomic EDM, the presence of a proton or neu- contributions to the nuclear SMt (Sushkov et al., 1984).
tron EDM, or of T -,P -odd hadronic interactions, can mix For example, in many theories quarks acquire a chromo-
nuclear states to induce a nuclear EDM. However, within EDM (cEDM), d˜q , which is the strong-interaction ana-
an atom the motion of a nucleus is deeply nonrelativistic. logue of the ordinary EDM. The color field resulting
Hence, according to Schiff’s theorem, any nuclear EDM from the cEDM induces a T -,P -odd strong interaction—
is very effectively screened from external fields and leads typically described as an effective T,PV nucleon-nucleon
to negligible energy shifts. Nevertheless, the same T -,P - interaction—between a valence nucleon and the remain-
odd hadronic effects can induce changes in the nuclear der of the nucleus. This mechanism generally leads to a
charge and current distributions corresponding to elec- larger nuclear SMt than that from the ordinary nucleon
tromagnetic moments other than an EDM. These modi- EDM (Fischler et al., 1992), by a factor of ∼ 40, for the
fied distributions, unlike a nuclear EDM, can give rise to same size of d˜q and dn (Khriplovich and Lamoureaux,
T,PV energy shifts in an electrically-polarized atom or 1997).4 Hence these experiments are particlularly sensi-
molecule. tive to new physics at high energy scales that is related
The primary mechanism for these shifts is associated to quark cEDMs (Engel et al., 2013; Pospelov and Ritz,
with the finite size of the nucleus. Penetration of va- 2005).
lence electrons into the finite nuclear volume allows them In addition, there is the possibility in quantum chro-
to interact with a local (intra-nuclear) E-field different modynamics (QCD) of an irreducible CPV interaction
from that of a point dipole, which would be completely (see, for example, the reviews by Peccei (2008), Kim and
screened. The charge distribution that leads to the Carosi (2010), and Sikivie 2012), described by the La-
lowest-order observable T,PV energy shift is known as grangian density
the Schiff moment (SMt), S = SI/I (Schiff, 1963).
Z Z Z  Lθ = −θ̄QCD (αs /8π)(µνκλ /2)Gaµν Gaκλ .
Ze 2 3 5 3 2 3
S≡ ρZ (r)rr d r − ρZ (r)rd r ρZ (r)r d r ,
10 3

Rwhere ρZ3 is the nuclear charge density normalized as


ρZ (r)d r = 1. Physically, S corresponds to the charge
4 The color field within the nucleon from d˜q induces dn ∼ ed˜q .
36

Here, αs is the strong-interaction analogue of α in electro- Experiments of this type contend with certain common
magnetism; µνκλ is the completely antisymmetric ten- issues related to the fact that mj -dependent energy shifts
sor; Ga is the gluon field tensor; and θ̄QCD is a dimension- easily can be caused not only by T,PV effects, but also by
less constant parameterizing the strength of this term rel- magnetic fields B due to their interaction with the mag-
ative to the ordinary strong interaction. Lθ also leads to netic moment µ ∝ j. Random B-field fluctuations can
an effective T,PV nucleon-nucleon interaction (Crewther degrade the signal-to-noise ratio. Nearly all experiments
et al., 1979; Engel et al., 2013; Pospelov and Ritz, 2005). minimize this effect by reversing E, in order to reverse
Typical calculations in spherical nuclei of the relation be- the system polarization P, as frequently as possible. It
tween θ̄QCD and S yield S ∼ 10−3 θ̄QCD eR03 (Khriplovich is also common to perform measurements on side-by-side
and Lamoureaux, 1997). Searches for nuclear SMts (and regions with opposing E-fields; common-mode magnetic
the bare neutron’s EDM (Baker et al., 2006)) set a strong field shifts cancel in the difference between energy shifts
bound θ̄QCD < −10
∼ 10 , while naively one expects dimen- in these regions. In addition, B-fields correlated with E
sionless fundamental parameters to have values of or- can lead to systematic errors that mimic ∆ETP . These
der unity. The hypothetical particle known as the axion can arise due to leakage currents associated with the E-
was first devised as a mechanism to solve this so-called field and due to motional effects (since a particle mov-
“strong CP problem.” Axions are discussed further in ing in an electric field E with velocity v experiences a
Secs. VII.A.2 and Sec. IX. magnetic field Bmot = E × v). Frequently, experiments
While the SMt makes the dominant contribution to replicate the measurements on an EDM-insensitive sys-
T,PV energy shifts in diamagnetic systems, in para- tem [e.g., a lighter species as in (Regan et al., 2002)] or on
magnetic systems with nuclear spin I ≥ 1, an- a state of the same system with opposite sign of P [e.g.,
other mechanism typically leads to larger nuclear spin- the excited state of the pair in the toy model of section
dependent T,PV effects (Khriplovich and Lamoureaux, V.B, as first used in (DeMille et al., 2001; Eckel et al.,
1997; Sushkov et al., 1984). Here, the underlying 2013)] and hence also ∆ETP . These “co-magnetometers”
hadronic T,PV physics leads to a current distribution in act as a useful probe for systematic errors.
the nucleus that corresponds to a magnetic quadrupole
moment (MQM). This MQM couples to the gradient
of the magnetic field produced by the electron and 2. Experiments on paramagnetic systems
mixes s1/2 and p3/2 atomic orbitals. The nuclear spin-
dependent T,PV energy shifts associated with the MQM The ACME collaboration (Yale/Harvard) recently
can exceed those due to the nuclear SMt by 10 − 100×. completed the most sensitive experiment using a para-
magnetic system (Baron et al., 2014). In ACME, ThO
molecules are prepared in a metastable triplet state with
D. State-of-the-art experiments two valence electrons. In this state (labeled H 3 ∆1 ), one
electron is in a σ1/2 orbital—roughly, a linear combina-
1. General remarks tion of s1/2 and p1/2 atomic Th orbitals—and provides
excellent sensitivity to T,PV effects. The second electron,
All recent atomic and molecular experiments that have in a δ3/2 orbital, nearly cancels the magnetic moment of
set stringent limits on T,PV effects rely on the same basic the first electron, and also gives rise to high polarizabil-
measurement principle. A T -,P -odd Hamiltonian HTP , ity due to a small Ω-doublet splitting (Meyer and Bohn,
together with an applied electric field E = E ẑ, results 2008; Meyer et al., 2006; Vutha et al., 2010). In the ex-
in an energy shift ∆ETP of the state |j, mj i, given by periment (see Fig. 10), a beam of ThO molecules is pro-
∆ETP = PδTP sgn(mj ). Since δTP grows rapidly with Z, duced with a cryogenic source that yields, relative to con-
all experiments use heavy atoms. To measure ∆ETP , an ventional molecular beam sources, a low forward velocity,
equal superposition of states with ±mj is prepared and low internal temperature, and high flux. The sequence
allowed to freely evolve for time τ . This state is typi- of events experienced by molecules in the beam proceeds
cally prepared with high efficiency by optical pumping, as follows. First, a set of “rotational cooling” lasers opti-
sometimes in combination with radiofrequency spin-flips. cally pumps ground-state ThO molecules to accumulate
The energy splitting leads to a relative phase accumu- population in a single rotational level. Next, they en-
lated between the states, φ = 2∆ETP τ /~. The superpo- ter a magnetically-shielded interaction region where an
sition state corresponds to an orientation or alignment of electric field E with magnitude E ∼ 100 V/cm is ap-
j in the x − y plane, and the phase evolution is equiv- plied to achieve polarization P ∼ = 1. Once in this region,
alent to a precession of j about E by angle φ. For a a laser pumps population from the enhanced rotational
single particle, ∆ETP can be measured with minimum level into the H state. Next, another laser is used to spin-
uncertainty ~/(4τ ); hence with N uncorrelated particles, align the H-state molecules in a direction perpendicular
the T,PV energy shift
√ can be measured with uncertainty to E, after which they fly freely for a distance of ≈ 20
δ(∆ETP ) = ~/(4τ N ). cm. In the slow molecular beam, this corresponds to spin
37

evolution time τ ≈ 1 ms, comparable to the metastable


level’s lifetime, τH ≈ 2 ms. A magnetic field B is ap-
plied parallel to E to provide a bias (typically π/4 rad)
to the spin precession. After the free-flight region, the
final direction of the spin alignment axis is detected by
the relative strength of laser-induced fluorescence when
molecules are excited by a laser beam with alternating
orthogonal polarizations. To suppress a wide range of
systematic errors, the measurement is performed in both
the positively- and negatively-polarized states of the Ω-
doublet and at different magnitudes of the applied field
FIG. 10 Schematic of the ACME eEDM experiment. The
E. With a rate dN/dt ∼ 5 × 104 /s of detected molecules figure shows only the magnetically-shielded region where spin
and ∼ 2 weeks of data, ACME was sensitive to an energy precession takes place. Reproduced from (Baron et al., 2014).
shift ∆ETP /h < 2 mHz. Given the calculated sensitiv-
ity of the ThO H state to the eEDM (with E eff ≈ 80
GV/cm) and the pseudoscalar electron-scalar nucleon 3. Experiments on diamagnetic systems
SLI, this corresponds to limits de < 9 × 10−29 e·cm and
CS < 6 × 10−9 (both at 90% c.l.). By far the most sensitive experiment using a diamag-
netic system is the long-running Hg EDM search at the
University of Washington (Graner et al., 2016; Swallows
Very recently, results from a new experiment at JILA et al., 2013). Here, 199 Hg atoms (with a 1 S0 closed-shell
were reported (Cairncross et al., 2017). Here, HfF+ ground state) are contained at high density in vapor cells
molecular ions in a metastable 3 ∆1 state are exposed to a (see Fig. 11). Their nuclear spins (I = 1/2) are polarized
rotating E-field (E ∼ 20 V/cm) that serves both to fully by optical pumping with a resonant laser beam, whose
polarize the Ω-doublet levels and to trap the ions (Gresh intensity is modulated at the precession frequency of the
et al., 2016; Leanhardt et al., 2011; Loh et al., 2013). A atoms in the nominally uniform and static applied B-
small, static quadrupolar magnetic field is applied; since field. A stack of four nominally identical cells is used;
molecules orbit a finite distance from the center of the the inner cells have strong, equal and opposite E fields
trap, this lab-frame field gradient causes them to expe- along the B-field axis, while the outer cells have E = 0.
rience a rotating B-field, parallel to the rotating field This configuration makes it possible to cancel fluctua-
E. All state preparation and readout operations are per- tions not only in the average value of B, but also in its
formed in synchrony with the rotating fields. The spin first-order gradient. At the applied field E ≈ 10 kV/cm,
precession frequency is measured relative to this rotating the atomic Hg reaches a polarization P ∼ 3 × 10−5 . The
frame, with a Ramsey measurement sequence of two π/2 cells are filled with ∼ 0.5 atm of CO buffer gas to slow
pulses to prepare a superposition of ±m states and then diffusion of the Hg atoms to the walls, which are coated
to transfer information on the final direction of the spin with paraffin to suppress spin relaxation. After initial
to the populations of these states. Metastable state pop- polarization, the spins freely precess over τ = 170 s, af-
ulation and spin polarization along E is achieved with a ter which nearly all remain polarized. The final spin
series of laser pulses. The π/2 pulses are induced by a direction is probed by monitoring the angle by which the
rotation-induced 3rd -order coupling between the states, linear polarization of a near-resonant probe laser beam is
amplified by briefly reducing E. The population in one rotated as it passes through the atomic vapor. Decades
m state is detemined by a series of laser pulses that pho- of development led to cells with extremely low leakage
todissociates molecules only in that state, and detection currents (< 40 fA). The slow diffusion ensures small mo-
of resulting Hf+ ions. A remarkably long spin coherence tional field effects. The primary systematic errors were
time τ ≈ 700 ms is achieved. However, ion-ion Coulomb associated with nm-scale voltage-induced movements of
interactions in the trap limit the useful molecular den- the vapor cells together with uncontrolled B-field gradi-
sity, leading to a low counting rate dN/dt ∼ 10/s. With ents. With N ∼ 1014 atoms detected in each measure-
∼ 2 weeks of data, this experiment was sensitive to an ment cycle and ∼ 250 days of data, the experiment was
energy shift ∆ETP /h < 0.8 mHz. With the calculated sensitive to an energy shift ∆ETP /h < 20 pHz. From
value E eff ≈ 23 GV/cm for the HfF+ 3 ∆1 state, this cor- the calculated sensitivity of the atomic EDM to the nu-
responds to de < 13 × 10−29 e·cm (90% c.l.), only slightly clear SMt SHg , this set a limit SHg < 3 × 10−13 e·fm3
less sensitive than the ACME result. Earlier experiments, (95% c.l.). This can be interpreted in terms of underly-
one using a beam of YbF molecules (Hudson et al., 2011) ing mechanisms that give rise to S. For example, this
and the other using side-by-side beams of both Tl and yields a limit on the neutron EDM, dn < 1.6 × 10−26 e
sodium (Na, Z=11) atoms (Regan et al., 2002), each set cm, that is more stringent than the best limit from di-
limits about 10× less stringent than those of ACME. rect measurements with free neutrons by a factor of ∼ 2.
38

HV coax cable (RG58) Magnet coil


Magnetic shields
windings
Abe et al. (2014); Denis and Fleig (2016); Skripnikov
Electrode (+10kV)
Wollaston (2016)]. Calculations of electronic structure for diamag-
prism
OT
netic systems—both atoms (Hg, Ra, Xe) and molecules
cell (TlF)—give the ratio between observable energy shift
MT
cell E B0 and nuclear SMt with accuracy < ∼ 20% [see, for example,
MB E Dzuba et al. (2002)]. For the null results from all current
cell
OB
cell
EDM experiments, these small uncertainties have negli-
λ/2 plates gible impact on the limits that can be set on underlying
Vessel (0V) Groundplane y Beam separator
y
physics.
a) HV feedthrough z b) Vessel Electrode x By contrast, theoretical uncertainties associated with
strongly interacting particles are not negligible for in-
FIG. 11 Schematic of the 199 Hg EDM experiment. Purple terpretation of underlying hadronic T,PV parameters.
arrows show probe laser beams, polarization-analyzed by the
There are difficulties with the relations both between
combination of Wollaston prisms and photodiodes. (a) Sec-
tion through the y-z plane of the vessel containing all four va- quark- and nucleon-level parameters (e.g., what value of
por cells, showing probe beams through the outer cells (where the proton EDM dp results from a given value of θ̄QCD
E = 0). (b) Section through the x-y plane showing probe or the up-quark chromo-EDM d˜u ) and between nucleon-
beams through the inner cells. Reproduced with permission and nucleus-level parameters (e.g., what value of a nu-
from (Graner et al., 2016). clear SMt arises from dp or from a given strength of ef-
fective nucleon-nucleon T,PV interaction). In the former
case, the uncertainties are estimated to be at the level of
Similarly, the 199 Hg experiment sets the best limits on ∼ 100%; in the latter, they can be as large as ∼ 500%,
quark cEDMs (d˜u − d˜d < 6 × 10−27 cm) and on the ob- i.e. even the sign of the relation is not reliably known
servable QCD θ-parameter, θ̄QCD < 1.5 × 10−10 as well (Engel et al., 2013; Pospelov and Ritz, 2005). These un-
as on hadronic T -,P -odd couplings, pseudoscalar-scalar certainties are typically not folded into quoted limits on
and tensor-tensor SLI couplings, and the proton EDM dp . fundamental parameters from diamagnetic system EDM
Remarkably, despite having no sensitivity to the scalar- experiments; if properly included, the corresponding lim-
pseudoscalar SLI at lowest order, the limit on CS from its would typically be weaker by factors of a few.
199
Hg is only ∼ 2× less strict than that from ACME.
Other experiments with diamagnetic systems also have
set limits on nuclear SMts and nuclear spin-dependent E. Impact on particle physics
SLI couplings. The most sensitive include searches for
an EDM of 129 Xe (Rosenberry and Chupp, 2001) and To discuss the impact of these experiments, it is useful
225
Ra atoms (Bishof et al., 2016), and for a T,PV energy to begin with crude estimates for the size of the under-
shift in 205 TlF molecules (Cho et al., 1989, 1991). All of lying effects in models with new T,PV physics at a high
these are several orders of magnitude less sensitive to the energy scale (Commins and DeMille, 2009; Pospelov and
underlying physics than is the 199 Hg experiment. Never- Ritz, 2005). First, consider effects associated with the
theless, their sensitivity to different linear combinations EDMs (and cEDMs) of the light fundamental fermions
of the large set of parameters needed to describe T,PV that make up atoms: the electron and the up and down
in these systems makes them useful for providing global quarks. The non-renormalizable EDM Lagrangian LEDM
constraints (Chupp and Ramsey-Musolf, 2015). describes the effect of radiative corrections (Feynman
loop diagrams) in the underlying theory. If the asso-
cated diagram for a fermion with mass mf has n` loops
4. Role of low-energy theory
that contain heavy new particles with mass up to mX , a
typical size of the associated EDM will be
Interpreting the results of atomic and molecular EDM µf n` 2 2
d∼ sin φCP g 2 /2π mf /mX ,
experiments in terms of underlying physical parameters c
requires knowledge of electronic wavefunctions. Calcu- where µf = e~/(2mf ) is the magnetic moment for a
lations of EDM sensitivity (i.e. ratio of atomic EDM Dirac fermion and g is a dimensionless coupling strength
to eEDM or SLI coupling strength) for paramagnetic (e.g., g 2 = α for electromagnetic interactions). The fac-
atoms are similar to those needed to interpret APV ex- tor 1/m2X is associated with the propagator of the heavy
periments, and have accuracy < ∼ 5% for single valence- particle in the loop.
electron atoms such as Tl, Cs, and Fr [see, for ex- In the SM, electron and quark EDMs appear only at
ample, Dzuba and Flambaum (2009b)]. Remarkably, the four- and three-loop level, respectively (Khriplovich
calculations for paramagnetic molecules with one va- and Lamoureaux, 1997). There is a strong addi-
lence electron (YbF, BaF) or even two (ThO, HfF+ ) tional suppression of EDMs in the SM due to a near-
now have accuracy of 10% or better [see, for example, cancellation in the sum over all contributing amplitudes
39

(Hoogeveen, 1990; Nanopoulos et al., 1979; Shabalin, cleons can lead to T,PV SLIs (Barr, 1992b), and two-loop
1978). This mechanism, which arises from the explicit diagrams including the Higgs can lead to fermion EDMs
linkage of flavor mixing and CPV via the CKM ma- and cEDMs (Barr and Zee, 1990). The relative impor-
trix, makes the SM predictions for EDMs extraordinarily tance of the SLI and EDM contributions to T,PV sig-
small—some 5-10 orders of magnitude below current lim- nals depends on the details of parameters in the theory.
its for dn and de , respectively. By contrast, for an uncan- Broadly speaking, however, in these models the 199 Hg
celled 1-loop diagram (n` = 1) and with sin φCP ∼ 1, the and ACME experiments set limits of
current limit on the eEDM corresponds to mX < ∼ 10 TeV;
bounds from 199 Hg on the quark chromo-EDMs probe a sin φh Mh−2
0
< (5 TeV/c2 )−2 ,

similar scale (Barr, 1993; Engel et al., 2013; Pospelov and
Ritz, 2005). where Mh0 is the mass of a new Higgs particle (Barr,
Detailed calculations of the size of the relevant T,PV 1993; Engel et al., 2013; Pospelov and Ritz, 2005). Again,
parameters have been made in a wide range of theoreti- this substantially exceeds direct LHC bounds on Mh0 , if
cal models. Among the most widely explored are models φh ∼ 1.
that incorporate Supersymmetry (SUSY) that is broken A few authors have also begun to explore the impli-
near the electroweak scale, i.e. which predict superpart- cations of EDM limits on the possible existence of new
ner particles with mass MSUSY ∼ MZ ∼ 0.1 TeV. Weak- particles with mass below the electroweak scale, but with
scale SUSY naturally includes many attractive features very weak couplings to ordinary matter—“dark sector”
(Kane, 2002): it stabilizes the Higgs mass against radia- particles and their associated dark forces, discussed in
tive corrections at around its observed value; includes Secs. IV and VII.B.5. It has been argued (Le Dall et al.,
candidate particles for dark matter; modifies the energy- 2015) that within a broad class of models where new par-
dependent running of strong, weak, and electromagnetic ticles appear only at low mass scales, EDM limits provide
couplings so that they converge at a sensible scale for less stringent constraints than other types of experiments
grand unification; and includes new CPV phases δSUSY (aside from limits on θ̄QCD , which is not associated with
that could produce the cosmic baryon asymmetry. a high mass scale and can appear in such a scenario).
However, more generally (e.g., where new particles are
The simplest weak-scale SUSY models include one-
present at both low and high mass), EDMs can provide
loop diagrams that lead to EDMs much larger than the
−2 < (10 TeV)−2 very strict limits on T,PV couplings to dark force carriers
experimental limits, unless δSUSY MSUSY ∼ with mass below the electroweak scale. The first example
(Barr, 1993; Engel et al., 2013; Feng, 2013; Pospelov and
of such an analysis (for tensor SLIs) was carried out in
Ritz, 2005). Improved EDM sensitivity by 1-2 orders of
Gharibnejad and Derevianko (2015).
magnitude will either yield a discovery or conclusively
rule out SUSY models, such as these, that are compat-
ible with electroweak baryogenesis (Balazs et al., 2017; F. Future directions
Cirigliano et al., 2010; Huber et al., 2007). There is grow-
ing interest in models where only a few of the new par- Here we briefly review ongoing or planned T,PV ex-
ticles have MSUSY ∼ MZ while all other SUSY partners periments known to us. A few themes are common in
have much higher mass (Arkani-Hamed and Dimopoulos, new experimental approaches. For example, because of
2005). Here, the primary contribution to EDMs usually the importance of obtaining long spin precession times,
comes from two-loop diagrams. Even in these scenarios techniques of laser cooling (to obtain lower velocities)
the eEDM and quark cEDM limits correspond to lower and trapping (for the longest hold times) are beginning
bounds of ∼ 2 − 4 TeV on the masses of the lighter SUSY to be used. Several new and ongoing experiments are
particles, if δSU SY ∼ 1 (Giudice and Romanino, 2006; also exploiting the high polarizability of polar molecules
Nakai and Reece, 2016). This is well beyond the direct for enhanced sensitivity. A few groups plan to employ
reach of the Large Hadron Collider (LHC) for these types both these concepts, leveraging the recent initial demon-
of particles (Patrignani, 2016). Some typical Feynman strations of laser cooling and trapping of polar molecules.
diagrams leading to particle EDMs are shown in Fig. 12. These methods require optical cycling behavior, which in
SUSY is a well-motivated and thoroughly investi- itself enables both efficient detection (when each molecule
gated extension to the SM. However, in nearly every emits many fluorescent photons) and cooling of internal
model that predicts new physics near the electroweak states, such as rotation, via optical pumping.
scale, new CPV phases appear and T,PV signals in
atomic/molecular experiments should arise at values near
current experimental bounds. For example, there may be 1. Paramagnetic systems
additional scalar fields in nature, analogous to the Higgs
boson. In such multi-Higgs models, the relative phase be- Improvements in all the recent experiments using para-
tween the fields, φh , can lead to CPV (Weinberg, 1976). magnetic molecules are underway. The ACME collabora-
Exchange of the Higgs bosons between electrons and nu- tion plans several upgrades to improve statistics. In the
40

FIG. 12 Example Feynman diagrams leading to particle EDMs. Crosses represent CPV phases and tildes indicate SUSY
partners of SM fields. (a) One-loop diagram leading to an eEDM in a SUSY model. The vertical photon represents the
coupling of the EDM to an electric field. The CPV phase arises from the mechanism that leads to breaking of SUSY at low
energies. (b) Two-loop diagram leading to an up quark cEDM in a multi-Higgs doublet model. Hi,j represent different Higgs
fields, and the CPV phase arises from mixing between them. The dominant diagram includes t quarks since their large mass
indicates a strong coupling to the Higgs field. (c) Dominant two-loop diagrams leading to an eEDM in SUSY models where
partners of fermions are heavy. Here ω ± are SUSY partners of W ± bosons.

ongoing second generation of ACME, improved efficiency coherence times, ∼ 3 s, are anticipated, along with large
of state preparation and detection is anticipated to im- counting rates to compensate for the low Z and low P in
prove the sensitivity to de by ∼20× (Panda et al., 2016). Cs atoms, which yield an effective field Fe E ∼ 107 V/cm
A third generation with increased molecular beam flux with Fe (Cs)= 120 and E ∼ 105 V/cm. A group at To-
from a new source and electrostatic focusing, enhanced hoku University is planning a similar experiment using
detection efficiency via optical cycling, and longer inte- Fr atoms (Inoue et al., 2015), with Fe = 900. Finally,
gration time could yield another ∼30-50× improvement, a group at the University of Groningen is constructing
corresponding to sensitivity at the level de ∼ 10−31 e·cm an apparatus to electrically decelerate (Mathavan et al.,
(Vutha et al., 2010). Simultaneously, the YbF exper- 2016) a beam of BaF molecules (with PEeff ≈ 5 GV/cm),
iment at Imperial College is also improving beam flux then apply transverse laser cooling to obtain a very bright
and velocity by use of a cryogenic beam and rotational and slow molecular beam (Hoekstra, 2016).
cooling, plus optical cycling fluorescence for efficient de-
tection (Rabey et al., 2016). YbF reaches P ≈ 60% at
E = 10 kV/cm, yielding PEeff ≈ 15 GV/cm, ≈ 5× smaller 2. Diamagnetic systems
than Eeff in ThO (Kara et al., 2012). Future plans call
for a dramatic increase in interaction time by use of a A new generation of the 199 Hg experiment is planned,
laser-cooled molecular fountain similar to that used for with various technical improvements to increase sensitiv-
atomic Cs clocks (Tarbutt et al., 2013). Kozyryev and ity by 2-3× (Heckel, 2016). A newer effort at Argonne
Hutzler (2017) recently proposed using certain types of National Lab uses laser-cooled and optically trapped
225
polyatomic paramagnetic molecules that could be laser Ra atoms (Bishof et al., 2016). Here there is a large
cooled, and also have a favorable energy level structure enhancement in the SMt of 225 Ra, for the same size of
for eEDM measurements (similar to the Ω-doublet states underlying parameters such as quark cEDMs. An oc-
used in ACME). Meanwhile, the JILA trapped molecular tupole deformation of the 225 Ra nucleus leads to closely-
ion experiment plans a new trap electrode geometry to spaced opposite-parity nuclear energy levels, analogous
allow use of much larger ion clouds. A ∼10× improved to Ω-doublet levels in polar molecules, and a similar en-
sensitivity to de is projected. In the longer term, use of hancement of the induced SMt due to T,PV effects in the
the heavier species ThF+ will improve Eeff by a factor nucleus (Auerbach et al., 1996). However, the relatively
of 1.5×, and possibly also enable longer spin coherence short half-life of 225 Ra (∼ 15 days) complicates the ex-
time since here 3 ∆1 is the ground state (Cairncross et al., perimental protocol. Calculations of the SMt for given
2017; Gresh et al., 2016). Kawall (2011) also proposed to values of the microscopic T,PV parameters (Ban et al.,
perform similar experiments in a storage ring to further 2010) generally indicate a 100-1000× larger SMt in 225 Ra
increase the trapping volume.
Several other efforts are also under development. A
group at Pennsylvania State University is using laser suggested by Lamoreaux (1989), about the time when it was
cooled and optically trapped Cs atoms, with co-trapped realized that magnetometers separated from the EDM experi-
mental region do not generally provide an adequate measure of
Rb as a comagnetometer5 (Weiss et al., 2003). Here long
magnetic fields correlated with the application of an electric field.
A comagnetometer is different species or atomic/molecular state
with negligible sensitivity to EDMs that occupies the same vol-
ume simultaneously as the species sensitive to EDMs used in the
5 The idea of a comagnetometer for such experiments was first particular search.
41

as compared to 199 Hg; the atomic structure of Ra gives VI. TESTS OF THE CP T THEOREM,
another 3× enhancement (Dzuba et al., 2002). In addi- MATTER-ANTIMATTER COMPARISONS
tion, it may be possible to make the uncertainty in the re-
lation between fundamental parameters and nuclear SMt Current physical laws are believed to be invariant un-
smaller in octupole-deformed nuclei such as 225 Ra than in der the CP T transformation (the CP T theorem), i.e.
more spherical nuclei such as 199 Hg. The Argonne group combined transformations of charge conjugation, spatial
recently reported a limit on the EDM of atomic 225 Ra, inversion and time reversal. Within conventional field
dRa < 1.4×1023 e·cm (95% c.l.), corresponding to a limit theory, the CP T symmetry is closely related to Lorentz
on the SMt of SRa < 2 × 10−7 e·fm3 (Bishof et al., 2016). invariance; however, in more general frameworks such as
While this is ∼ 1000× less sensitive than the 199 Hg exper- string theory, there is a possibility in principle to violate
iment at present, dramatic improvements are anticipated one symmetry without violating the other (Greenberg,
in trapped atom number, detection efficiency, and E-field 2002). This topic has been a subject of recent research
strength (Bishof et al., 2016). Another effort to take ad- and lively debates (Dolgov and Novikov, 2012; Kostelecký
vantage of an enhanced SMt due to octupole deformation and Vargas, 2015; Tureanu, 2013).
is underforway at TRIUMF using 233 Rn, which can be Since weak interactions are not invariant under charge
collected in a vapor cell after production at a radioactive conjugation and also violate CP , a prudent question is
beam facility (Tardiff et al., 2014). Groups at Munich whether violation of these symmetries may result in a
(Kuchler et al., 2016), Mainz (Zimmer, 2017), and Tokyo difference of properties between particles and antiparti-
(Sato et al., 2015) are preparing new measurements of the cles. As it turns out, within the framework of conven-
129 tional field theory, CP T invariance ensures the equality
Xe EDM. All will use vapor cells, where extraordinar-
ily long spin coherence times can be achieved; all also of masses and total lifetimes between particles and an-
use 3 He as a comagnetometer. However, Xe (Z = 54) tiparticles (Lüders and Zumino, 1957) and the same is
has lower intrinsic sensitivity than 199 Hg; moreover, the true for the magnitude of the magnetic moments (Bluhm
inaccessability of optical transitions in Xe forces the use et al., 1997).
of direct magnetic field sensing of the nuclear spins, with Comparison of particle and antiparticle properties,
lower signal/noise than is routinely achieved with laser- therefore, provides tests of the CP T theorem and de-
based detection methods. tection of any discrepancies will be an unambiguous sig-
nal of new physics, motivating such experiments which
have seen significant progress in recent years. Tests of
Finally, groups at Yale, Columbia, and University of the CP T theorem were recently reviewed by Yamazaki
Massachusetts are constructing a new experiment (CeN- and Ulmer (2013), Gabrielse et al. (2014), and Keller-
TREX) to measure the SMt of 205 Tl. CeNTREX will bauer (2015). Here, we present only a brief account of
use a cryogenic beam of TlF molecules, with rotational recent results and progress toward future CP T tests with
cooling and electrostatic focusing for a large useful flux, antiprotons and antihydrogen.
plus optical cycling for efficient detection of laser-induced The ALPHA experiment at CERN demonstrated trap-
fluorescence (Norrgard et al., 2017). The 19 F nucleus will ping of antihydrogen (H) atoms for 1000 s in 2011 (An-
be used as a co-magnetometer. With near-unit polariza- dresen et al., 2011). With the goals of performing spec-
tion, the sensitivity of TlF to the SMt is ∼ 104 × larger in troscopy of the 1S − 2S and hyperfine transitions for
TlF than in the 199 Hg experiment (Dzuba et al., 2002). a comparison with their values in hydrogen, the AL-
This helps to overcome a small spin coherence time of PHA team carried out a proof-of-principle experiment
∼ 15 ms in the TlF cryogenic beam. Future genera- using resonant microwave radiation to flip the spin of
tions will employ transverse laser cooling for improved the positron in magnetically trapped antihydrogen atoms
beam flux and, eventually, optically trapped molecules (Amole et al., 2012). The spin flip caused trapped H to
for long spin coherence time. New experiments also have be ejected from the trap and detected via the annihi-
been proposed to search for a nuclear MQM (Flambaum lation. While this experiment was not aimed at preci-
et al., 2014). Here two enhancement mechanisms can be sion frequency measurement, it bounded the resonance
employed. First, using a nucleus with large quadrupole within 100 MHz of the hydrogen hyperfine frequency,
deformation enhances the MQM by a factor of ∼ 10 − 20 corresponding to a relative precision of about 4 × 10−3
relative to spherical nuclei (Flambaum et al., 1994). Sec- (Amole et al., 2012). The ATRAP collaboration reported
ond, using molecules in a 3 ∆1 state gives the unpaired accumulation of 4 × 109 electron-cooled positrons in a
electron spin needed to couple to the nuclear MQM, high Penning trap for production and storage of antihydrogen
polarization P, and suppressed magnetic moment rela- atoms for future tests of CP T and antimatter gravity
tive to typical paramagnetic systems, just as in the ThO (Fitzakerley et al., 2016).
eEDM experiment. For the same underlying T,PV pa- In 2014, the ASACUSA experiment succeeded for the
rameters, the energy shifts in such a system could be first time in producing a beam of antihydrogen atoms; de-
∼ 107 × larger than in the 199 Hg experiment. tection of 80 antihydrogen atoms 2.7 metres downstream
42

of their production was reported (Kuroda et al., 2014). ment established a limit
This result represents a milestone towards precision spec- (q/m)p
troscopy of the ground-state hyperfine splitting of anti- − 1 = 1(69) × 10−12 (49)
(q/m)p
hydrogen using beam spectroscopy.
and gave a bound on sidereal variations in the measured
An experimental limit on the charge Qe of antihydro-
ratio of < 720 parts per trillion (Ulmer et al., 2015).
gen, in which e is the elementary charge, was reported
Three-body metastable antiprotonic helium p̄He+ con-
by ALPHA collaboration (Amole et al., 2014). In 2016,
sists of an α-particle, an electron and an antiproton, p̄.
they further improved this bound to |Q| < 0.71 parts per
When He captures a slow p̄ in an atomic collision, p̄He+
billion (one standard deviation) (Ahmadi et al., 2016).
is often formed in a high Rydberg state of p̄ orbiting
Assuming charge superposition and using the best mea-
He+ . Such states are amenable to precision laser spec-
sured value of the antiproton charge (Olive et al., 2015),
troscopy in order to determine the antiproton-to-electron
this measurement placed a new limit on the positron
mass ratio and to test the equality between the magni-
charge anomaly, i.e. the relative difference between the
tudes of antiproton and proton charges and masses. Two-
positron and elementary charge, of about one part per
photon spectroscopy of p̄He+ performed by Hori et al.
billion (ppb).
(2011) resulted in the determination of the antiproton-
In December of 2016, ALPHA reported a long awaited to-electron mass ratio mp̄ /me = 1836.1526736(23). Re-
breakthrough result (Ahmadi et al., 2017): they have cently, Hori et al. (2016) employed buffer-gas cooling and
further improved the efficiency of antihydrogen produc- performed single-photon spectroscopy of p̄He+ yielding a
tion (trapping about 14 antiatoms per trial), and em- slightly more precise value of 1836.1526734(15), which
ployed two-photon laser excitation with 243 nm light to agrees with the CODATA recommended value of mp /me
drive the 1S − 2S transition. The initial measurements at a level of 0.8 ppb. Laser spectroscopy of pionic helium
of the transition frequency indicated that it is equal to atoms to determine the charged-pion mass was proposed
its hydrogen counterpart at the level of 2 × 10−10 with by Hori et al. (2014).
further significant improvements anticipated in the near The experimental efforts on matter-antimatter com-
future. In 2017, the results of a microwave spectroscopy parisons aimed at testing whether antimatter is affected
experiment which probed the response of antihydrogen by gravity in the same way as matter are described in
over a controlled range of frequencies were reported (Ah- Sec. X.D.
madi et al., 2017) providing a direct, magnetic-field- Due to the deep intrinsic connection between CP T
independent measurement of the hyperfine splitting of and other symmetries such as Lorentz invariance, test-
1420.4±0.5 MHz, consistent with expectations for atomic ing CP T does not always require antimatter (Pospelov
hydrogen at the level of four parts in 104 . and Romalis, 2004). A recent review of “magnetometry”
ATRAP collaboration (DiSciacca et al., 2013) mea- experiments in this area was given by Jackson Kimball
sured the antiproton magnetic moment with a 4.4 parts et al. (2013b); see also Sec. XI of this review.
per million (ppm) uncertainty with a single particle. The
BASE experiment aims at comparisons of antiproton and
proton magnetic g-factors with a fractional precision of VII. REVIEW OF LABORATORY SEARCHES FOR
EXOTIC SPIN-DEPENDENT INTERACTIONS
δg/g ∼ 10−9 , using the double Penning-trap method
(Smorra et al., 2015). The BASE collaboration observed A. Early work
the first spin flips with a single trapped proton (Ulmer
et al., 2011) and performed a direct measurement of the Ever since the discovery of intrinsic spin [see the his-
magnetic moment of a single trapped proton with a preci- torical review by Commins (2012)], a central question
sion of 3.3 ppb, which is the most precise measurement of in physics has been the role of spin in interactions be-
gp to date (Mooser et al., 2014). The collaboration pro- tween elementary particles. Leptons and quarks, the
poses to use quantum-logic technologies (Dubielzig et al., fundamental fermions, are spin-1/2 particles which in
2013; Heinzen and Wineland, 1990) to trap and probe principle can possess only two possible multipole mo-
(anti)protons by coupling the (anti)proton to an atomic ments: monopole moments (such as mass and charge)
“qubit” ion trapped in its vicinity via Coulomb interac- and dipole moments (such as the magnetic moment). A
tion. This coupling will be used for both ground-state particle’s dipole moment is necessarily proportional to its
cooling of single (anti)protons and for the state readout. spin based on the Wigner-Eckart theorem. In fact, the
The BASE collaboration also performed a comparison inception of the idea of spin was based on the observation
of the charge-to-mass ratio for the antiproton (q/m)p of the anomalous Zeeman effect, a consequence of the in-
to that for the proton (q/m)p using high-precision cy- teraction of the electron’s magnetic dipole moment with
clotron frequency comparisons of a single antiproton and an external magnetic field. It is natural to ask what other
a negatively charged hydrogen ion (H− ) carried out in a sorts of spin-dependent interactions might exist between
Penning-trap system (Ulmer et al., 2015). This experi- fermions apart from the magnetic dipole interaction.
43

1. Torsion in gravity tial of these proposals was the suggestion that a light
spin-0 boson, the axion (Dine et al., 1981; Kim, 1979;
There were a number of hypothetical dipole interac- Shifman et al., 1980; Weinberg, 1978; Wilczek, 1978),
tions postulated and searched for soon after the discov- could possess a coupling to dipoles that might be de-
ery of intrinsic spin. An early theoretical question was tectable in laboratory experiments (Moody and Wilczek,
how to incorporate the concept of spin into the frame- 1984). As Moody and Wilczek note, a spin-0 field ϕ can
work of general relativity. The fact that intrinsic spin couple to fermions in only two possible ways: through a
possessed all the usual properties of angular momentum scalar vertex or through a pseudoscalar vertex. In the
but yet could not be understood as arising from the phys- nonrelativistic limit (small fermion velocity and momen-
ical rotation of an object posed a deep question for at- tum transfer), a fermion coupling to ϕ via a scalar vertex
tempts to extend our understanding of gravity to the acts as a monopole and a fermion coupling to ϕ via a
quantum level. There were indeed general relativistic in- pseudoscalar vertex acts as a dipole. This can be under-
teractions, such as frame-dragging (Lense and Thirring, stood from the fact that in the particle’s center of mass
1918; Thorne and Hartle, 1985), between macroscopic ro- frame, there are only two vectors from which to form
tating bodies possessing angular momentum. But it was a scalar/pseudoscalar quantity: the spin s and the mo-
unclear if analogous effects would exist for particles with mentum p (since the field ϕ is a scalar), so the either
spin since general relativity, being a geometrical theory, the vertex does not involve s (monopole coupling) or if it
did not directly include the possibility of intrinsic spin. does, it depends on s · p, which is a P -odd, pseudoscalar
At the macroscopic scale, mass-energy adds up due to its term. Hence it is the pseudoscalar coupling of ϕ that is
monopole character and leads to observable gravitational the source of new dipole interactions.
effects. On the other hand, spin, due to its dipole charac- The axion emerged from an elegant solution to the
ter, tends to average out for astrophysical bodies such as strong-CP problem (see Sec. V.C.4). The strong-CP
stars and planets. Thus any gravitational effects related problem is that the observable CP -violating phase that
to spin would tend to be difficult to detect through as- can appear in the QCD Lagrangian, θ̄QCD , is known from
tronomical observations, which are the principal vehicles EDM limits to be extemely small: θ̄QCD < −10
∼ 10 . This
for tests of general relativity to this day. Nonetheless, presents a so-called fine-tuning problem, since naı̈vely
soon after the invention of general relativity by Einstein one would expect θ̄QCD ∼ 1. The solution to the strong-
(1916), Cartan proposed an extension of general rela- CP problem proposed by Peccei and Quinn (1977a,b)
tivity that opened the possibility of incorporating spin was that θ̄QCD does not possess a constant value, but
through its effect on the torsion of spacetime (Cartan, rather evolves dynamically. In this model, θ̄QCD is re-
1922, 1923, 1924, 1925). Torsion quantifies the twisting of placed in the Lagrangian by a term representing a dy-
a coordinate system as it is transported along a curve. In namical field, and the quantum of this field is known as
Einstein’s general relativity, mass-energy generates cur- the axion (or, more specifically, the QCD axion). The
vature of spacetime but the torsion is zero, and so vectors underlying physics of the Peccei-Quinn solution to the
curve along geodesics via parallel transport but do not strong-CP problem is closely related to the physics be-
twist. In Cartan’s extension, spin generates nonzero tor- hind the Higgs mechanism endowing particles with mass
sion, and so frames transported along geodesics curve due in the Standard Model: there exists a global continuous
to the effect of mass-energy and twist due to the effect of symmetry in QCD that is spontaneously broken, and a
spin [see, for example, the review by Hehl et al. (1976)]. result of the spontaneous symmetry breaking is the ap-
The consequence is that gravitational dipole interactions pearance of a new “pseudo-Nambu-Goldstone” boson (in
are possible within the framework of Einstein-Cartan the- this case the axion, which is analogous to the Higgs bo-
ory. From another point-of-view, assuming there is a way son). It turns out that the mass of the axion is very small
to parameterize gravity in terms of a quantum field the- [upper limits on the axion mass based on astrophysical
ory, in addition to the spin-2 graviton (the hypothetical observations are < ∼ 10 meV (Raffelt, 1999)], thus produc-
quantum of the gravitational field associated with Ein- ing long-range dipole forces that can be searched for in
stein’s general relativity), there might exist spin-0 and laboratory experiments (Moody and Wilczek, 1984). The
spin-1 gravitons associated with the torsion field. idea of axions spurred theorists to consider other possi-
bilities for light bosons that could mediate dipole interac-
tions between fermions, such as familons (Gelmini et al.,
2. Axions and axion-like-particles (ALPs) 1983; Wilczek, 1982), majorons (Chikashige et al., 1981;
Gelmini and Roncadelli, 1981), arions (Ansel’m, 1982),
The above ideas involved new dipole couplings to and new spin-0 or spin-1 gravitons (Carroll and Field,
known fields, gravitational and electric. It was later re- 1994; Neville, 1980, 1982; Scherk, 1979). Familons are
alized that another possibility existed: there could be pseudo-Nambu-Goldstone bosons arising from sponta-
heretofore undiscovered fields generating dipole couplings neous breaking of flavor symmetry; majorons were devel-
between fermions. Among the earliest and most influen- oped to understand neutrino masses and are constrained
44

by searches for neutrinoless double-β decay; and arions the axion have been extended to a diverse array of prob-
are the bosons corresponding to a spontaneous breaking lems opening new frontiers of research. The numerous
of the chiral lepton symmetry. light pseudoscalar bosons proposed to address a panoply
of theoretical problems in modern physics are known col-
lectively as axion-like particles (ALPs).
3. Early experiments

On the experimental front, early work searching for 1. Axion-like-particles (ALPs) in string theory
new dipole interactions focused on EDMs of neutrons,
nuclei, and electrons (discussed in Sec. V of this review). ALPs generically arise in string theory as excitations of
Later, some attention turned to the role of spin in gravity. quantum fields that extend into compactified spacetime
Morgan and Peres (1962) proposed a test of the equiva- dimensions beyond the ordinary four (Bailin and Love,
lence principle for a spin-polarized body and Leitner and 1987; Svrcek and Witten, 2006). It has been further pro-
Okubo (1964) pointed out that a gravitational monopole- posed by Arvanitaki et al. (2010) that, in fact, because of
dipole interaction would violate P and T (time-reversal) the topological complexity of the extra-dimensional man-
symmetries. If a gravitational monopole-dipole interac- ifolds of string theory, if string theory is correct and there
tion existed, the energy of a particle would depend upon are indeed spacetime dimensions beyond the known four,
the orientation of its spin relative to the local gravita- there should be many ultralight ALPs, possibly popu-
tional field of the Earth. Since no such dependence had lating each decade of mass down to the Hubble scale of
been experimentally observed, Leitner and Okubo were 10−33 eV, a so-called Axiverse.
able to derive corresponding constraints on monopole-
dipole couplings based on the absence of gravitationally
induced splitting of Zeeman sublevels in measurements 2. The hierarchy problem
of the ground state hyperfine structure of hydrogen. A
later experiment searching for such a gravitational dipole Another intriguing hypothesis where axions and ALPs
moment (GDM) of the proton by Velyukhov (1968) in appear is a novel proposed solution to the electroweak
fact found a nonzero value for the proton GDM, but this hierarchy problem (Graham et al., 2015b). The elec-
result was later proved erroneous by Young (1969) and troweak hierarchy problem is essentially the question of
Vasil’ev (1969). Wineland and Ramsey (1972) searched why the Higgs boson mass is so much lighter than the
for a nuclear GDM with orders of magnitude greater sen- Planck mass, for one would expect that quantum cor-
sitivity than previous experiments by using a deuterium rections would cause the effective mass to be closer to
maser. Ramsey (1979) established the first precise con- the Planck scale. Phrased another way, it is surpris-
straints on exotic dipole-dipole interactions between pro- ing that the electroweak interaction should be so much
tons by comparing the measured magnetic dipole inter- stronger than gravity. Attempts to solve the hierar-
action between protons in molecular hydrogen with the- chy problem include, for example, supersymmetry (Di-
oretical calculations. mopoulos and Georgi, 1981) and large (sub-mm) extra
dimensions (Arkani-Hamed et al., 1998; Randall and Sun-
drum, 1999b). Graham et al. (2015b) propose that in-
B. Theoretical motivation stead the hierarchy problem is solved by a dynamic relax-
ation of the effective Higgs mass from the Planck scale to
Speculation concerning the possibility of a spin-gravity the electroweak scale in the early universe that is driven
coupling manifesting as a GDM of elementary fermions by inflation and a coupling of the Higgs boson to a spin-
(Hari Dass, 1976; Kobzarev and Okun, 1962; Leitner and 0 particle dubbed the relaxion, which could be the QCD
Okubo, 1964; Morgan and Peres, 1962; Peres, 1978) or a axion or an ALP. Inflation in the early universe causes
torsion field (Neville, 1980) stood as a principal theoreti- the relaxion field to evolve in time, and because of the
cal impetus encouraging experimental searches for exotic coupling between the relaxion and the Higgs, the effective
spin-dependent interactions for some time until the ap- Higgs mass evolves as well. The coupling between the re-
pearance of the idea of spin-dependent potentials gener- laxion and the Higgs generates a periodic potential for
ated by light spin-0 particles such as the axion (Moody the relaxion once the Higgs’ vacuum expectation value
and Wilczek, 1984) and arion (Ansel’m, 1982). The the- becomes nonzero. When the periodic potential barriers
oretical motivation to search for axions was significantly become large enough, the evolution of the relaxion stalls
boosted when it was realized that axions could be the and the effective mass of the Higgs settles at its observed
dark matter permeating the universe [see, for example, value. A key idea of this scenario is that the electroweak
Sec. IX and also the reviews by Duffy and van Bibber symmetry breaking scale is a special point in the evolu-
(2009), Raffelt (1999), and Graham et al. (2015a)]. tion of the Higgs mass, and that is why the Higgs mass
More recently, the ideas underpinning the concept of eventually settles at the observed value relatively close
45

to the electroweak scale and far from the Planck scale. Model fields, quantum excitations of scale-invariant in-
teractions cannot be described in terms of particles (like
the photon): rather they are objects known as unpar-
3. Dark energy ticles that are unconstrained by any dispersion relation
and without definite mass. The coupling of unparticles
A further theoretical motivation for ALPs comes from to fermions results in long-range spin-spin interactions
attempts to explain the observed accelerating expansion that depend on a nonintegral power of distance between
of the universe, attributed to a so-called dark energy per- the fermions (Liao and Liu, 2007) that can be searched
meating the universe (Peebles and Ratra, 2003). Arkani- for in laboratory experiments.
Hamed et al. (2004) have proposed an infrared (i.e., at
very low energy scales corresponding to the large dis-
tances over which the accelerating expansion of the uni- 5. Paraphotons, dark/hidden photons, and new Z 0 bosons
verse is observed) modification of gravity that posits dark
energy is a ghost condensate, a constant-velocity scalar An entirely different source of new spin-dependent in-
field permeating the universe. The ghost condensate acts teractions are exotic spin-1 bosons. There are twelve
as a fluid filling the universe which turns out to behave known gauge bosons in the standard model: the photon,
identically to a cosmological constant by possessing a the W± and Z bosons, and the eight gluons. Generally
negative kinetic energy term, and thus matches astro- speaking, a massless spin-1 boson accompanies any new
physical observations. The direct coupling of the ghost unbroken U (1) gauge symmetry [such symmetries arise
condensate to matter leads to both apparent Lorentz- quite naturally, for example, in string theory (Cvetic and
violating effects and new long-range spin-dependent in- Langacker, 1996) and other standard model extensions;
teractions (Arkani-Hamed et al., 2004; Arkani-Hamed U (1) refers to the unitary group of degree 1, the collec-
et al., 2005). Along these same lines, Flambaum et al. tion of all complex numbers with absolute value 1 under
(2009) point out that if dark energy is a cosmological multiplication]. Massless spin-1 bosons are referred to
scalar/pseudoscalar field (which could be considered to as paraphotons γ 0 (Holdom, 1986) in analogy with pho-
be a spin-0 component of gravity) there would be a spin- tons, the quanta arising from the U (1) gauge symmetry
gravity coupling. This implies that fermions would pos- of electromagnetism. If paraphotons couple directly to
sess GDMs (as discussed in Sec. VII.A.3), and also pre- standard model particles, in order to generate fermion
dicts spatial and temporal variations of particle masses masses and avoid gauge anomalies (quantum corrections
and couplings. that break the gauge symmetry and lead to theoretical in-
In general, it should be noted that most other such the- consistencies), the gauge symmetry corresponding to the
ories proposing that cosmic acceleration is due to the dy- paraphoton must be U (1)B−L (Appelquist et al., 2003),
namical evolution of a scalar field (termed quintessence), where B − L refers to difference between the baryon (B)
by virtue of possessing a conventional kinetic energy and the lepton (L) number: in other words, the “charge”
term, require a certain level of fine-tuning at least at the of standard model particles with respect to γ 0 is given
level of invoking a nonzero cosmological constant, see for by B − L (so, for example, a proton has B − L = 1
example the review by Joyce et al. (2015). For example, and an electron has B − L = −1). However, if the
in many quintessence models there must exist a screening paraphoton coupling to standard model particles is in-
mechanism of some kind in order to avoid existing astro- direct, i.e., through higher-order processes [so that all
physical and laboratory constraints from tests of gravity standard model particles have zero charge under the new
(see also Sec. VIII). U (1) symmetry], this restriction on the possible charge
is removed and the coupling of quarks and leptons to γ 0
can take on a range of possible values (Dobrescu, 2005).
4. Unparticles Such couplings generate long-range spin-dependent inter-
actions (Dobrescu and Mocioiu, 2006). A closely related
Yet another theoretical idea that motivates searches hypothesis is that of the dark photon, which would com-
for spin-dependent interactions is the unparticle (Georgi, municate a “dark” electromagnetic interaction between
2007). It is possible in the context of quantum field dark matter particles, and could be detectable via mix-
theory that interactions may be scale invariant (Banks ing with photons (Ackerman et al., 2009). Of course,
and Zaks, 1982; Wilson, 1970). A scale-invariant inter- it is also possible that exotic spin-1 bosons possess non-
action’s strength is independent of the energy of the in- zero mass, as does the Z boson in the standard model.
teracting particles. This is not the case for Standard A non-zero mass for such a hypothetical Z 0 boson could
Model fields: in quantum electrodynamics, for example, arise from the breaking of a new U (1) gauge symme-
the strength of the electromagnetic interaction is energy- try. There is a plethora of theoretical models predict-
dependent because of the appearance of virtual particles ing new Z 0 bosons and theoretically motivated masses
(i.e., higher-order processes). In fact, unlike Standard and couplings to quarks and leptons extend over a broad
46

range [see, for example, the review by Langacker (2009)]. possible new bosons and derive limits on theories intro-
Z 0 bosons that do not directly interact with standard duced in Sec. VII.B.
model particles (and therefore reside in the so-called hid-
den sector) are commonly referred to as hidden photons
(Holdom, 1986).
2. Moody-Wilczek-Dobrescu-Mocioiu (MWDM) formalism

6. Conclusions
Generally speaking, the most commonly employed
framework for the purpose of comparing different ex-
Even this brief survey portrays a compelling case for
perimental searches for exotic spin-dependent interac-
experimental searches for exotic spin-dependent interac-
tions is that introduced by Moody and Wilczek (1984)
tions. Such interactions are a ubiquitous feature of the-
to describe long-range spin-dependent potentials asso-
oretical extensions to the standard model and general
ciated with the axion and extended by Dobrescu and
relativity, and furthermore are intimately connected to
Mocioiu (2006) to encompass long-range potentials as-
the mysteries of dark energy, dark matter, the strong
sociated with any generic boson exchange; here we
CP problem, and even the hierarchy problem and grand
denote this framework the Moody-Wilczek-Dobrescu-
unification.
Mocioiu (MWDM) formalism. Given basic assumptions
within the context of quantum field theory (e.g., rota-
C. Parametrization tional invariance, energy-momentum conservation, local-
ity), interactions mediated by new bosons can gener-
1. Introduction ate sixteen independent, long-range potentials between
fermions in the nonrelativistic limit (small fermion ve-
Considering the vast theoretical jungle filled with hy- locity and low momentum transfer). Most laboratory ex-
pothetical new particles (and even unparticles) possess- periments search for interactions between the most com-
ing unknown properties outlined in Sec. VII.B, a reader mon fermions: electrons (e) and nucleons [either protons
may ask: ‘How are we to systematically search for their (p) or neutrons (n)]. In general, because of their differ-
effect on atomic systems and quantify their existence or ent quark content, the couplings of protons and neutrons
lack thereof?’ To set up a general system enabling com- may be expected to differ [for example, in one of the
parison between different experiments that search for most widely studied models of the QCD axion, the so-
the effects of such new particles and fields, let us con- called Kim-Shifman-Vainshtein-Zakharov (KSVZ) model
sider the related question: if a heretofore undiscovered (Kim, 1979; Shifman et al., 1980), the axion coupling to
spin-dependent force exists, how might it affect atoms the proton is >∼ 30 times stronger than that of the neu-
and their constituents: electrons, protons, and neutrons? tron (Raffelt, 1999)]. Thus there are six fermion pairs
It turns out that based on rather general principles, a (ee, ep, en, pp, nn, np) that can couple with sixteen
framework to describe all possible types of interactions different potentials. The potentials are ascribed dimen-
between electrons, protons, and neutrons can be quan- sionless scalar coupling constants, fiXY , between different
tified by “exotic physics coupling constants” for a range fermions (which in general are momentum-dependent,
of length scales. Thus experimental goals are clarified: but can be approximated as momentum-independent in
an experiment searches for an exotic interaction, and if the nonrelativistic limit). Here XY denotes the possible
nothing is found, a limit or constraint is established for fermion pairs: X, Y = e, n, p and i = 1, 2, . . . , 16 labels
coupling constants at the studied length scale for par- the corresponding potential. The potentials can be writ-
ticular forms of interactions. Experimentalists seek to ten in terms of a dimensionless r-dependent function y(r)
explore regions of parameter space that have not been that is determined by the exact nature of the propaga-
previously studied to determine if as-yet-undiscovered tor describing the exotic boson exchange. The potentials
physics exists with such properties. Then particle the- enumerated 1-8 by Dobrescu and Mocioiu (2006) encom-
orists can interpret the experimental results in terms of pass all possible P -even (scalar) rotational invariants:
47

~c
V1 = f1XY y(r) , (50)
r
~c
V2 = f2XY (σ̂ X · σ̂ Y )y(r) , (51)
r
~3 1 2 d2
    
d d
V3 = f3XY
σ̂ X · σ̂ Y 1 − r − 3(σ̂ X · r̂)(σ̂ Y · r̂) 1 − r + r y(r) , (52)
m2 cr3 dr dr 3 dr2
~2
 
d
V4,5 = −f4,5
XY
(σ̂ X ± σ̂ Y ) · (v × r̂) 1 − r y(r) , (53)
2mcr2 dr
~2
 
d
V6,7 = −f6,7
XY
[(σ̂ X · v)(σ̂ Y · r̂) ± (σ̂ X · r̂)(σ̂ Y · v)] 1 − r y(r) , (54)
2mcr2 dr
~
V8 = f8XY (σ̂ X · v)(σ̂ X · v)y(r) , (55)
cr
and those enumerated 9-16 encompass all possible P -odd (pseudoscalar) rotational invariants:
2
 
XY ~ d
V9,10 = −f9,10 (σ̂ X ± σ̂ Y ) · r̂ 1 − r y(r) , (56)
2mr2 dr
2
 
XY ~ d
V11 = −f11 (σ̂ X × σ̂ Y ) · r̂ 1 − r y(r) , (57)
mr2 dr
XY ~
V12,13 = f12,13 (σ̂ X ± σ̂ Y ) · vy(r) , (58)
2r
XY ~
V14 = f14 {σ̂ X × σ̂ Y } · vy(r) , (59)
r
3~3 1 2 d2
 
d
V15 = −f15XY
{[σ̂ X · (v × r̂)](σ̂ Y · r̂) + (σ̂ X · r̂)[σ̂ Y · (v × r̂)]} 1 − r + r y(r) , (60)
2m2 c2 r3 dr 3 dr2
~2
 
d
V16 = −f16XY
{[σ̂ X · (v × r̂)](σ̂ Y · v) + ( σ̂ X · v)[σ̂ Y · (v × r̂)]} 1 − r y(r) . (61)
2mcr2 dr

In Eqs. (50) - (61), ~ is Planck’s constant, c is the speed of field theory, y(r) takes on a Yukawa-like form:
light, m is the fermion mass, r is the distance between the
1 −r/λ
fermions and r̂ is the unit vector along the line between y(r) = e , (62)
them, σ̂ i is a unit vector in the direction of the spin of 4π
fermion i, and v is the average velocity of the fermion where
in the center of mass frame. It is interesting to note
~
that the MWDM formalism applies whether or not the λ= (63)
Mc
underlying theory obeys Lorentz invariance (so long as
rotational invariance is preserved) and also applies in the is the reduced Compton wavelength of the new boson
case of multi-boson exchange between the fermions in of mass M , which sets the scale of the new interaction.
question. Thus the MWDM formalism is quite general If there is multi-boson exchange or Lorentz invariance
in nature and serves as a useful framework for comparing is violated, other forms of y(r) can arise, but the spin
different experiments. dependence of the potential functions is preserved. Gen-
erally an experimental setup characterized by a distance
scale ` is sensitive to new bosons of mass M < ∼ ~/(c`).
[Note that the derivative operators with respect to r are
understood to act only on y(r) and not on wave func-
3. MWDM formalism for Lorentz-invariant, single-boson tions.]
exchange If particular spin and parity properties of the new
boson are specified, correlations between the coupling
A specific form can be obtained for y(r) if some as- strengths are found. For example, if the new boson is a
sumptions are made about the propagator. Assuming spin-0 particle such as an axion or ALP, f3XY = gpX gpY /4,
one-boson exchange within a Lorentz-invariant quantum where gpX,Y parameterizes the vertex-level pseudoscalar
48

coupling (denoted by the subscript p) of the spin-0 field Under these assumptions, for example, the dipole-dipole
to the fermions. The quantity gp2 /(~c) is dimensionless. potential of Eq. (52) can be written in the form most
commonly encountered in the literature,

gpX gpY ~2
    
1 1 1 3 3
V3 (r) = σ̂ X · σ̂ Y + − (σ̂ X · r̂)( σ̂ Y · r̂) + + e−r/λ . (64)
16πmX mY c2 λr2 r3 λ2 r λr2 r3

If the new interaction possesses both scalar and pseu- Moody and Wilczek (1984). Of related interest is the
XY
doscalar couplings, for example, f9,10 = gpX gsY /2 (where fact that the Higgs boson (Aad et al., 2012; Chatrchyan
the subscript s denotes the scalar coupling) one obtains et al., 2012), a spin-0 particle, is predicted to induce a
the following monopole-dipole potential for coupling of Yukawa-like interaction between fermions (Haber et al.,
polarized fermions X to a monopole source of fermions 1979), leading to a delta-function-like potential which
Y: could be searched for in precision atomic physics experi-
ments (Berengut et al., 2017; Delaunay et al., 2017; De-
gpX gsY ~
 
1 1
V9,10 (r) = σ̂ X · r̂ + 2 e−r/λ . (65) launay et al., 2016). The Higgs interaction can even pro-
8πmX c rλ r duce a P -odd, T -odd electron-nucleon interaction gen-
Monopole-dipole and dipole-dipole potentials, and in- erating EDMs of atoms and molecules (Barr, 1992a,b).
deed the vast majority of the potentials enumerated Because the mass of the Higgs boson is ≈ 125 GeV, the
Eqs. (50) - (61), can also be generated by exchange range of any force mediated by the Higgs is ∼ 10−17 m
of spin-1 particles, such as a Z0 boson (Dobrescu and (the Higgs Compton wavelength), and thus meaningful
Mocioiu, 2006; Gomes Ferreiraa et al., 2015). For ex- constraints on Higgs-mediated interactions have not yet
ample, if the new boson is a spin-1 boson, f3XY = been experimentally obtained.
X Y
gA gA + gVX gVY /4, where the subscripts A and V refer A closely related point is that measurements of per-
to the axial vector and vector couplings, respectively. An manent electric dipole moments (EDMs), discussed in
axial vector coupling also generates the dipole-dipole po- Sec. V, also constrain some exotic spin-dependent forces.
tential This is because a P - and T -violating interaction be-
X Y tween particles will naturally induce a P - and T -violating
gA gA ~c
V2 (r) = σ̂ X · σ̂ Y e−r/λ , (66) atomic EDM, and indeed a number of the potentials Vi
4π~c r
violate P and T symmetries [V9,10 , V11 , and V15 – see
which has a different scaling with particle separation as Eqs. (56), (57), and (60)]. Gharibnejad and Derevianko
compared to the V3 (r) potential described by Eq. (64). (2015) have reinterpreted the results of the Hg EDM ex-
A good example of a complete enumeration of the co- periment (Griffith et al., 2009) to constrain a P ,T -odd
efficients fiXY in terms of vertex-level couplings for the interaction of electrons and nucleons through the ex-
particular case of electrons (X = e, Y = e) is given in change of a massive gauge boson, and have excluded vec-
the paper by Leslie et al. (2014). tor bosons with masses > ∼ 1 MeV with coupling strengths
> 10−9 .

4. Contact interactions

Another detail to be aware of is that the potentials de- 5. Position representation and permutation symmetry
scribed in Eqs. (50) - (61) are long-range potentials that
assume the fermions under investigation are separated by Furthermore, in atomic and molecular calculations for
a finite distance. In searches for exotic spin-dependent velocity-dependent potentials, it is useful to convert v
interactions in atoms and molecules one must also take into the relevant momentum operator in position space
into account the possibility of wave function overlap and [the equations of the MWDM formalism (Dobrescu and
the contribution of terms in the potentials proportional Mocioiu, 2006) express the potentials V1 − V16 in a
to Dirac delta functions δ 3 (r). For example, the term “mixed” representation rather than a pure position repre-
gpX gpY ~2 sentation]. Furthermore, for identical particles care must
σ̂ X · σ̂ Y δ 3 (r) (67) be taken to account for permutation symmetry. For ex-
12mX mY c2
ample, the V8 potential [Eq. (55)], which can arise from
must be added to the expression for the dipole-dipole the exchange of axial-vector bosons, can be written for
interaction generated by an ALP given in Eq. (64), see identical particles 1 and 2 as
49

gA ~3
X X
e−r/λ
  
gA
V8 (r) = σ̂ 1 · (∇ 1 − ∇ 2 ), σ̂ 2 · (∇ 1 − ∇ 2 ), , (68)
4π~c 4m2X c r

where ∇i is the vector differential operator in position and the fundamental theories predicting exotic spin-
space for particle i and {, } denotes the anticommu- dependent interactions outlined in Sec. VII.B, although
tator. Details concerning this point are addressed by there are exceptions such as the predicted potentials gen-
Ficek et al. (2017). erated by unparticles (Sec. VII.B.4).
Consider, for example, the standard QCD axion dis-
cussed in Sec. VII.A.2. An axion (or ALP) is character-
6. Quantum field theory details ized by a symmetry breaking scale fa and an interaction
scale Λ, which in the case of the QCD axion is the QCD
In order to check if different experiments are truly mea- confinement scale Λ ∼ 200 MeV (ALPs may have differ-
suring the same quantity, it can sometimes be important ent values for Λ). These scales determine, for example,
to consider further specifics regarding the origin of a spin- the mass of the axion
dependent coupling within quantum field theory. For ex-
ample, an ALP field ϕ can generate the potential V3 (r) Λ2
described by Eq. (64) between fermions ψ of mass m in ma c2 = . (71)
fa
two different ways: either through a Yukawa-like cou-
pling described by the Lagrangian (Moody and Wilczek, The interaction of an axion with a fermion X is deter-
1984; Vasilakis et al., 2009) mined by a dimensionless coupling constant CX which
can be predicted in the context of a specific theory, and
LYuk = −igp ψ̄γ 5 ψϕ , (69)
related to the coupling constants in the MWDM formal-
or through a derivative coupling described by the La- ism. For instance, the pseudoscalar coupling
grangian
gp gX M c2
LDer = ψ̄γµ γ 5 ψ∂ µ ϕ , (70) √p = CX X . (72)
2m ~c fa
where in Eqs. (69) and (70) we have used the Dirac γ ma-
For a particular manifestation of the QCD axion referred
trices. Although experimental searches for dipole-dipole
to as the KSVZ axion (Kim, 1979; Shifman et al., 1980),
interactions are sensitive to both Yukawa-like and deriva-
Cp ≈ −0.34 for the proton, Cn ≈ 0.01 for the neu-
tive couplings, various searches for spin-independent in-
tron, and Ce = 0 for the electron (Raffelt, 1999). Note
teractions and some astrophysical phenomena are sensi-
that in this specific theoretical model a single param-
tive only to one or the other type of coupling (Fischbach
eter, the symmetry breaking scale fa , determines both
and Krause, 1999; Raffelt, 2012; Raffelt and Weiss, 1995).
the axion mass and the coupling strength to particu-
Similarly, Mantry et al. (2014) have shown that by delv-
lar fermions. This formalism connects searches for ex-
ing deeper into the quantum field theoretic origins of ex-
otic spin-dependent interactions to the broader context
otic spin-dependent interactions one can distinguish the
of QCD axion searches: most QCD axion searches exploit
effects of the QCD axion from generic ALPs by com-
the axion-photon coupling, also proportional to 1/fa , but
paring the results of nuclear EDM searches with results
with a different coupling constant. For example, the Ax-
of searches for new spin-dependent forces [see also the
ion Dark Matter eXperiment [ADMX, (Asztalos et al.,
analysis of Gharibnejad and Derevianko (2015)]. It is
2010)] searches for axions converted into detectable mi-
also important to note that QCD axion models (Dine
crowave photons using the inverse Primakoff effect as first
et al., 1981; Kim, 1979; Shifman et al., 1980; Zhitnitskii,
outlined by Sikivie (1983). Since experiments such as
1980) have a definite relationship between the interaction
ADMX probe a different coupling and generally speak-
strength and the axion mass, whereas for a generic ALP
ing a different axion mass range as compared to searches
the mass and the interaction strength are independent
for spin-dependent interactions, these experimental ap-
parameters.
proaches are largely complementary [see Sec. IX and also
the reviews by Kim and Carosi (2010) and Graham et al.
7. Connection between the MWDM formalism and various (2015a)].
fundamental theories As another example, a standard propagating gravi-
tational torsion field (see Sec. VII.A.1) can generate a
In most cases there is a clear one-to-one correspon- dipole-dipole interaction identical to the V3 potential in
dence between potentials in the MWDM formalism the MWDM formalism (Adelberger et al., 2009; Ham-
50

mond, 1995; Neville, 1980, 1982), with the relationship context of quantum field theory, this theoretical bias can
be understood as follows. If an exotic field couples to
2
gp gp 18πGMX L then the field couples to particle current. However,
= β2 , (73)
~c ~c the lowest-order coupling to particle current vanishes if
the exotic interaction is mediated by a spin-0 particle
where G is Newton’s gravitational constant and the min-
such as an ALP (Dobrescu and Mocioiu, 2006). On the
imal torsion model predicts the torsion constant β = 1.
other hand, a coupling of a generic massive spin-1 bo-
In general, there is similar one-to-one correspondence
son to particle current is forbidden by gauge invariance
between the MWDM formalism and any model based on
(Dobrescu, 2005), and constraints on couplings of mass-
a quantum field theory with new force-carrying spin-0
less spin-1 bosons are already quite stringent (Appelquist
and spin-1 bosons.
et al., 2003). Thus, generally, couplings of exotic fields
to particle current, and thus L, are expected to be sup-
8. Relationship between coupling constants for atoms and pressed relative to spin couplings. Nonetheless, it should
elementary particles also be noted that there are theories that do postulate ex-
otic couplings to L. For example, hidden photons can mix
Furthermore, theoretical knowledge of atomic, molec- with ordinary photons, and thus can produce real mag-
ular, and nuclear structure is critical for interpretation netic fields in magnetically shielded regions that would
of experiments. In order to meaningfully compare ex- indeed couple to L (Chaudhuri et al., 2015).
perimental results, the coupling of the exotic field to the The relationship of the expectation value for total
atomic spin must be interpreted in terms of the coupling atomic angular momentum hFi to electron spin hSi
to electron, proton, and neutron spins. The basic scheme and nuclear spin hIi can be evaluated for the ground
of such a parametrization of spin couplings to new physics states of most low-to-intermediate mass atoms based on
can be cast in terms of an exotic atomic dipole moment the Russell-Saunders LS-coupling scheme (Budker et al.,
χ = χa F related to coupling constants χe , χp , and χn 2008):
for the electron, proton, and neutron, respectively, where
F is the total atomic angular momentum. It is generally hFi = hSi + hLi + hIi ,
assumed that such couplings do not follow the same scal-
hS · Fi hL · Fi hI · Fi
ing as magnetic moments. The coupling constants fiXY = hFi + hFi + hFi ,
describing the potentials enumerated in Eqs. (50) - (61) F (F + 1) F (F + 1) F (F + 1)
can then be written in terms of χe , χp , and χn depend- (74)
ing on the specific experiment, where for each different
potential Vi (r) the constants χe , χp , and χn may be dif- where L is the orbital angular momentum. It follows that
ferent. The nucleon coupling constants χp and χn can for the exotic atomic dipole moment coupling constant
in turn be related to quark and gluon couplings via mea- χa ,
surements and calculations based on QCD (Aidala et al.,
2013; Flambaum et al., 2004). hS · Fi hI · Fi
χa = χe + χN , (75)
It is generally assumed by most theories postulat- F (F + 1) F (F + 1)
ing new interactions (Arkani-Hamed et al., 2005; Do-
brescu and Mocioiu, 2006; Flambaum et al., 2009; Georgi, where χN is the exotic nuclear dipole coupling constant
2007; Graham and Rajendran, 2013; Liao and Liu, 2007; which can be expressed in terms of χp and χn .
Moody and Wilczek, 1984) that there is no coupling of The projection of S on F can be calculated in terms of
the exotic field to orbital angular momentum L. In the eigenvalues of the system according to:

hS · Ji
hS · Fi = hJ · Fi , (76)
J(J + 1)
[J(J + 1) + S(S + 1) − L(L + 1)][F (F + 1) + J(J + 1) − I(I + 1)]
= , (77)
4J(J + 1)

where J = S + L, and the projection of I on F is given by

1
hI · Fi = [F (F + 1) + I(I + 1) − J(J + 1)] . (78)
2
51

The next problem is a more difficult one: what is the D. Overview of experimental searches
relationship between χN and the nucleon coupling con-
stants, χp and χn ? Traditionally constraints from atomic A typical approach in experiments searching for exotic
experiments on exotic couplings to neutron and proton spin-dependent interactions is to develop a sensitive de-
spins have been derived using the single-particle Schmidt tector of torques or forces on particles (such as a torsion
model for nuclear spin [see, for example, Venema et al. pendulum) and then bring the detector in close proxim-
(1992)]. In this model, particular atomic species are sen- ity to an object that acts as a local source of the ex-
sitive to either neutron or proton spin couplings, but not otic field (for example, a large mass or highly polarized
both. The single-particle Schmidt model assumes that spin sample). The object sourcing the exotic field acts
the nuclear spin I is due to the orbital motion and intrin- analogously to how a charged object sources an electric
sic spin of one nucleon only and that the spin and orbital field. Usually the major difficulty in such measurements
angular momenta of all other nucleons sum to zero (Blatt is understanding and eliminating systematic errors: in
and Weisskopf, 1979; Klinkenberg, 1952; Schmidt, 1937): other words, distinguishing exotic torques and forces that
in other words, the nuclear spin I is entirely generated would be evidence of new physics from prosaic effects
by a combination of the valence nucleon spin (Sp or Sn ) such as magnetic interactions. For this reason, it is ad-
and the valence nucleon orbital angular momentum `, so vantageous if the source can be manipulated in such a
that we have way as to modulate the exotic field in order to distinguish
its effects from background processes. In lieu of this,
hSp,n · Ii possible sources of systematic errors can be constrained
χN = χp,n , (79) by independent measurements. Another approach, often
I(I + 1)
Sp,n (Sp,n + 1) + I(I + 1) − `(` + 1) used to probe exotic spin-dependent interactions at the
= χp,n , (80) atomic or molecular length scale, is to compare theory
2I(I + 1)
and measurement for some property of a system (such as
the energy splitting between different hyperfine states in
where it is assumed that the valence nucleon is in a well- an atom) that would change if an exotic spin-dependent
defined state of ` and Sp,n . However, it is well known that interaction existed.
nuclear magnetic moments are not accurately predicted As seen in Sec. VII.C, the basic features of an ex-
by the Schmidt model, since in most cases it is a consider- periment that characterize its particular sensitivity are
able oversimplification of the nucleus. Thus, in general, the identities and properties of the particle constituents
the nuclear spin content and magnetic moment cannot of the exotic field source and the detector (determining
be described by a single valence nucleon in a well-defined whether the experiment is searching for neutron-neutron
state of ` and Sp,n . While there have been attempts to interactions, electron-electron interactions, etc.) and the
apply semi-empirical models employing nuclear magnetic distance between the source and detector (which deter-
moment data to derive new constraints for non-valence mines the range of the interaction to which the experi-
nucleons (Engel and Vogel, 1989; Flambaum et al., 2009; ment is sensitive, or, alternatively, the mass of the exotic
Flambaum, 2006; Stadnik and Flambaum, 2015a), Jack- boson communicating the interaction). The precision of
son Kimball (2015) has shown that such models cannot the experiment determines the strength of the interac-
reliably be used to predict the spin polarization of non- tion to which it is sensitive. Depending on whether one
valence nucleons by analyzing known physical effects in or both of the source and detector employ polarized par-
nuclei and by comparisons with detailed large-scale nu- ticles and if the source and detector are in relative mo-
clear shell model calculations [see, for example, Brown tion, the experiment can be sensitive to different poten-
et al. (2016); Vietze et al. (2015)]. Thus while the sensi- tials among those enumerated in the MWDM formalism
tivity of valence nucleons and electrons to exotic physics (Sec. VII.C.2). Most experimental searches to date have
can be reliably estimated, evaluating the sensitivity of been for velocity-independent interactions (V1 , V2 , V3 ,
non-valence nucleons and electrons to exotic physics re- V9,10 , and V11 , see Sec. VII.C.2).
quires detailed theoretical calculations. While most experiments house a macroscopic source
and detector in a single laboratory, thus allowing proxim-
ities between source and detector to range from slightly
less than a millimeter to a few meters [see, for exam-
9. Conclusions ple, Terrano et al. (2015); Tullney et al. (2013); Vasilakis
et al. (2009); Youdin et al. (1996)], the longest-range ex-
Keeping in mind the above caveats, the MWDM frame- periments use the Earth as a source mass [see, for exam-
work introduced by Moody and Wilczek (1984) and Do- ple, Jackson Kimball et al. (2017); Venema et al. (1992);
brescu and Mocioiu (2006) [Eqs. (50) - (61)] provides a Wineland et al. (1991)] or a source of polarized electrons
useful tool to compare different experimental searches for (Hunter et al., 2013), and the shortest-range experiments
exotic spin-dependent effects. probe atomic or molecular structure [see, for example,
52

Ledbetter et al. (2013); Ramsey (1979)]. Experiments


with the Earth as an exotic field source have the partic- Monopole-dipole constraints
ular challenge of lacking a way to reverse or modulate 10 -18 for neutrons
the interaction. Atomic-range experiments suffer from a Wineland et al. (1991)
Venema et al. (1992)

Coupling strength (|g g |/Ñc)


similar challenge insofar as they generally must rely on 10 -22 Youdin et al. (1996)
Raffelt (2012)
a comparison between calculations of energy levels and

p s
Bulatowicz et al. (2013)
Pethukhov et al. (2010)
spectroscopic measurements. Chu et al. (2013)
10 -26 Tullney et al. (2013)
Experiments searching for exotic spin-dependent inter- Guigue et al. (2015)

actions typically employ magnetic shielding between the


source of the exotic field and the detector. Any such ex- 10 -30
periment must answer the basic question: what is the
effect of the magnetic shield system on the signal de-
10 -34
tected by the spin-polarized ensemble? This question laboratory astrophysical
was considered by Jackson Kimball et al. (2016a), and limits limits
the general conclusion is that for common experimen- 10 -38
10 -4 10 -2 10 0 10 2 10 4 10 6 10 8 10 10 10 12
tal geometries and conditions, magnetic shields do not
Length scale l (m)
significantly reduce sensitivity to exotic spin-dependent
interactions, especially when the technique of comagne-
tometry is used [where measurements are simultaneously Monopole-dipole constraints
10 -18
performed on two or more atomic species, see Lamore- for protons
aux (1989)]. However, exotic fields that couple to elec-
Coupling strength (|g g |/Ñc)
tron spin can induce magnetic fields in the interior of 10 -22
p s
shields made of a soft ferro- or ferrimagnetic material.
This induced magnetic field must be taken into account 10 -26
in the interpretation of experiments searching for new
spin-dependent interactions.
10 -30
In the next sections we review the experiments estab-
Youdin et al. (1996)
lishing the best laboratory constraints on various exotic Raffelt (2012)
-34
spin-dependent interactions. 10 Petukhov et al. (2010)
Chu et al. (2013) Laboratory Astrophysical
Jackson Kimball et al. (2017)
limits limits
10 -38
10 -4 10 -2 10 0 10 2 10 4 10 6 10 8 10 10 10 12
Length scale l (m)
E. Experimental constraints on monopole-dipole
interactions
Monopole-dipole constraints
10 -18
for electrons
Figure 13 shows the most stringent laboratory and as-
Coupling strength (|g g |/Ñc)

trophysical constraints on exotic monopole-dipole inter- 10 -22 Wineland et al. (1991)


actions, in particular the V9,10 potentials as described
p s

Heckel et al. (2008)


Youdin et al. (1996)
by the MWDM formalism [Eq. (65)], which can be in- Ni et al. (1999)
10 -26
terpreted as a scalar-pseudoscalar coupling. The hori- Terrano et al. (2015)
Hoedl et al. (2011)
zontal axes show the range of the interaction, inversely Raffelt (2012)
proportional to the mass of the boson communicat- 10 -30
ing the interaction [Eq. (63)]. The vertical axes show
the dimensionless coupling parameter |gp gs | /~c between
10 -34
the studied particles. Typically in experiments, the
laboratory astrophysical
monopole (scalar) coupling is to an unpolarized sample limits limits
with roughly equal numbers of protons, neutrons, and 10 -38
electrons, whereas the dipole (pseudoscalar) coupling is 10 -4 10 -2 10 0 10 2 10 4 10 6 10 8 10 10 10 12
Length scale l (m)
to a polarized sample of predominantly one species, so
the upper, middle, and lower plots in Fig. 13 can be in- FIG. 13 Laboratory constraints (shaded light blue, see text
terpreted as constraints on gpn gsX /~c, gpp gsX /~c, and for discussion of individual experiments) on monopole-dipole
(scalar-pseudoscalar) couplings, |gp gs | /~c [the V9,10 poten-
e X
gp gs /~c, respectively, where the superscripts n, p, and
e refer to neutrons, protons, and electrons, respectively, tials as described in Eq. (65)], for neutrons, protons, and elec-
trons as a function of the range λ of the interaction (gp and
and X = n, p, e for each case.
gs are the pseudoscalar and scalar coupling constants, respec-
tively). Astrophysical constraints (excluded parameter space
shaded light green) are from the analysis of Raffelt (2012).
53

1. Neutrons

At the longest interaction ranges probed by ex-


periments, the most stringent laboratory constraint
on monopole-dipole interactions between spin-polarized
neutrons and other particles is derived from the experi-
ment of Venema et al. (1992), establishing the limit dis-
played on the upper plot of Fig. 13 with a solid black
line. The experiment of Venema et al. (1992) illustrates
the principles involved in a broad class of experiments
that rely on optical measurements of the spin preces-
sion of various atomic species in the gas phase [for re-
views of these experimental techniques, see Budker et al.
(2002); Budker and Jackson Kimball (2013); Budker and
Romalis (2007)]. Venema et al. (1992) simultaneously
measured the spin-precession frequencies of two isotopes
of Hg (this exemplifies the technique of comagnetome-
try) as the orientation of a magnetic field B was changed
relative to the Earth’s gravitational field g. Since the
ground electronic state of Hg is 1 S0 , the ground-state
polarization is entirely due to the nuclear spin I, with
199
Hg having I = 1/2 and 201 Hg having I = 3/2. A
heretofore undiscovered long-range, monopole-dipole in-
teraction would generate spin precession about an axis
directed along the local gravitational field g. In the pres- FIG. 14 Experimental setup from Venema et al. (1992). LP
= linear polarizer, λ/4 = quarter-wave plate, I = iris, PMT
ence of only B and g, the spin precession frequencies for = photomultiplier tube, PEM = photoelastic modulator. Ar-
the two Hg isotopes are rows on right-hand side indicate computer control and data
acquisition. The angles φ± indicate the projection of r̂ (par-
Ω199 = γ199 B + χ199 g cos φ , (81) allel to g) along ẑ (parallel to B) for the two magnetic field
Ω201 = γ201 B + χ201 g cos φ , (82) orientations.

where γi is the gyromagnetic ratio and χi is the so-called


“gyrogravitational ratio” parameterizing the new inter- to generate spin polarization: the angular momentum
action (where the subscripts i denote the respective iso- of the light field is transferred to the atomic sample –
topes), and φ is the angle between B and g. As long as see, for example, reviews by Happer (1972) and Happer
χ199 /χ201 6= γ199 /γ201 (as generally expected), the ra- et al. (2010). In the probe stage, the magnetic field was
tio R = Ω199 /Ω201 acquires a dependence on B and φ if re-directed along ±ẑ in order to induce spin precession.
the χi ’s are nonzero, enabling a search for the long-range The light intensity was reduced so as not to significantly
monopole-dipole coupling. perturb the atomic states, and a photoelastic modula-
The experimental setup is shown in Fig. 14. The Hg tor (PEM) rapidly alternated the light polarization be-
atoms were contained in a cylindrical vapor cell situated tween right- and left-circular in order to reduce vector
at the center of a three-layer cylindrical µ-metal shield light shifts and enable lock-in detection. The detected
with internal coils to apply controlled magnetic fields to signal was demodulated at the PEM frequency (42 kHz)
the atoms. The axes of the concentric cylinders of the and the free-precession-decay signal was analyzed to ex-
shield system (defined to be z) and quadrupole axis of tract the precession frequencies.
the vapor cell (Q̂c ), as well as the magnetic field during Two important systematic errors required special con-
spin precession, were oriented along the Earth’s rotation sideration. The first arose due to a collisional interac-
axis (Ω̂E ). This orientation is designed to make system- tion of the 201 Hg atoms with the walls of the cylindrical
atic errors related to the Earth’s rotation quadratic in vapor cell, causing a ≈ 50 mHz quadrupolar shift. The
the misalignment angle of the apparatus, as discussed quadrupolar wall shift led to resolved splitting of the Zee-
below. The experimental procedure consisted of a pump man frequencies for 201 Hg. The quadrupolar wall shift,
stage and a probe stage. During the pump stage, the and an optical method to cancel it, has recently been
atoms were optically pumped in the presence of a small studied in detail by Peck et al. (2016) for Cs atoms – al-
magnetic field along x (Bx < ∼ 10 mG) by circularly polar- though it should be noted that in this case the quadrupo-
ized light propagating along x̂. Optical pumping involves lar wall shift turns out to be of electronic origin rather
exciting atomic transitions with polarized light in order than nuclear as is the case for Hg. The second systematic
54

error arose because the experimental apparatus was at- plotted in Fig. 13, the work of Jenke et al. (2014) es-
tached to the Earth, while the Hg spins were effectively tablishes the strongest bounds on |gp gs | /~c for distances
decoupled from the Earth’s rotation during the probe between 1-100 microns. In the experiment of Jenke et al.
stage (since the spins were freely precessing). Conse- (2014), transitions between quantum states of ultracold
quently, the Hg isotopes exhibited apparent precession at neutrons confined vertically above a horizontal mirror
the rotation rate of the Earth, ΩE ≈ 2π × 11.6 µHz. This by the Earth’s gravity were driven by resonantly os-
effect, known as the gyro-compass effect (Heckel et al., cillating the mirror position. At even shorter distance
2008), can be understood as the result of viewing an in- ranges, 10−10 m < <
∼ λ ∼ 10
−7
m, the most stringent
ertial system, the Hg spins, from a noninertial frame, the laboratory constraints on monopole-dipole interactions
surface of the rotating Earth. The gyro-compass effect come from measurement of the diffraction of a cold neu-
was studied with even greater precision in the work of tron beam as it passed through a non-centrosymmetric
Brown et al. (2010) and Gemmel et al. (2010). Both quartz crystal (Fedorov et al., 2013), setting the bound
systematic effects were constrained at or below the sta- |gp gs | /~c < −12
∼ 10 .
tistical sensitivity of the experiment by orienting the ap-
paratus so that uncertainty in the effects were quadratic
in the misalignment angles. 2. Electrons
The experiment establishing the strongest laboratory-
scale limit on monopole-dipole couplings of neutrons was The lower plot of Fig. 13 shows constraints on
that of Tullney et al. (2013), shown by the solid blue monopole-dipole interactions of electrons. Most of the
curve in the upper plot of Fig. 13. In the experiment best limits for electrons come from a series of experiments
of Tullney et al. (2013), the spin-precession frequencies using spin-polarized torsion pendulums carried out at the
of co-located gaseous samples of 3 He and 129 Xe were University of Washington in Seattle (Heckel et al., 2008;
measured using a multi-channel, low-Tc Superconducting Hoedl et al., 2011; Terrano et al., 2015), shown by the
QUantum Interference Device (SQUID) to monitor the purple curves. A diagram of the spin-polarized torsion
magnetization. This avoided issues related to light shifts pendulum setup used by Terrano et al. (2015) is shown
that can be problematic in optical atomic magnetome- in Fig. 15. The key piece of the experimental apparatus
try experiments (Acosta et al., 2006; Jackson Kimball was a ring of 20 equally magnetized segments of alter-
et al., 2013a). The source mass was a cylindrical unpo- nating high- and low-spin density materials. The 20-pole
larized BGO crystal (Bi4 Ge3 O12 ) whose position could spin ring was the active element of the torsion-pendulum
be modulated using a compressed-air driven piston be- detector. The high spin density material was alnico and
tween ≈ 2 mm and ≈ 200 mm from a 3 He/129 Xe cell the low spin density material was SmCo5 – a substan-
in order to modulate the strength of the exotic inter- tial degree of the magnetization of SmCo5 comes from
action. The BGO crystal was chosen as the source mass the orbital motion of electrons while the magnetization
based on its high nucleon number density, low conductiv- of alnico is almost entirely due to the electrons’ spin. The
ity (and thus low Johnson-Nyquist noise), and its vanish- magnetization of each alnico wedge is tuned by a localized
ingly small low-field magnetic susceptibility. external field so that the spin-polarized torsion pendulum
At sub-mm distance scales, limits on monopole-dipole has negligible variation in magnetization. Then either an
interactions of the neutron have been obtained by the unpolarized copper attractor or spin-polarized attractor
experiments of Bulatowicz et al. (2013), Petukhov et al. (identical to the pendulum detector) was rotated below
(2010), and Guigue et al. (2015) shown by short-dashed the torsion pendulum at a frequency ω, producing a mod-
red, solid red, and short-dashed blue curves, respectively, ulated torque at 10ω as the source’s high mass (or spin)
in the upper plot of Fig. 13. Bulatowicz et al. (2013) density wedges passed below the high or low spin den-
employed a dual species xenon nuclear magnetic reso- sity segments of the pendulum. The pendulum’s and
nance (NMR) gyroscope with polarized 129 Xe and 131 Xe both attractors’ four cylinders (either tungsten or vac-
to search for a monopole-dipole interaction when a zir- uum) provided gravitational calibration signals at 4ω.
conia rod was moved near the NMR cell. Again the The twisting of the pendulum was measured optically
technique of co-magnetometry was utilized: by simulta- using a reflector cube, and the torque was inferred from
neously comparing the precession frequencies of the two a harmonic analysis of the pendulum twist angle. The
Xe isotopes, magnetic field changes were distinguished experimental setup allowed the attractors to be moved
from frequency shifts due to the monopole-dipole cou- close to the pendulum, with a minimum separation of
pling between the polarized Xe nuclei and the zirconia ≈ 4 mm. The experiment of Heckel et al. (2008) used
rod source mass. The experiments of Petukhov et al. a similar spin-polarized torsion pendulum but with the
(2010) and Guigue et al. (2015) used measurements of hy- Earth and Sun as source masses. Hoedl et al. (2011) used
perpolarized 3 He to constrain the contribution of short- a semiconductor-grade silicon single crystal attached to
range monopole-dipole interactions to relaxation rates. an ultrapure titanium bar as the torsion pendulum in
Although it is outside the range of the parameter space order to have a highly non-magnetic detector, and then
55

mental limits are somewhat sparser. Most measurements


using atomic vapor comagnetometers to search for exotic
spin-dependent interactions use noble gases with valence
neutrons, and therefore, as discussed previously, they are
insensitive to proton couplings. Experiments using spin-
polarized torsion pendulums or solid-state systems are
sensitive to electron couplings. The laboratory-range ex-
periment of Youdin et al. (1996), whose established con-
straints are shown by the long-dashed black curve, is an
exception. Youdin et al. (1996) searched for monopole-
dipole couplings between a 475-kg lead mass and the
spins of 133 Cs and 199 Hg atoms using co-located atomic
magnetometers (consisting of a Cs vapor cell sandwiched
between a pair of 199 Hg cells contained within a system of
magnetic shields, with laser optical pumping and probing
of the atomic spins). Youdin et al. (1996) originally inter-
preted the results of their experiment to constrain only
electron and neutron spin couplings. However, because
FIG. 15 The left panel shows the 20-pole spin-polarized tor- the 133 Cs nucleus has a valence proton, Jackson Kimball
sion pendulum and the right panel shows the unpolarized (2015) noted that in this case the single-particle Schmidt
and polarized sources (upper and lower figures, respectively) model, semi-empirical models, and large-scale nuclear
used to search for monopole-dipole and dipole-dipole interac-
tions in the experiment of Terrano et al. (2015). The µ-metal
shell model calculations are all in reasonable agreement
shielding surrounding the spin-polarized pendulum and the concerning the contribution of the valence proton spin
sources is cut away to show the alnico (green) and SmCo5 to the nuclear spin of 133 Cs. Therefore the experiment
(blue) segments and one of the four pair of calibration cylin- of Youdin et al. (1996) reliably establishes laboratory
ders (red). The mirror cube (in the middle of the pendulum’s constraints on exotic monopole-dipole couplings of the
support structure) is used to monitor the pendulum twist an- proton. Similarly, at short ranges, the experiments of
gle. The entire apparatus is contained within a system of Petukhov et al. (2010) and Chu et al. (2013) establish
magnetic shields. The arrows on the spin attractor indicate
net spin density and direction.
constraints for protons because of the well-understood
contribution of proton spin to the nuclear spin of 3 He
(Anthony et al., 1996; Jackson Kimball, 2015). Chu et al.
used a ferromagnet as a dipole source – this setup en- (2013) search for a spin-precession frequency shift of po-
abled the spin source to be brought into close proximity larized 3 He when an unpolarized mass (either a ceramic
of the detector, allowing sensitivity to monopole-dipole block or a liquid mixture of ≈ 1% MnCl2 in pure water)
forces with ranges of fractions of a mm (i.e., boson masses was moved between 5 cm and 10 µm of the 3 He vapor cell.
> 1 meV). The particular source masses were chosen based on their

Strong laboratory constraints on monopole-dipole cou- nucleon densities, low magnetic impurities and magnetic
plings of electrons were also obtained by Ni et al. (1999) susceptibilities, and minimal influence on the NMR mea-
by using a paramagnetic salt (TbF3 ) and a dc SQUID surement procedure. Although the work of Tullney et al.
to search for induced spin polarization in the TbF3 sam- (2013) also uses polarized 3 He, because the technique of
ple caused by the proximity of a copper mass. Impor- comagnetometry with 129 Xe is employed and there is con-
tant constraints on both electron and neutron monopole- siderable uncertainty regarding the contribution of the
dipole interactions were also obtained in the experi- proton spin to the 129 Xe nuclear spin (Jackson Kimball,
ment of Wineland et al. (1991). They carried out mea- 2015), we do not infer a limit on monopole-dipole in-
surements on trapped 9 Be+ ions as an applied mag- teractions of the proton from this work. Recently Jack-
netic field was reversed relative to the local gravitation son Kimball et al. (2017) completed a search for a long-
field g: the resulting frequency shift between the 9 Be+ range monopole-dipole coupling of the proton spin to the
2
S1/2 |F = 1, M = 0i and 2 S1/2 |F = 1, M = −1i states mass of the Earth using a 85 Rb/87 Rb comagnetometer,
was constrained to be < 13.4 µHz, leading to the limits improving on the long-range limits of Youdin et al. (1996)
shown in Fig. 13 with the dotted black line. by over three orders-of-magnitude. The experiment of
Jackson Kimball et al. (2017) employed overlapping en-
sembles of 85 Rb and 87 Rb atoms contained within an
3. Protons evacuated, antirelaxation-coated vapor cell and simul-
taneously measured the spin precession frequencies us-
The middle plot of Fig. 13 shows constraints on ing optical magnetometry techniques (Budker and Jack-
monopole-dipole interactions of protons. Here experi- son Kimball, 2013) as the magnetic field was reversed
56

relative to the direction of the gravitational field, similar 10 0


Pseudoscalar dipole-dipole constraints
to the experiment of Venema et al. (1992) discussed ear- for protons and neutrons
lier. The measurement of Jackson Kimball et al. (2017)

Coupling strength (|gp gp|/Ñc)


10 -2
establishes the best constraint on the proton GDM. The
experiment was ultimately limited by systematic effects 10 -4
related to scattered light and magnetic field gradients.
10 -6

4. Astrophysical constraints Vasilakis et al. (2009) [nn]


10 -8
Ramsey (1979) [pp]

The green curves and light green shading in Fig. 13 Glenday et al. (2008) [nn]

show the parameter space excluded by astrophysical con- 10 -10 Ledbetter et al. (2013) [np,pp]
Adelberger et al. (2007) [nn,pp,np Yukawa only]
siderations. Raffelt (2012) argues that the coupling con- Klimchitskaya et al. (2015) [nn,pp,np Yukawa only]
stants gs and gp are individually constrained, and thus 10 -12 -12
10 10 -10 10 -8 10 -6 10 -4 10 -2 10 0 10 2 10 4
constraints on their product gs gp can be derived. The
Length scale l (m)
scalar coupling constant gs is constrained by laboratory 10 0
searches for monopole-monopole interactions [the poten- Pseudoscalar
10 -2
tial V1 (r), Eq. (50) – see the review by Adelberger et al. dipole-dipole

Coupling strength (|gp gp|/Ñc)


(2009) and also Sec. VIII]. The pseudoscalar coupling 10 -4 constraints
constant gp for nucleons is constrained by the measured for electrons
10 -6
neutrino signal from supernova 1987A: the ≈ 10 s du-
10 -8
ration of the signal excludes excessive new energy losses
(Raffelt and Seckel, 1988; Turner, 1988), although this 10 -10
constraint is based on the bremsstrahlung process in the 10 -12
collapsed supernova core and thus suffers from signifi-
10 -14
cant uncertainties related to dense nuclear matter ef- Kotler et al. (2015)
fects (Janka et al., 1996). The pseudoscalar coupling 10 -16 Leslie et al. (2014) [e+e-]
Ni et al. (1994)
constant gp for electrons is constrained by star cooling 10 -18 Terrano et al. (2015)
rates (Raffelt and Weiss, 1995). Although the astro- Ficek et al. (2017)
10 -20 -12
physical constraints on |gp gs | /~c are more stringent than 10 10 -10 10 -8 10 -6 10 -4 10 -2 10 0 10 2 10 4
the laboratory limits in all cases, there is both a de- Length scale l (m)
gree of model specificity (Massó and Redondo, 2005) and
some degree of uncertainty regarding the accuracy of stel- FIG. 16 Laboratory constraints (shaded light blue, see text
for discussion of individual experiments) on pseudoscalar
lar models. Furthermore, it is possible that a so-called
dipole-dipole couplings, |gp gp | /~c [the V3 potential as de-
“chameleon mechanism” that screens interactions in re- scribed in Eq. (64)], between nucleons and electrons as a
gions of space with high mass density could invalidate function of the range λ of the interaction. The short- and
astrophysical bounds on new interactions (Jain and Man- long-dashed red lines in the upper plot show constraints de-
dal, 2006). Thus direct laboratory measurements play a rived from spin-independent measurements that apply only to
crucial, comparatively less ambiguous role in determining the Yukawa form of the pseudoscalar interaction [Eq. (69)].
the existence of exotic spin-dependent interactions even The long-dashed red line in the lower plot shows constraints
based on positronium spectroscopy which in order to be com-
when they are somewhat less sensitive than astrophysical
pared with electron-electron constraints must assume CP T
bounds. invariance.

F. Experimental constraints on dipole-dipole interactions recent progress in laboratory searches for exotic dipole-
dipole interactions.
Experimental searches for monopole-dipole interac-
tions have certain appeal because such couplings violate
invariance under both time reversal and spatial inver- 1. Constraints on V3 (r)
sion, and hence one expects negligible background from
standard-model physics. Dipole-dipole couplings, on the The best limit on long-range pseudoscalar dipole-
other hand, are even under both T and P and can arise dipole interactions [of the form given by the V3 poten-
from standard model physics. In this sense, dipole-dipole tial described in Eq. (64)] between neutrons was achieved
couplings may be problematic for exotic physics searches in the experiment of Vasilakis et al. (2009) (solid black
because one must carefully account for standard-model curve in the upper plot of Fig. 16) using the setup shown
physics effects. Nonetheless, there has been impressive in Fig. 17. The measurement technique is based on the
57

moments, leaving the K-3 He comagnetometer sensitive


to inertial rotation (Kornack et al., 2005) and anomalous
spin couplings (Vasilakis et al., 2009). Thus the self-
compensating K-3 He comagnetometer enables one to use
high-sensitivity SERF magnetometry techniques for de-
tection of anomalous spin-dependent interactions causing
precession about axes transverse to ẑ.
The K spin-polarization along x̂ is determined by mea-
suring optical rotation of an off-resonant, linearly po-
larized probe light beam [see, for example, the review
by Budker et al. (2002) for a discussion of optical rota-
tion]. After residual magnetic fields and light shifts are
eliminated using zeroing routines [described in detail by
Kornack et al. (2005)], the K spin-polarization along x̂
can only arise due to non-magnetic, spin-dependent in-
teractions — offering a highly sensitive probe of such
anomalous interactions. The spin source in the exper-
iment of Vasilakis et al. (2009) consisted of a dense
(≈ 3×1020 cm−1 ), highly polarized (≈ 15% polarization)
FIG. 17 Experimental setup of Vasilakis et al. (2009). PD = 3
photodiode, SP = stress plate to control polarization of the He gas located approximately 50 cm from the cell. The
probe beam, T = translation stage to shift the probe beam, nuclear spin direction of the 3 He sample was reversed at
P = polarizer, PMF = polarization maintaining fiber, OA = a 0.18 Hz rate by adiabatic fast passage. After approxi-
optical amplifier, LCW = liquid crystal wave plate, PEM = mately one month of data acquisition, no anomalous ef-
photoelastic modulator, λ/4 = quarter-wave plate, LDA = fect was detected at a level corresponding to a magnetic
laser diode array. field value less than an attoTesla (10−14 G).
Constraints on pseudoscalar dipole-dipole couplings
between protons at the molecular scale were deduced
principles of spin-exchange-relaxation-free (SERF) mag- by Ramsey (1979) based on molecular-beam experiments
netometry Allred et al. (2002); Kominis et al. (2003); with hydrogen (H2 ). Comparing the measurements of
Kornack et al. (2005); Kornack and Romalis (2002). Harrick et al. (1953) to calculations of the magnetic
The atomic sample consists of overlapping ensembles of dipole-dipole interaction between the protons in H2 lim-
potassium (K) and 3 He at relatively high vapor den- ited the possible contribution of an exotic dipole-dipole
sities (3 He density ≈ 1020 atoms/cm3 and K density interaction to spin-dependent energy splittings, estab-
≈ 1014 atoms/cm3 ). The K sample is polarized through lishing the constraint shown by the short-dashed black
optical pumping and the 3 He sample is polarized through curve in the upper plot of Fig. 16. Ledbetter et al. (2013)
spin-exchange collisions with K. The vapor cell is located obtained the constraints on proton-proton and neutron-
within a five-layer µ-metal shield fitted with internal coils proton pseudoscalar dipole-dipole couplings shown by the
used to cancel residual magnetic fields and create a small long-dashed black curve in the upper plot of Fig. 16
field B parallel to the propagation direction of the pump by comparing NMR measurements to theoretical cal-
beam. culations of indirect nuclear dipole-dipole coupling (J-
Under these experimental conditions, for which the coupling) in deuterated molecular hydrogen (HD). The
Larmor frequencies are comparable to or smaller than the Hamiltonian describing J-coupling has the form JI1 · I2
frequency of spin-exchange collisions, the spin-exchange (I1,2 are the nuclear spins and J parameterizes the inter-
interaction between K atoms and the polarized 3 He va- action strength) and arises due to a second-order hyper-
por strongly couples the two spin ensembles (Kornack fine interaction where the interaction between the nuclear
et al., 2005; Kornack and Romalis, 2002). In a spheri- spins is mediated through the electron cloud. The mea-
cal cell this coupling can be represented as an effective surements from which Ledbetter et al. (2013) extracted
magnetic field B eff experienced by one spin species due constraints were performed with HD in the gas phase:
to the average magnetization M of the other. The ap- thus the internuclear vector r̂ was randomly reoriented
plied field B is tuned so that it approximately cancels the due to collisions. This effect leads to an averaging of
B eff experienced by the K atoms. The K atoms are then Eq. (64), so that its distance scaling becomes propor-
effectively in a zero-field environment. Because the 3 He tional to e−r/λ /(λ2 r). The collisional averaging reduces
magnetization M adiabatically follows B, components of sensitivity to exotic dipole-dipole forces for which λ dif-
B transverse to ẑ are automatically compensated by B eff fers significantly from the mean internuclear separation,
to first order. Such cancelation only occurs for interac- as seen in Fig. 16. Of interest in regard to the constraints
tions that couple to spins in proportion to their magnetic derived from J-coupling in HD are more recent measure-
58

ments and calculations indicating tension between theory 10 -10


and experiment (Garbacz, 2014; Neronov and Seregin, Axial-vector
10 -15 dipole-dipole
2014).

Coupling strength (|gA gA|/Ñc)


constraints
Other notable experiments searching for exotic dipole- 10 -20
dipole couplings of nucleons include the work of Glen-
day et al. (2008), an experiment similar to that of Vasi- 10 -25

lakis et al. (2009) that employed a dual-species 3 He-


129 10 -30
Xe maser as the detector, and constraints from Adel-
berger et al. (2007) and Klimchitskaya and Mostepa- 10 -35 Ledbetter et al. (2013) [pp,np]
Karshenboim (2011) [en,ep]
nenko (2015) based on short-range tests of the gravita- Hunter et al. (2013) [ee,en]
tional inverse-square law and the Casimir effect. The 10 -40 Kotler et al. (2015) [ee]
Vasilakis et al. (2009) [nn]
work of Adelberger et al. (2007) and Klimchitskaya and 10 -45 Heckel et al. (2013) [ee]
Mostepanenko (2015), which actually search for spin- Ritter et al. (1990) [ee]

independent interactions, constrain only the Yukawa 10 -50 -15


10 10 -10 10 -5 10 0 10 5 10 10
form of the pseudoscalar coupling [Eq. (69)] and are thus
Length scale l (m)
more model-specific than the other laboratory searches
considered. The constraints from Adelberger et al. (2007) FIG. 18 Laboratory constraints (shaded light blue, see
and Klimchitskaya and Mostepanenko (2015) do not ap- text for discussion of individual experiments) on axial-vector
ply to the derivative form [Eq. (70)] that would be ex- dipole-dipole couplings, |gA gA | /~c [the V2 potential as de-
pected for Goldstone bosons such as the axion. Con- scribed in Eq. (66)], for various particles as a function of the
straints on spin-dependent interactions derived from ex- range λ of the interaction.
perimental searches for spin-independent interactions are
also considered by Aldaihan et al. (2017).
For electrons, the experiments of Terrano et al. (2015) (2013) took advantage of the large number of polarized
and Ni et al. (1994) establish the most stringent con- electrons in the Earth: there are ∼ 1049 unpaired elec-
straints on pseudoscalar dipole-dipole forces at interac- tron spins in the Earth, yielding ∼ 1042 polarized geo-
tion ranges >∼ 1 mm (solid and long-dashed black curves, electrons polarized by the Earth’s magnetic field. Thus
respectively, in the lower plot of Fig. 16). The experiment the number of polarized geoelectrons exceeds that of a
of Terrano et al. (2015) was addressed in the preceding typical laboratory source by a factor of > 17
∼ 10 . How-
section on monopole-dipole interactions (see Fig. 15 and ever, a typical laboratory source of polarized electrons
surrounding discussion). Ni et al. (1994) used a SQUID can be placed closer than a meter away from a detector
to measure the magnetization of a paramagnetic salt whereas the mean distance of a polarized geoelectron is
(TbF3 ) induced by dipole-dipole interactions with rotat- > 105 m from a detector on the surface of the Earth. For

ing spin-polarized samples (Dy6 Fe23 and HoFe3 ). From pseudoscalar dipole-dipole interactions, the 1/r3 distance
atomic scales up to a mm, the agreement between energy scaling makes searches for exotic interactions with geo-
structure calculations and spectroscopic measurements in electrons less competitive than searches employing po-
He (Ficek et al., 2017) provide the most stringent con- larized laboratory sources. On the other hand, the 1/r
straints, shown by the black dot-dashed line in the lower scaling of the axial-vector interaction makes the huge
plot of Fig. 16. number of polarized geoelectrons a much stronger source
with which to search for long-range interactions. Hunter
et al. (2013) used data from optical atomic magnetome-
2. Constraints on V2 (r) ters (Peck et al., 2012; Venema et al., 1992) and a spin-
polarized torsion pendulum (Heckel et al., 2008) to de-
Figure 18 shows the laboratory constraints on axial- rive the limits shown in Fig. 18. The experiments of Peck
vector dipole-dipole couplings, |gA gA | /~c, described by et al. (2012) and Heckel et al. (2008) utilized rotatable
the V2 potential in the MWDM formalism [Eq. (66)]. In mounts for their entire experimental apparatus in order
terms of experiments, the critical difference between the to modulate the signal from the polarized geoelectrons,
V2 and V3 potentials is the scaling with particle sepa- a technique also employed in the experiment of Brown
ration: the V2 potential scales as 1/r whereas the V3 et al. (2010).
potential scales as 1/r3 . Thus experiments searching for Many of the experiments searching for exotic dipole-
dipole-dipole interactions can have vastly different sensi- dipole interactions previously discussed also place strong
tivities to the two different potentials. constraints on axial-vector interactions between various
An excellent example illustrating the importance of particles (Heckel et al., 2008; Ledbetter et al., 2013; Vasi-
the distance scaling is the work of Hunter et al. (2013) lakis et al., 2009). Between a micron and a mm, the
which established the long-range axial-vector constraints best direct constraint on axial-vector dipole-dipole inter-
shown by the solid black curve in Fig. 18. Hunter et al. actions between electrons comes from the measurement
59

of the magnetic dipole-dipole interaction between two (61)]. For example, Vasilakis et al. (2009) and Hunter
trapped 88 Sr+ ions (Kotler et al., 2014, 2015). Kotler et al. (2013) specifically searched for the V11 potential
et al. (2014) trapped two 88 Sr+ ions using a linear radio- [Eq. (57)]. Jackson Kimball et al. (2010) used measure-
frequency Paul trap, and the ions were initialized in an ments and calculated cross sections for spin exchange
entangled state that was insensitive to spatially homo- between alkali metal atoms and noble gases (specifi-
geneous magnetic field noise. This technique enabled cally sodium and helium) to constrain anomalous spin-
precise measurement of the magnetic dipole-dipole in- dependent forces between nuclei at the atomic scale, and
teraction between the ions, which when compared to established the first limits on the V8 potential [Eq. (55)].
a straightforward calculation gave good agreement at a Hunter and Ang (2014) used polarized geoelectrons to
level of ≈ 200 µHz. Kotler et al. (2015) then used the constrain many of the other velocity-dependent poten-
agreement between experiment and theory to limit the tials: V6,7 , V8 , V14 , V15 , and V16 [Eqs. (54), (55), (59) -
strength of exotic dipole-dipole interactions as shown (61), respectively]. Yan and Snow (2013) used measure-
by the short-dashed red curve in Fig. 18. Ritter et al. ments of a P -odd spin rotation when a cold neutron beam
(1990) (solid purple curve in Fig. 18) carried out an ex- passed through a liquid 4 He target to set limits on V12,13
periment with a spin-polarized torsion pendulum made at short ranges (10−6 m < <
∼ λ ∼ 1 m), and Yan et al.
3
from Dy6 Fe23 , which had the characteristic that at a (2015) used He spin-relaxation rates with the Earth as
particular temperature (between 265 K to 280 K) the an unpolarized source mass to constrain V12,13 at long
magnetization due to the orbital motion of the electrons ranges (λ > ∼ 1 m). Piegsa and Pignol (2012) were able to
approximately canceled the magnetization from the elec- establish bounds on V4,5 at the mm scale for neutrons us-
tron spins, allowing a torsion pendulum with large net ing Ramsey’s method of separated oscillatory fields with
intrinsic spin but small magnetic moment, a similar idea a cold neutron beam that travelled past a nearby cop-
to that behind the later work of the University of Wash- per plate. Heckel et al. (2008) constrained long-range
ington group discussed previously (Heckel et al., 2008; velocity-dependent potentials between their torsion pen-
Terrano et al., 2015). Karshenboim (2010a,b,c) com- dulum and the Moon and Sun.
pared spectroscopic measurements of hyperfine structure It bears mentioning that there have been several tests
to QED calculations for various atomic systems in or- of the universality of free fall (UFF) performed with spin-
der to derive constraints on axial-vector interactions, the polarized objects, in particular with cold atoms (Duan
strongest constraints coming from hydrogen, deuterium, et al., 2016; Fray et al., 2004; Tarallo et al., 2014). At
and 3 He+ . present, such experiments are orders of magnitude less
sensitive to the potentials described in Eqs. (50) - (61)
than the experiments described in Secs. VII.E and VII.F.
3. Astrophysical constraints The basic reason for this is that free-fall experiments es-
sentially measure the spatial derivative of Vi whereas the
It should be noted that laboratory limits on pseu- experiments using optical atomic magnetometers or tor-
doscalar interactions are weaker than relevant astrophys- sion pendulums measure the energy shift due to Vi di-
ical constraints on gp from the neutrino signal from SN rectly. Sec. X discusses UFF tests using both polarized
1987A (Engel et al., 1990), the metallicity of stars (Hax- and unpolarized test masses along with other experimen-
ton and Lee, 1991), the maximum brightness of red gi- tal probes of the equivalence principle.
ants (Raffelt and Weiss, 1995), and null searches for ax- As noted in Sec. VII.B, there are a variety of other
ion emission from the Sun (Derbin et al., 2009). How- theories predicting spin-dependent interactions that are
ever, these astrophysical constraints do not apply to not well-described by the potentials outlined in Eqs. (50)
axial-vector or vector interactions (Dobrescu and Mo- - (61), and several experiments have specifically sought
cioiu, 2006). Since both the V2 and V3 potentials can be to measure such effects. Glenday et al. (2008), Vasi-
generated by spin-1 bosons, astrophysical constraints – lakis et al. (2009), and Heckel et al. (2008) searched for
specific to the vertex-level interactions for spin-0 bosons the hypothetical ghost condensate resulting from spon-
(Dobrescu and Mocioiu, 2006; Raffelt, 1999) – do not taneous breaking of Lorentz symmetry (Arkani-Hamed
apply in general to the V2 and V3 potentials and are et al., 2004; Arkani-Hamed et al., 2005). Vasilakis et al.
therefore not shown in Figs. 16 and 18. (2009) and Hunter et al. (2013) searched for the poten-
tials arising from unparticles (Georgi, 2007; Liao and
Liu, 2007). Many experiments have analyzed their re-
G. Experimental constraints on other forms of sults in terms of gravitational torsion (Heckel et al., 2008;
spin-dependent interactions Jackson Kimball et al., 2010; Kotler et al., 2015; Led-
better et al., 2013; Lehnert et al., 2014, 2015). Ivanov
A number of experiments have searched for some of and Snow (2017) have proposed that gravitational tor-
the other forms of exotic spin-dependent potentials enu- sion generates a new type of P -even and T -odd potential
merated by Dobrescu and Mocioiu (2006) [Eqs. (50) - that can be probed using spin-polarized particles moving
60

through unpolarized matter that is rotating in the lab- sensitivity of a precessing needle magnetometer can far
oratory frame. Undoubtedly, the rich theoretical land- surpass that of magnetometers based on spin precession
scape of exotic spin-dependent interactions will continue of atoms in the gas phase. Under conditions where noise
to inspire a vibrant array of experiments. from coupling to the environment is subdominant, the
scaling with measurement time t of the quantum- and
detection-limited magnetometric sensitivity is t−3/2 . If a
H. Emerging ideas magnetometer based on a precessing ferromagnetic nee-
dle can be experimentally realized, a measurement of nee-
A major new direction in the search for exotic spin- dle precession averaged over ≈ 103 s could reach a sensi-
dependent interactions is the push to study oscillating tivity to exotic electron-spin-dependent couplings at an
and transient signals from fields comprised of new bosons energy scale of ∼ 10−26 eV. If such an experimental
such as axions, ALPs, and hidden photons that may con- sensitivity could be achieved in practice, it would probe
stitute dark matter or dark energy. These ideas are dis- exotic spin-dependent interactions more than five orders
cussed in Sec. IX, and include global networks of optical of magnitude weaker than present laboratory limits.
atomic magnetometers (Pospelov et al., 2013; Pustelny
et al., 2013) and atomic clocks (Derevianko and Pospelov,
2014) to search for correlated transient signals herald- VIII. SEARCHES FOR EXOTIC SPIN-INDEPENDENT
ing new physics that might arise from topological defects INTERACTIONS
(Pospelov et al., 2013; Stadnik and Flambaum, 2014b) or
A. Introduction
clumps of virialized ultra-light fields (Derevianko, 2016).
There are also new experiments using NMR (Budker
One of the exotic potentials described by the
et al., 2014), atomic spectroscopy (Stadnik and Flam-
MWDM formalism deserves special attention, namely V1
baum, 2014a), and resonant electromagnetic detectors
[Eq. (50)] — the sole potential among those discussed
(Chaudhuri et al., 2015) to search for coherently oscil-
in Sec. VII that has no dependence on the spins of the
lating dark matter fields. A related proposal is that of
interacting fermions. Experimental searches for such ex-
Romalis and Caldwell (2013), who have noted that cos-
otic spin-independent interactions have a long history,
mological scalar fields, which may explain dark energy,
mostly from the perspective of tests of the inverse-square
have local spatial gradients that could have detectable
law (ISL) of gravity. Originally the idea was to see if
electromagnetic couplings.
the gravitational force law followed the form (Adelberger
There are also new ideas being developed for novel
et al., 2003)
sources and detectors that can be used to search for ex-
otic spin-dependent interactions. Chu et al. (2015) pro- mX mY
FG (r) = −G r̂ , (83)
posed the use of new paramagnetic insulators, in particu- r2+
lar gadolinium gallium garnet (Gd3 Ga5 O12 , or GGG), to
search for spin-dependent interactions. Ledbetter et al. where FG is the gravitational force between test masses
(2012) have proposed a new class of liquid state nuclear mX and mY separated by a distance r, G is Newton’s
spin comagnetometers with potential sensitivities in the gravitational constant, and  is a parameter character-
10−11 Hz range for one day of measurement. Another izing deviation from the ISL. Since the r−2 scaling of
new concept being developed by Arvanitaki and Geraci the gravitational force law derives from the geometry of
(2014) combines the techniques used in short-distance three-dimensional space, it turns out, generally, that a
tests of gravity employing torsion pendulums (Kapner force law of the form given by Eq. (83) is difficult to moti-
et al., 2007) and micro-cantilevers (Geraci et al., 2008) vate from a theoretical perspective. Instead, the modern
with those used in NMR experiments in order to search perspective follows the MWDM formulation, positing a
for short-range monopole-dipole interactions. Yukawa-like deviation from the ISL; the common α − λ
Jackson Kimball et al. (2016b) have recently predicted parametrization (Talmadge et al., 1988) found in the lit-
that a ferromagnetic needle will precess about the axis erature proposes a modified form of the gravitational po-
of a magnetic field at a Larmor frequency Ω when IΩ  tential given by
N ~, where I is the moment of inertia of the needle about GmX mY  
the precession axis and N is the number of polarized V 0 (r) = − 1 + αe−r/λ , (84)
r
spins in the needle. In this regime the needle behaves
as a gyroscope with spin N ~ maintained along the easy where α characterizes the strength and λ characterizes
axis of the needle by the crystalline and shape anisotropy. the range of the modified gravitational interaction. From
Such a precessing ferromagnetic needle is a correlated the point-of-view of quantum field theory, such a mod-
system of N spins which can be used to measure mag- ification of the gravitational interaction is equivalent to
netic fields for long times. In principle, by taking ad- effects generated by the exchange of a new boson as in
vantage of rapid averaging of quantum uncertainty, the the MWDM formalism. Typically in the literature such
61
p
a Yukawa-like, spin-independent interaction is referred to mass M Pl ∗ = ~c/G is given by
as a fifth force (Fischbach et al., 1986; Fujii, 1971). Cor-  n
respondence between the two viewpoints can be made ∗ 2 2 R cn R n
(M Pl ) ≈ M Pl ≈ M Pl 2+n , (86)
explicit: exchange of scalar or vector bosons between `Pl ~n
fermions X and Y can be described with
where `Pl = ~/(M Pl c) is the “true” Planck length. Set-
X Y ting M Pl ≈ M EW , for n = 2, R ∼ 100 µm. Although re-
~c gs,v gs,v
α=− , (85) cent experiments (Bezerra et al., 2011; Chen et al., 2016b;
4πG mX mY Kamiya et al., 2015; Kapner et al., 2007; Sushkov et al.,
2011a; Tan et al., 2016; Yang et al., 2012) and astrophys-
X,Y
where gs,v characterizes the vertex-level scalar (sub- ical constraints (Arkani-Hamed et al., 1999; Barger et al.,
script s) or vector (subscript v) coupling generating a 1999; Cullen and Perelstein, 1999; Hall and Smith, 1999)
long-range V1 potential [Eq. (50)]. The range λ is under- have excluded the n = 2 possibility, scenarios with n ≥ 3
stood in this case to be the reduced Compton wavelength and variations on the ideas of Arkani-Hamed et al. (1998)
of the new scalar or vector boson. Although there have involving, for example, extra dimensions with nonuni-
been numerous alternative theoretical proposals for spe- form compactification scales (Lykken and Nandi, 2000)
cific forms of modified gravitational potentials, to a large and alternative metrics for the extra dimensions (Ran-
degree these considerations are moot for experimental dall and Sundrum, 1999b), including the possibility of
work since all searches for ISL violations have to date infinite-sized extra dimensions (Randall and Sundrum,
returned null results; Eq. (84) is entirely adequate for 1999a), are still viable and provide motivation for con-
phenomenological comparison of different experimental tinued tests of the ISL.
constraints. In the event a violation is detected, how- A second aspect of the conflict between general relativ-
ever, it will be necessary to pursue determination of the ity and quantum field theory is the cosmological constant
specific form of the new interaction. problem or vacuum energy catastrophe (Weinberg, 1989).
There have been a number of recent comprehensive Observational evidence suggests that the accelerating ex-
reviews on the topic of ISL tests and searches for ex- pansion of the universe may be explained by a nonzero
otic spin-independent interactions, we refer the reader cosmological constant associated with a vacuum energy
to the works by Adelberger et al. (2003); Antoniadis density ρvac ∼ 4 × 103 eV/cm3 , the so-called dark en-
et al. (2011); Fischbach and Talmadge (2012); Gundlach ergy. However, rough estimates of ρvac based on the Stan-
(2005); Lamoreaux (2012); Murata and Tanaka (2015); dard Model assuming no new physics up to the Planck
Newman et al. (2009); Onofrio (2006) for more details scale suggest a vacuum energy density ∼ 10122 eV/cm3 , a
on this subject. In this section we offer a brief overview staggering discrepancy. The vacuum energy scale derived
of the field and recent developments. from cosmological observations corresponds to a length
scale
s
~c
B. Motivation and Theoretical Landscape `vac ≈ 4 ∼ 100 µm . (87)
ρvac

Theories motivating searches for ISL violations and A suggested theoretical path toward resolving the cos-
fifth forces are often inspired by the inherent conflict be- mological constant problem is the proposal that some-
tween general relativity and quantum field theory. One how the gravitational interaction with vacuum fluctua-
aspect of this conflict is the hierarchy problem, the enor- tions “cuts off” at length scales <
∼ `vac (Sundrum, 1999),
mous gulf between the Higgs mass and the Planck mass indicating that one might generically expect a change
(discussed in Sec. VII.B). An influential theoretical sug- in gravitational physics below ∼ 100 µm. It is sugges-
gestion that inspired a new generation of short-range ISL tive that two of the most significant theoretical problems
tests was the proposal by Arkani-Hamed et al. (1998, confronting quantum theories of gravity both indicate a
1999) that the hierarchy problem could be resolved if benchmark scale of ∼ 100 µm where a deviation from the
there existed relatively large (sub-mm scale) extra com- ISL might be expected.
pact spatial dimensions in which gravitons could prop- As noted in Sec. VIII.A, the existence of new scalar or
agate but Standard Model particles could not. In this vector bosons could also give rise to apparent violations of
scenario, n extra dimensions beyond the ordinary four the ISL due to the appearance of a new Yukawa potential
are compactified with characteristic radius R and the hi- between fermions. Such new bosons commonly appear in
erarchy problem is resolved by setting the “true” Planck grand unification theories such as string theory (Bailin
mass M Pl ≈ M EW , the electroweak scale. The observed and Love, 1987) as well as in related theories involving ex-
long-range strength of gravity is a result of the dilu- tra dimensions such as those discussed above (Antoniadis
tion of the field through the extra dimensions, so from et al., 1998), supersymmetric theories (Taylor, 1990), and
Gauss’s Law the apparent “four-dimensional” Planck many others (Adelberger et al., 2003; Antoniadis et al.,
62

2011; Dobrescu and Mocioiu, 2006). Two specific exam- cantilever and a source mass with alternating 100 µm-
ples from string theory are often cited as possible targets wide gold and silicon strips that is moved beneath the
of searches: radions (Brans and Dicke, 1961), which are cantilever (Geraci et al., 2008).
scalar bosons related to the radius of extra dimensions, A feature common to all recent torsion-pendulum tests
and dilatons (Arvanitaki et al., 2015), which are scalar of the ISL and micro-cantilever experiments is the use
bosons that determine the interactions between particles of a thin conducting membrane between the source and
in string theory. Particles such as radions and dilatons test masses that acts as an electrostatic shield. Be-
are known collectively as moduli, scalar bosons whose cause of the challenges related to manufacturing conduct-
expectation values determine key parameters in string ing membranes thinner than a few microns, experimen-
theory (Schellekens, 2013). tal tests of the ISL below a few microns have generally
Another important theoretical motivation to search for had to contend with distant-dependent electromagnetic
new scalar bosons is the idea of quintessence, the pro- forces due to the Casimir effect (Lamoreaux, 1997) and
posal that the accelerating expansion of the universe is a electrostatic patch potentials (Kim et al., 2010; Sushkov
result of the potential energy of a scalar field; for reviews et al., 2011b). The Casimir effect [reviewed by Lamore-
see Frieman et al. (2008); Linder (2008); Padmanabhan aux (2004), for example] is the attraction or repulsion
(2003); Peebles and Ratra (2003); Tsujikawa (2013). Fur- between objects due to modification of the electromag-
thermore, there have been a number of proposals that at- netic vacuum modes in the space between the objects,
tempt to explain dark energy as a modification of gravity which appears as an additional short-range force. Pre-
at cosmological distance scales; for a review, see Joyce cise comparisons between Casimir effect measurements
et al. (2015). To produce the observed accelerating ex- and calculations provide some of the best constraints on
pansion, the modification of gravity would correspond to fifth forces for 10−7 m < < −5 m (Bezerra et al.,
∼ λ ∼ 10
a long-range scalar interaction. However, modified grav- 2011; Masuda and Sasaki, 2009; Sushkov et al., 2011a).
ity at such large distance scales immediately confronts Experiments by Chen et al. (2016b) employing a mi-
stringent observational tests at the solar system scale and cromechanical torsional oscillator have recently probed
shorter distances (Will, 2014) and is ruled out. To avoid the 4 × 10−8 m < <
∼ λ ∼ 10
−5
m range by coating the
these observational constraints, there have been a num- surface of an alternating density source mass with gold
ber of proposals that the new scalar component of gravity in order to keep the Casimir effect uniform as the po-
is somehow screened within the solar system, for exam- sition of the source mass is varied (Decca et al., 2005;
ple via self-interactions (Khoury and Weltman, 2004b; Matloob and Falinejad, 2001), improving on the Casimir
Olive and Pospelov, 2008), modified Newtonian dynamics effect measurement constraints on α by several orders of
[MOND; see, for example, the work of Milgrom (1983)], magnitude.
or other nonlinearities (Vainshtein, 1972). These screen- At even smaller length scales, on the order of 0.01 –
ing mechanisms are, in turn, associated with new par- 10 nm, the best constraints on fifth forces come from ex-
ticles such as chameleons (Khoury, 2013) and galileons periments measuring the scattering of neutrons off of no-
(Nicolis et al., 2009) that can be searched for in labora- ble gas atoms (Kamiya et al., 2015; Nesvizhevsky et al.,
tory experiments. 2008; Pokotilovski, 2006). Atomic and molecular spec-
troscopy can also produce meaningful constraints at this
length scale. In particular, spectroscopy of atomic hy-
C. Laboratory tests drogen (Dahia and Lemos, 2016a; Wan-Ping et al., 2015)
and molecular hydrogen [H2 , HD, and D2 , (Gato-Rivera,
Many experimental searches for fifth forces and tests 2015; Salumbides et al., 2015)] have been used in conjunc-
of the ISL employ torsion pendulums, an experimen- tion with theoretical calculations of atomic and molecular
tal technique discussed in Sec. VII.E in the context of energy levels to constrain the models of gravity postulat-
searches for spin-dependent interactions (see Fig. 15 and ing extra dimensions discussed in Sec. VIII.B.
surrounding discussion). Torsion-pendulum tests of the Another closely related class of experimental probes of
ISL are reviewed by Adelberger et al. (2009) and New- gravity involve tests of the Einstein equivalence principle
man et al. (2009); recent torsion-pendulum experiments (EEP) that underpins general relativity. The EEP states
by Kapner et al. (2007), Yang et al. (2012), and Tan et al. that any local experiment (local in the sense that gravita-
(2016) have established the most stringent constraints on tional tidal effects may be neglected) cannot distinguish
α for 10−5 m < <
∼ λ ∼ 10
−2
m [Eq. (84)], probing the between a gravitational field and an acceleration of the
theoretically interesting region of parameter space cov- laboratory. Recent precise measurements of the gravita-
ering up to three orders of magnitude below the nomi- tional redshift using atom interferometry by Müller et al.
nal strength of gravity around the dark energy scale of (2010); Poli et al. (2011); Zhou et al. (2015) verify the
`vac ∼ 100 µm. Between 5 and 15 microns, the best predictions of general relativity with an accuracy better
constraint on a fifth force comes from measurements em- than 10−8 and represent some of the best tests of the EEP
ploying a ∼ 1 µg test mass attached to cryogenic micro- to date (Hohensee et al., 2011). The technique of atom
63

interferometry has demonstrated significant potential to IX. SEARCHES FOR LIGHT DARK MATTER
test the EEP and other aspects of general relativity with
even greater precision in the near future (Biedermann A. Introduction
et al., 2015; Dimopoulos et al., 2008a).
An alternative approach to testing the EEP employing A variety of astrophysical and cosmological measure-
atomic spectroscopy has achieved a sensitivity matching ments (Bertone et al., 2005; Feng, 2010; Gorenstein and
that of atom interferometer experiments: Hohensee et al. Tucker, 2014) strongly suggest that over 80% of all mat-
(2013a) used measurements of the transition frequency ter in the Universe is invisible, nonluminous dark matter
between two nearly degenerate opposite-parity states of (DM). Understanding the microscopic DM nature is one
atomic dysprosium over the course of two years. The of the paramount goals of cosmology, astrophysics, and
long-term measurement of the energy splitting between particle physics, since it will not only reveal the origins of
the nearly degenerate dysprosium states served as a sen- the dominant constituent of matter in the Universe but
sitive clock comparison test: the EEP implies that the also offer insights into the cosmology of the early Uni-
relative frequencies of any set of clocks at relative rest verse, uncover new physical laws, and potentially lead to
and located within a sufficiently small volume of space- the discovery of other fundamental forces.
time must be independent of the velocity, orientation, The evidence for DM is derived from observations of
and gravitational potential of their rest frame. The mea- DM’s gravitational effects at the galactic scale and larger:
surements of Hohensee et al. (2013a), showing that the galactic rotation curves (Rubin and Ford Jr., 1970; Ru-
transition frequency varied by less than 1 Hz over two bin et al., 1980; Zwicky, 1933), gravitational lensing (Re-
years, constrained electron-related anomalies in gravita- fregier, 2003; Tyson et al., 1998), the Bullet Cluster
tional redshifts at the 10−8 level. (Clowe et al., 2006) and other galactic cluster studies
As mentioned in Sec. VIII.B, theoretical attempts to (Lewis et al., 2003), large-scale structure of the Uni-
ascribe the accelerating expansion of the universe to a verse (Allen et al., 2003), supernovae surveys (Perlmutter
long-range modification of gravity appear to require a et al., 1999; Riess et al., 1998), and the cosmic microwave
screening mechanism in order to evade experimental lim- background radiation (Komatsu et al., 2011). All these
its on fifth forces. Experiments using atom interferom- observations point toward the existence of DM. In or-
etry have established the most stringent constraints on der to fully elucidate the nature of DM, terrestrial ex-
such theories. Hamilton et al. (2015) used a Cs matter- periments seek to measure non-gravitational interactions
wave interferometer near a spherical source mass in an of DM with Standard Model particles and fields. How-
ultra-high vacuum chamber, thereby reducing any screen- ever, the vast extrapolation from the > ∼ 10 kpc distances
ing mechanisms by searching for a fifth force with individ- associated with astrophysical observations to particle-
ual atoms rather than bulk matter (in contrast to the tor- physics phenomena accessible to laboratory-scale experi-
sion pendulum, microcantilever, and Casimir-effect ex- ments leaves open a vast number of plausible theoretical
periments discussed above). possibilities worth exploring.
It is notable that the types of scalar particles that In order to develop an experimental strategy for terres-
would mediate fifth forces, such as dilatons (Van Tilburg trial DM detection, it is useful to consider what can be
et al., 2015), may also constitute the dark matter (in surmised about the local DM environment of our solar
the same way that the axions and ALPs mediating spin- system based on astrophysical observations. The local
dependent interactions can be dark matter, as mentioned DM density is best estimated from the galactic rotation
in Sec. VII.H). Consequently, atomic physics techniques curve of the Milky Way, which, it turns out, is rather
can be employed to search for dark matter scalar bosons challenging to measure from the vantage point of our so-
as discussed in detail in Sec. IX. There are also a num- lar system. Furthermore, in the end, the galactic rotation
ber of new proposals on the horizon that promise im- curve only offers incomplete information on the local DM
proved sensitivity to spin-independent interactions at density since it is sensitive to the integrated mass density
various length scales and new ways to test the EEP and between our location and the center of the galaxy, and
ISL: examples include experiments employing optically the mass density near the galactic center is notoriously
trapped microspheres and nanospheres (Geraci and Gold- difficult to determine. Nonetheless, based on numerical
man, 2015; Geraci et al., 2010), Bose-Einstein conden- models (Bergström et al., 1998) and observations of other
sates (Dimopoulos and Geraci, 2003), and novel atom similar spiral galaxies (Salucci and Borriello, 2003), it is
interferometry (Hohensee et al., 2012) and atomic-clock believed that the Milky Way is embedded within and ro-
experiments (Abele et al., 2010). An alternative way to tates through a spherical DM halo.
look for exotic interactions is to see if, for example, a The commonly used standard halo model predicts that
mass can source a scalar field that changes fundamental the DM energy density local to the Solar system is
constants; such an experiment can be competitive with ρDM ≈ 0.3 − 0.4 GeV/cm3 ; this corresponds to a mass
those searching directly for new forces as surveyed in this density equivalent to one hydrogen atom per a few cm3 .
section (Leefer et al., 2016). Further, in the galactic rest frame the velocity distribu-
64

tion of DM objects is isotropic and quasi-Maxwellian, ble II): the classical field can cause precession of nu-
with dispersion v ' 290 km/s and a cutoff above the clear/electron spins, drive currents in electromagnetic
galactic escape velocity of vesc ' 550 km/s (Freese et al., systems, and induce equivalence-principle-violating ac-
2013). The Milky Way rotates through the DM halo celerations of matter (Graham and Rajendran, 2013).
with the Sun moving at ∼ 220 km/s roughly towards They could also modulate the values of the fundamen-
the Cygnus constellation. Therefore one may think of tal “constants” of nature, which can induce changes in
a DM “wind” impinging upon the Earth, with typical atomic transition frequencies (Arvanitaki et al., 2015;
relative velocities vg ∼ 300 km/s ≈ 10−3 c. The speed of Derevianko and Pospelov, 2014) and local gravitational
the Earth with respect to the DM halo is also seasonally field (Geraci and Derevianko, 2016). Some of these phe-
modulated at a level of ∼ 10% due to the Earth’s orbit nomena have been searched for in other contexts de-
around the Sun. Furthermore, the prevailing view based scribed throughout this review (see Secs. II,VII,VIII,XI,
on astrophysical observations is that the DM is cold, i.e., and X). Here we examine the particular characteristics
moving with velocities much smaller than the speed of of effects induced by light DM fields and how preci-
light. sion measurement techniques such as nuclear magnetic
To date, experimental efforts to detect DM have resonance (NMR), atomic and SQUID (Superconduct-
largely focused on Weakly Interacting Massive Parti- ing QUantum Interference Device) magnetometry, elec-
cles (WIMPs), with masses between 10 and 1000 GeV tromagnetic resonators, atomic/optical interferometers,
(Bertone et al., 2005; Feng, 2010). Despite considerable and atomic clocks can be used to search for these ef-
effort, there are no conclusive signs of WIMP DM inter- fects. When the mass of the particle constituting the
actions, even as experimental sensitivities have improved DM is sufficiently light, the classical DM field leads to
rapidly in recent years. While the WIMP is theoretically persistent time-varying signals that are localized in fre-
well-motivated, it is by no means the only DM candidate. quency at the DM mass, enabling rejection of technical
Observational limits permit the mass of DM constituents noise while permitting signal amplification through res-
to be as low as 10−33 GeV or as high as 1048 GeV. A onant schemes. The classical fields sourced by “clumpy”
number of candidates inhabit this vast parameter space, DM could cause transient signals that can be observed by
ranging from ultra-light axions and axion-like particles correlating output from multiple synchronized detectors.
(ALPs), which are discussed in relation to new interac- The entire field of laboratory cosmology, where table-
tions in Secs. VII and VIII, to more complex dark sectors top-scale precision measurement experiments search for
that lead to composite DM “clumps.” terrestrial signatures of effects related to light DM, has
While particle detectors work by measuring energy emerged as a vibrant research area over the last few years
deposition, precision measurement techniques are well with a number of promising new proposals joining sev-
suited for detecting candidates that act as coherent en- eral ongoing experiments. As noted above, based purely
tities on the scale of individual detectors (or networks of on the known properties of DM, the range of parameter
detectors). Aided by recent advances in fields such as space to be explored is vast. However, experiments can
optical and atomic interferometry, magnetometry, and be guided by clues from other fields of physics suggesting
atomic clocks, several promising new experimental con- mysteries that can be solved by postulating, for example,
cepts have been recently proposed to employ these tech- new DM candidates with particular properties — this is
nologies to search for DM candidates with masses be- what distinguishes the most theoretically well-motivated
tween 10−33 GeV and 10−12 GeV. Methods to probe light DM candidates (by the Occam’s razor principle).
ultra-heavy, composite DM candidates with astrophys- Among the most well-motivated light DM candidates is
ical and terrestrial measurements have also emerged. the quantum chromodynamic (QCD) axion, discussed in
The key idea behind these concepts is that light DM Sec. VII.A.2; experimental axion searches were recently
particles have a large mode occupation numbers and their reviewed, for example, by Graham et al. (2015a). Axions
phenomenology is described by a classical field. For this can naturally constitute a significant fraction of DM: for
mass range the DM candidates are necessarily bosonic; example, they can be produced in sufficient abundance in
non-interacting fermionic candidates would require larger the early universe via the so-called vacuum realignment
masses to reproduce the standard halo model velocity process (Abbott and Sikivie, 1983; Dine and Fischler,
distribution. The requirement that the field is classi- 1983; Preskill et al., 1983). This process results from a
cal translates into the upper mass limit of ∼ 10 eV [see misalignment between the axion field generated when ax-
detailed discussion in Derevianko (2016)]. The lower ions are first produced via spontaneous symmetry break-
mass limit of 10−24 eV comes from requiring that the ing and the minimum of the potential generated by QCD
de Broglie wavelength is smaller than the size of galax- interactions. Since the axion field is initially perturbed
ies where gravitational signatures of DM have been ob- from the QCD potential minimum, it will oscillate; these
served. While such classical fields may arise in a wide oscillations are not significantly damped over the age of
variety of DM models, their effects on Standard Model the universe and in fact in most models it is the energy
particles include a finite number of possibilities (see Ta- stored in these coherent oscillations of the axion field
65

Spin Type Operator Interaction DM effects Searches


scalar ϕ h† h, φn OSM Higgs portal / dilaton fund.-constant variation Atomic clocks, GPS.DM
0 a Gµν G̃µν axion-QCD nucleon EDM CASPEr-Electric
pseudo-scalar a F µν F̃µν axion-E&M EMF along B field ADMX, CULTASK, MADMAX
(∂µ a)ψ̄γ µ γ5 ψ axion-fermion spin torque CASPEr-Wind, GNOME, QUAX
0
Fµν F µν vector–photon mixing EMF in vacuum DM Radio, ADMX
0
1 vector Fµν ψ̄σ µν ψ dipole operator spin torque CASPEr-Wind
axial-vector A0µ ψ̄γ µ γ 5 ψ minimally coupled spin torque CASPEr-Wind

TABLE II Current experimental efforts in searches for bosonic ultralight dark matter candidates. The Table lists illustrative
couplings of bosonic ultralight dark matter candidates [scalar ϕ, axion a and dark photon A0µ ] to SM fields, and their exper-
imental effects. h, Gµν , F µν , and ψ represent respectively SM Higgs, gluon, photon, and fermion fields, or operators of that
form. OSM stand for terms from the SM Lagrangian density. n = 1, 2 for linear/quadratic couplings. Free fields cause signal
oscillations at the field Compton frequency and self-interacting fields forming DM “clumps” can cause transient effects. Specific
experiments are discussed in Sec. IX.B. The table is not exhaustive, as for example, the GPS.DM and GNOME experiments
could be sensitive to monopole topological defects which require vector fields. Modified table from Graham et al. (2016b).

that constitute the mass-energy ascribed to DM (Dine paradigm for the birth of the universe. This (along with
and Fischler, 1983; Duffy and van Bibber, 2009; Preskill other factors) led to alternative possibilities for axion pro-
et al., 1983). Similar scenarios describe the production duction in the early universe. As several authors have
of most light bosons. Another production axion mecha- pointed out, these allow a much larger mass range for the
nism is through the decay of topological defects such as QCD axion, and in fact bestows the lighter axions with
domain walls or strings (Chang et al., 1998; Davis, 1985; a strong theoretical motivation (Freivogel, 2010; Linde,
Nagasawa and Kawasaki, 1994), where the topological 1988).
defects interpolate between regions of space with differ- Going beyond the QCD axions, SM extensions offer a
ent axion vacuum fields which can exist, for example, due plenitude of ultralight DM candidates. We collectively
to nontrivial vacuum structure (i.e., multiple equivalent refer to such candidates as virialized ultralight fields
local minima in the self-interaction potential). (VULFs). Possibilities are compiled in Table II, where
QCD axions couple to photons, gluons, and fermion the fields are characterized by their spin and intrinsic par-
spins over a predictable range of axion mass-coupling ity. In the considered mass range (< 10 eV) the DM can-
strength parameter space (Abbott and Sikivie, 1983; didates are necessarily bosonic. In particular, spin-1 par-
Dine and Fischler, 1983; Preskill et al., 1983). There ticles, commonly referred to as dark or hidden photons
are three possible interactions of axions with Standard (see Sec. VII.B.5) could conceivably constitute a substan-
Model particles and fields: axions can couple to electro- tial fraction of the DM. Table II also indicates various
magnetism, induce electric dipole moments (EDMs) for potential DM couplings to SM fields. More broadly, the
nucleons (see Sec. V) via interaction with the gluon field, potential non-gravitational couplings can be classified us-
and can cause precession of electron and nucleon spins ing “portals” corresponding to different gauge invariant
(see Table II). operators of the SM fields coupling to operators that con-
There are robust astrophysical constraints on QCD tain fields from the dark sector. This phenomenological
axions with masses > approach is widely used in particle physics for searches
∼ 10 meV based on the observa-
tion of the neutrino signal from supernova 1987A and of DM and dark forces (Rouven et al., 2013). For ex-
star cooling (Raffelt, 1999). Heavier axions would have ample for scalar DM fields ϕ, the following portals may
produced observable effects in astrophysical objects, and arise (Derevianko and Pospelov, 2014):
much heavier axions would already have been seen in ∂µ ϕ X
terrestrial experiments. Constraints have also been con- L1 = cψ ψ̄γµ γ5 ψ axionic portal,
Λ
sidered for QCD axions with masses < ∼ 1 µeV based on
SM particles
ϕ X (s)
cosmology. However, these constraints depend upon as- L2 = cψ mψ ψ̄ψ linear scalar,
sumptions about unknown initial conditions of the uni- Λ
SM particles
verse. Such lighter-mass QCD axions were never ruled ϕ2
(2s)
X
out either by experimental or astrophysical observations, L3 = cψ mψ ψ̄ψ quadratic scalar,
Λ2
but theory prejudice held that they were less likely based SM particles
on cosmology. It has now been realized that this was iϕ∗ ∂µ ϕ X
based on a particular scenario for the earliest epochs in L4 = gψ ψ̄γµ ψ current − current .
Λ2
SM particles
the universe, a time about which we know little. Since the
inception of this cosmological argument against lower- Here ψ are SM fermion fields with associated masses mψ ,
mass QCD axions, inflation has become the dominant Λ are the energy scales, and ci are individual coefficients
66

that can take on different values depending on type of from an intense laser light field and a static field of a
ψ. This classification can be generalized to include the strong magnet which facilitates mixing between photons
SM gauge bosons, for example of the form ϕ Fµν F µν for and axions. These axions are then detected by convert-
electromagnetism and extended further to non-scalar DM ing them back to photons after they cross a wall that is
fields. transparent to them but completely blocks the light. In
While the phenomenological portal classification is “helioscope” experiments (Graham et al., 2015a; Raffelt,
broad, one should be aware of certain existing astrophys- 1999), the task of producing axions or ALPs is “sub-
ical and laboratory constraints on the coupling strengths contracted” to the Sun (hence the name), but detec-
or energy scales Λ. For example, the hypothesized tion is accomplished as in LSW experiments. Finally,
DM fields can mediate forces and thus the limits from “haloscopes” directly detect the DM from the galactic
fifth-force experiments (Sec. VIII) immediately apply. halo. In a somewhat complementary approach, indirect
Thereby an experiment searching for DM signatures experiments search for modifications of the known in-
through a specific portal must probe yet unexplored pa- teractions via exchange of virtual exotic particles. Such
rameter space. In some cases, the broader search may experiments include the “fifth-force” searches and exper-
soften such constraints. For example, the discussed iments looking for exotic spin-dependent interactions or
bounds on the QCD axion are relaxed for ALPs (Massó modification of fundamental constants in the presence of
and Redondo, 2005). ALPs are pseudo-scalar particles massive and/or spin-polarized objects that presumably
similar in nature to the QCD axion that do not solve the act as sources of the virtual exotic particles. We discuss
strong CP problem, but rather emerge naturally from some examples of direct experimental searches of differ-
other frameworks such as string theory. ALPs may also ent kinds below, while indirect searches are discussed in
have the properties necessary to solve the hierarchy prob- Secs. VII and VIII. The direct detection of “clumpy”
lem, as discussed in Sec. VII.B. DM objects requires networks of precision measurement
As mentioned earlier, a distinct theoretical possibility tools, and we discuss here two ongoing searches with mag-
is that DM is not distributed uniformly but rather occurs netometers and atomic clocks.
in the form of “clumps.” Even the ever-present grav-
itational interaction leads to instabilities and clumping.
Examples of “clumpy” objects include “dark stars” (Bar- 1. Microwave cavity axion experiments
ranco and Bernal, 2011; Braaten et al., 2016; Iwazaki,
2015; Kolb and Tkachev, 1993), Q-balls (Coleman, 1985; Microwave cavity searches for dark-matter axions were
Kusenko and Steinhardt, 2001), solitons, and clumps reviewed by Bradley et al. (2003). The first experiment
formed due to dissipative interactions in the DM sec- to search for light DM composed of QCD axions was the
tor. Alternatively, a significant fraction of the DM mass- Axion DM eXperiment (ADMX), which began its work
energy could be stored in “topological defects” manifest- in the 1990s (Asztalos et al., 2001, 2010). This exper-
ing as monopoles, strings, or domain walls (Vilenkin, iment exploits the coupling of the QCD axion to the
1985). If DM takes such a form, terrestrial detectors electromagnetic field to convert axions into microwave
would not register a continuous oscillating signal as- photons in a strong magnetic field B (Fig. 19). In gen-
sociated with VULFs, but rather would observe tran- eral, pseudoscalar particles such as axions and ALPs can
sient events associated with the passage of the detec- be produced by the interaction of two photons via a pro-
tor through such a DM object (Budker and Derevianko, cess known as the Primakoff effect (Primakoff, 1951), and
2015; Derevianko and Pospelov, 2014; Pospelov et al., consequently an axion/ALP interacting with an electro-
2013). Self-interacting fields can include bosonic and magnetic field can produce a photon via the inverse Pri-
fermionic DM candidates. The characteristic spatial ex- makoff effect (Raffelt and Seckel, 1988). This process was
tent of topological defects is determined by the Compton proposed by Sikivie (1983, 1985) as a method to search
wavelength of the underlying DM field. For an Earth- for DM axions (haloscope experiment) as well as axions
sized object, the characteristic mass is ∼ 10−14 eV which emitted by the Sun (helioscope experiment). The non-
places such DM fields in the category of ultralight candi- relativistic Lagrangian describing this interaction is
dates. α a(r, t)
Laγγ = gγ E ·B , (88)
π fa
B. Experimental Searches where gγ is the coupling constant describing the strength
of the axion-photon interaction, α is the fine structure
Axion and ALP searches can be classified in different constant, a(r, t) is the axion field, fa is the symmetry
categories depending on where the detected particles are breaking scale associated with the axion (see Sec. VII.C),
produced. For example, in “light shining through walls” and E and B are the electric and magnetic fields. This
(LSW) experiments (Redondo and Ringwald, 2011; Ro- interaction corresponds to the axion-E&M entry in Ta-
billiard et al., 2007), axions are created in the experiment ble II. In the ADMX experiment, B is generated with
67

space corresponding to standard QCD axion models,


namely the Kim-Shifman-Vainshtein-Zakharov (KSVZ)
and Dine-Fischler-Srednicki-Zhitnitskii (DFSZ) family of
models (Dine et al., 1981; Kim, 1979; Shifman et al.,
1980; Zhitnitskii, 1980), in a band of axion masses near
≈ 2 − 4 × 10−6 eV. A new effort to extend the ADMX
experiment to search for higher mass axions using corre-
spondingly higher frequency microwave cavities, known
as HAYSTAC – Haloscope At Yale Sensitive To Axion
Cold dark matter (van Bibber and Carosi, 2013), has re-
cently produced its first results (Brubaker et al., 2016).
HAYSTAC was able to probe higher axion masses with
improved sensitivity by pushing to lower temperatures
and leveraging recent progress in quantum electronics;
HAYSTAC has probed the KSVZ parameter space in a
band of axion masses near ≈ 24 × 10−6 eV (Brubaker
et al., 2016). The ADMX and HAYSTAC collaborations
plan a systematic search for QCD axions with masses
between ma c2 ≈ 10−6 eV and ≈ 10−4 eV by 2021.
Another significant microwave cavity experimental
program is underway at the Center for Axion and Preci-
FIG. 19 Schematic diagram of the ADMX experiment from sion Physics Research (CAPP) at KAIST in South Korea
Asztalos et al. (2001). Photons produced in the microwave (Semertzidis, 2017; Youn, 2016), where researchers are
cavity by the interaction of an axion field a(r, t) with the developing stronger magnets, new low-noise amplifiers
magnetic field B [Eq. (88)] are detected by the electric-field (e.g., based on Josephson parametric amplifiers), and
probes. Tuning rods enable the resonant frequency of the superconducting cavities with novel designs to increase
cavity to be scanned to search for axions of different masses.
their Q and expand their volume. The CAPP haloscope,
(Fields from the RF signal source can be sent through the
setup for calibration purposes.) The signals are recorded after known as CULTASK (CAPP’s Ultra Low Temperature
multiple amplification stages and heterodyning. The 2001 ex- Axion Search in Korea), aims to target an axion mass
periment employed cryogenic heterojunction field-effect tran- range near ≈ 10−5 eV.
sistors built by the National Radio Astronomy Observatory A new broadband axion DM haloscope experiment
(NRAO), while new versions of ADMX employ SQUID am- aimed at detecting axions with ma ≈ 10−4 eV proposed
plifiers (Asztalos et al., 2010). by Jaeckel and Redondo (2013) is under development at
the Max Planck Institute for Physics (Majorovits et al.,
2016). This project, named MADMAX, is based on
a superconducting solenoid and E is the electric field of
axion-photon conversion at the transition between dif-
the resultant microwave photon produced by the inverse
ferent dielectric media. By using ∼ 80 dielectric discs
Primakoff effect. The resonant frequency of the cavity
immersed in a ∼ 10 T magnetic field, the emitted power
can be tuned so that it matches the frequency of the mi-
is enhanced by a factor of ∼ 105 over that from a single
crowave photons produced by the interaction of a(r, t)
mirror (flat dish antenna).
with B, which have energy corresponding to
1
Eγ ≈ ma c2 + ma c2 β 2 , (89) 2. Spin-precession axion experiments
2
where ma is the axion mass and β = v/c ∼ 10−3 is the A new experiment recently proposed by Budker et al.
relative velocity of the laboratory with respect to the rest (2014) to search for lighter QCD axions and ALPs us-
frame of the axion field. As noted in the introduction ing different couplings from those exploited in ADMX
to this section, the dispersion of the axion velocities is and similar microwave cavity experiments is the Cosmic
roughly on the same order as β, i.e., ∆β ∼ 10−3 , so the Axion Spin Precession Experiment (CASPEr). CASPEr
axionic DM would produce a relatively narrow microwave exploits both the axion-gluon coupling, which generates
resonance: a time-varying electric dipole moment (EDM) of nuclei6
(CASPEr Electric), and the coupling of the axion to
∆ω 2
∼ (∆β) ∼ 10−6 . (90)
ω
ADMX is to date the first and only experiment to 6 As noted by Schiff (1963), the interaction of an EDM of a point-
probe the particularly interesting region of parameter like particle with an applied electrostatic field is screened in
68

externally-applied bias magnetic field and search for a


non-zero magnetometer response, which is a signature of
spin coupling to the axion DM field. CASPEr Electric
has the potential to reach sensitivity to QCD axions over
a mass range of 10−14 eV < 2 < −9
∼ ma c ∼ 10 eV and search
a significant fraction of unexplored parameter space for
ALPs up to masses of ≈ 10−7 eV (Budker et al., 2014).
CASPEr Wind is an example of an experiment specif-
ically sensitive to ALP DM (at least in its present form,
it will not have sufficient sensitivity to reach parameter
FIG. 20 Experimental concept of CASPEr (Budker et al.,
2014). The oscillating axion field a(r, t) acts as a pseudo- space corresponding to the QCD axion). CASPEr Wind
magnetic field B1∗ , either by inducing an oscillating electric is analogous to CASPEr Electric, except that the pseudo-
dipole moment (EDM) via the axion-gluon interaction that magnetic field B1∗ is generated by a different mechanism:
couples to an electric field (CASPEr Electric), or via the in- the coupling of nuclear spins to the spatial gradient of
teraction of spins with the gradient of a(r, t) arising from the the ALP DM field (the so-called “ALP wind”). This en-
motion of the sample through the axion field (CASPEr Wind). ables the use of materials such as liquid xenon without
The oscillating B1∗ causes polarized nuclear spins to tip away
electric fields. Xenon can be efficiently spin-polarized to
from the leading field B0 and precess at the Larmor frequency.
The approach is based on the principles of NMR experiments. enhance the signal.
In the KSVZ family of QCD axion models (Kim, 1979;
Shifman et al., 1980), the coupling of the axion to elec-
nuclear spins (CASPEr Wind), see Graham and Rajen- tron spins is nominally zero, whereas in the DFSZ family
dran (2013). CASPEr uses nuclear magnetic resonance of models (Dine et al., 1981; Zhitnitskii, 1980), the ax-
(NMR) techniques for detecting spin precession caused ion is predicted to couple to the electron spin. Thus,
by background axion DM. This approach complements in addition to searches for axion couplings to nuclear
ADMX, HAYSTAC, and CULTASK which are sensitive spins as searched for in CASPEr, it is of interest to
to higher axion masses and a different coupling. search for axion-electron couplings: this is the target of
The key idea underlying the CASPEr concept is that the QUAX (QUaerere AXion) experiment (Ruoso et al.,
axion DM can cause the precession of nuclear spins 2016). The essence of the experiment, originally out-
in a manner similar to that discussed for exotic spin- lined by Barbieri et al. (1989); Kakhidze and Kolokolov
dependent interactions and EDMs (Secs. VII and V, (1991); Krauss et al. (1985); Turner (1990), is quite simi-
respectively). Nuclear spins in a non-centrosymmetric lar to that of CASPEr, with the important difference that
crystal, such as a ferroelectric, experience a large effec- in the QUAX experiment a Yttrium Iron Garnet (YIG)
tive electric field (Leggett, 1978; Mukhamedjanov and sphere is used as the sample of polarized electron spins
Sushkov, 2005), a phenomenon analogous to the large as opposed to the polarized nuclear samples studied in
internal effective electric fields experienced by electrons CASPEr.
in polar molecules (Graham and Rajendran, 2011). The
coupling of the axion DM field to nuclear spins (via the
generation of electric dipole moments through the axion- 3. Radio axion searches
gluon coupling) in such a material has the form of a
pseudo-magnetic field B1∗ oscillating at the axion Comp- ADMX, CASPEr, and related experiments are also
ton frequency. If the external bias magnetic field B0 is sensitive to another class of particles known as dark
set to a value such that the nuclear spin splitting matches or hidden photons (Wagner et al., 2010), discussed in
this frequency, a resonance condition is achieved, and the Sec. VII.B.5. Like ordinary photons, hidden photons
nuclear spins and the corresponding magnetization M tilt are vector particles with spin 1. However, hidden pho-
and undergo Larmor precession (see Fig. 20). A sensitive tons have mass and could constitute the DM in a man-
magnetometer, such as a Superconducting Quantum In- ner similar to axions and ALPs (Arkani-Hamed et al.,
terference Device (SQUID), placed next to the sample, 2009a). Hidden-photon DM can be described as a weakly
detects the oscillating transverse magnetization. The ex- coupled “hidden electric field,” oscillating at the hidden-
perimental protocol of CASPEr-Electric is to sweep the photon Compton frequency, and able to penetrate shield-
ing (Jackson Kimball et al., 2016a). At low frequencies
(where the wavelength is long compared to the size of the
shielding), the interaction of electrons in the shielding
atomic systems, since the constituent charged particles redis-
tribute themselves to cancel the field. However, the screening is
material with the hidden-photon field generates a real,
incomplete because of finite nuclear size and relativistic effects, oscillating magnetic field. It has recently been proposed
which can even enhance the atomic EDM relative to the electron that such hidden-photon DM can be searched for using
or nuclear EDM in heavy atoms (see Sec. V). a tunable, resonant LC circuit designed to couple to this
69

magnetic field, a “dark matter radio” (Chaudhuri et al., renormalized atomic mass. For atomic interferometers,
2015). Hidden-photon DM has an enormous range of the effects of atomic mass variation during the interfero-
possible Compton frequencies, but current experiments metric sequence and also DM-induced renormalization of
(such as ADMX, which is also sensitive to hidden pho- the local gravity g dominate over the direct DM-induced
tons) search only over a few narrow parts of that range forces (Geraci and Derevianko, 2016). Accelerometers
(Wagner et al., 2010). In contrast, DM Radio has poten- are particularly sensitive to vector and scalar VULFs.
tial sensitivity many orders of magnitude beyond current
limits over an extensive range of hidden photon masses,
from 10−12 eV < 2 <
∼ mγ 0 c ∼ 10
−3
eV, where mγ 0 is the 5. Exotic spin-dependent forces due to axions/ALPs
hidden photon mass.
Related proposals for broadband axion/ALP detection Section VII explores the exotic spin-dependent interac-
with LC circuits were developed by Sikivie et al. (2014) tions generated by axions, ALPs, and dark/hidden pho-
and Kahn et al. (2016). The concept of these experiments tons. A recent proposal by Arvanitaki and Geraci (2014)
can be understood by noting that the axion-photon cou- to search for short-range monopole-dipole interactions
pling effectively modifies Maxwell’s equations (Sikivie, between nuclei using NMR techniques, the Axion Res-
1983, 1985) such that dark-matter axions/ALPs generate onant InterAction Detection Experiment (ARIADNE),
an oscillating current density in the presence of a mag- has particular relevance to axion DM searches. The aim
netic field. These ideas also apply to the high-frequency of ARIADNE is to detect monopole-dipole interactions
(10-100 GHz) axion search proposed by the MADMAX between the spins of 3 He nuclei and a rotating unpolar-
collaboration (Jaeckel and Redondo, 2013). ized tungsten attractor. The geometry of the experiment
is specially designed to be sensitive to QCD axions in
the range 10−6 eV < 2 < −3
∼ ma c ∼ 10 eV. The upper end of
4. Atomic clocks and accelerometers, and spectroscopy the axion mass range to be explored by ARIADNE, well
within the astrophysically and cosmologically allowed re-
As noted in the introduction to this section, one of the gion, is particularly difficult for DM detection experi-
generic signals VULFs can produce are time-oscillating ments such as ADMX and CASPEr to access, and so
interactions. An example is DM consisting of dilatons, ARIADNE has the potential to fill in an important gap
ultralight scalar particles arising in string theories (Ar- in the explored axion parameter space.
vanitaki et al., 2015). Like axions and ALPs, dilatons
form a gas described as a scalar field oscillating at the
Compton frequency of the dilaton. This field feebly in- 6. Magnetometer and clock networks for detection of transient
teracts with normal matter leading to temporal varia- dark matter signals
tion of fundamental “constants” which in turn affects the
“ticking” rates of atomic clocks. Since clocks based on If a detection the QCD axion or other VULF candi-
distinct atomic transitions have different sensitivities to a dates is made, a network of such experiments can be
change of constants such as the fine structure constant α, used to verify it, since the signal in all of them should be
comparisons between such clocks is a sensitive way to de- centered at the axion/ALP Compton frequency, a fun-
tect the variation of the constants (see Sec. II), including damental constant. A network would also enable the
those caused by a time-varying DM field. A “differential study of spatial coherence of the DM field (Derevianko,
atomic clock” based on microwave transitions between 2016) and search for deviations from the standard halo
nearly-degenerate metastable states in dysprosium was model predictions due to non-uniform/non-isotropic DM
used by Van Tilburg et al. (2015) to search for dilatons flows (Duffy and Sikivie, 2008). Networks of sensors are
in the mass range of 10−24 to 10−16 eV, improving exist- crucial in order to search for “clumpy” DM. Here the
ing constraint on the electron coupling of a DM dilaton searches rely on the characteristic time delay of DM-
by up to four orders of magnitude. These limits were induced signals between the nodes (see Fig. 21), as on
further improved by Hees et al. (2016). Modern atomic average the “clumps” would sweep the network at galac-
clocks based on single trapped ions (Huntemann et al., tic velocities.
2016) and ensembles of neutral atoms in optical lattices The Global Network of Optical Magnetometers to
(Nemitz et al., 2016) are reaching into relative frequency search for Exotic physics (GNOME) collaboration
instability levels of a part in 1018 , promising a boost in (Pospelov et al., 2013; Pustelny et al., 2013) is search-
the sensitivity of dilaton searches by about two orders of ing for such transient signals due to passage of the Earth
magnitude in the near future. through DM objects that couple to atomic spins (simi-
Also of note is that recently Graham et al. (2016b) lar to the ALP wind coupling searched for by CASPEr).
proposed using using accelerometers (e.g., torsion bal- While a single magnetometer system could detect such
ances and atom interferometers) to search for DM- transient events, it would be exceedingly difficult to
induced forces, −∇[Ma (r, t)c2 ], where Ma is the DM- confidently distinguish a true signal generated by light
70

eter Gravitational Wave Observatory (LIGO) collabora-


tion has developed sophisticated data analysis techniques
12 12

9
10
11 1
2

3
vg
9
10
11 1
2

3
to search for similar correlated “burst” signals from a
8
7
6
5
4 8
7
6
5
4

worldwide network of gravitational wave detectors (Allen


et al., 2012; Anderson et al., 2001), and the GNOME
difference in clock readings

collaboration has demonstrated that these data analy-


sis techniques can be adapted to analyze data from the
GNOME (Pustelny et al., 2013). Presently the GNOME
consists of 5-10 dedicated atomic magnetometers located
at stations throughout the world.
If DM leads to variation of fundamental constants, DM
“clumps” can also manifest themselves as “glitches” of
l /v g
atomic clocks, for example those onboard satellites of
time Global Positioning System (Derevianko and Pospelov,
2014): if particular interactions exist, the clocks would
FIG. 21 Spatially-separated and initially-synchronized iden- become desynchronized as they are swept by a DM ob-
tical clocks are expected to exhibit a distinct de- ject (Fig. 21). The glitches would propagate through
synchronization and re-synchronization pattern due to an en- the GPS constellation at galactic velocities, ∼ 300 km/s,
counter with a DM object. Two clocks are separated by a dis- characteristic of DM halo. The GPS.DM collaboration
tance `, and because the wall propagates through the network is mining over a decade of archival GPS data to hunt for
with a speed vg ∼ 300 km/s, the characteristic “hump” per- such DM objects, effectively using the GPS constellation
sists over time `/vg . Adopted from Derevianko and Pospelov
(2014).
as a 50,000 km-aperture DM detector. While the ini-
tial search has not found evidence for such DM objects
at the current search sensitivity, it improves the current
constraints on certain DM couplings by several orders of
magnitude (Roberts et al., 2017). Recently, it has been
also shown that a single optical atomic clock (composed
of two independent clock ensembles probed by the same
laser) can be sensitive to transient DM interactions and
constraints on scalar quadratic DM-clock couplings have
been obtained by Wcislo et al. (2016). While this co-
located clock technique can be used to place limits on
DM-clock couplings, the positive DM detection still re-
quires a geographically-distributed network. Transient
variations of fundamental constants can be also searched
for with a global network of laser interferometers (Stad-
nik and Flambaum, 2015b, 2016).

X. GENERAL RELATIVITY AND GRAVITATION

FIG. 22 Schematic figure of an axion/ALP domain-wall A. Tests of the Einstein equivalence principle
crossing event as searched for by the GNOME; figure from
Pospelov et al. (2013). The crossings recorded in four distinct
locations (marked with stars) allow determination of the nor- The equivalence principle can be traced back to the
mal velocity v⊥ to the wall and prediction of the timing of sixteenth-century observation that all bodies fall to Earth
the 5th event. at the same rate of acceleration (Will, 2014). This was a
remarkable discovery, for it leads to the conclusion that a
body’s mass is proportional to its weight. The constant
DM from “false positives” induced by occasional abrupt of proportionality seems to be independent of material
changes of magnetometer operational conditions (e.g., composition or any other detail of the body. That is the
magnetic-field spikes, laser-light-mode jumps, electronic basic principle of the equivalence of gravitational mass
noise, etc.). Effective vetoing of false positive events re- and inertial mass.
quires an array of magnetometers. Furthermore, there Within the framework of Einstein’s theory of general
are key benefits in terms of noise suppression and event relativity (GR), there is the Einstein equivalence princi-
characterization to widely distributing the magnetome- ple (EEP), which includes the following postulates (Will,
ters geographically; see Fig. 22. The Laser Interferom- 2014):
71

1. The weak equivalence principle (WEP): the trajec- The “Eöt-Wash” torsion-balance experiments tested
tory of a freely falling “test” body is independent of WEP to 10−13 by comparing differential accelerations
its internal structure and composition. All bodies of beryllium-aluminum and beryllium-titanium test-body
in a common gravitational field fall with the same pairs (Schlamminger et al., 2008; Wagner et al., 2012).
acceleration according to WEP. This is also called Lunar laser-ranging experiments, which measure the dif-
the universality of free fall (UFF). ferential accelerations of the Earth and Moon towards the
Sun, provided similarly stringent limits to the violation
2. Local Lorentz invariance (LLI): the outcome of
of the equivalence principle (Williams et al., 2004, 2012).
any local non-gravitational experiment conducted
Both the torsion-balance and lunar-laser ranging WEP
in free fall is independent of the velocity and the
tests are close to their fundamental limits of accuracy.
orientation of the apparatus.
Macroscopic (i.e., classical) Earth-based free fall WEP
3. Local position invariance (LPI): the outcome of any tests are less accurate, reaching the 10−10 level (Kuroda
local non-gravitational experiment is independent and Mio, 1989).
of where and when in the universe it is performed. Significant improvement in probing WEP is expected
to come from future space-based missions. MicroSCOPE
Different versions of the EEP appear in the literature; is a Centre National d’Études Spatiales (CNES)/ Euro-
precise formulations of its variants are discussed by Ca- pean Space Agency (ESA) gravity-research minisatellite
sola et al. (2015). AMO tests of LLI are discussed in mission (Bergé et al., 2015) that aims to test the WEP in
Sec. XI. Tests of LPI include searches of the temporal space to 10−15 by comparing the acceleration experienced
and spatial variation of fundamental constants, as dis- by two free-falling test masses in the Earth’s gravity field.
cussed in detail in Sec. II. The satellite was launched in April 2016 and the mission
Both GR and the SM are assumed to be low-energy is planned to last two years. Other macroscopic propos-
limits of a more complete theory at the high-energy scale. als to test the WEP include the Sounding Rocket Prin-
The EEP implies a universal coupling between matter ciple of Equivalence Measurement [free-fall stratosphere
and gravity, i.e. all forms of matter-energy respond to experiment], Galileo Galilei (Nobili et al., 2012) [space-
gravity in the same way. However, this may not be the based torsion-balance experiment], the Satellite Test of
case for most theories aimed at unifying all four funda- the Equivalence principle (Overduin et al., 2012) [free-fall
mental interactions, such as string theories. Any theories space based experiment], and others.
in which the coupling constants are spatially dependent The theoretical framework for WEP tests in the quan-
violate the WEP, as discussed in Sec. II. On the other tum domain is discussed by Herrmann et al. (2012).
hand, the WEP can be tested with experiments comple- Weak-equivalence tests using quantum matter were made
mentary to those used to test fundamental constant vari- possible by techniques for production and control of ul-
ation. Thus WEP tests provide additional opportunities tracold atoms and by the attainment of dilute atomic
to open a low-energy window into the nature of unifi- Bose-Einstein condensates (BECs) (Cornell and Wie-
cation theories. Various theoretical arguments that the man, 2002; Ketterle, 2002). Atom interferometers mea-
EEP is violated at small but measurable levels are dis- suring the difference in phase between matter waves trav-
cussed in detail by Damour (2012). “Runaway dilaton” eling along different paths can be used as accelerometers,
models (Damour, 2012; Damour et al., 2002; Damour and offering potential precision tests of GR with quantum
Polyakov, 1994), estimate that onset of WEP violation rather than classical matter. A good review of atom in-
may start just beyond the sensitivity of current experi- terferometry is given by Cronin et al. (2009). Testing
ments. the limits of quantum mechanical superpositions with dif-
Tests of the equivalence principle as well as more gen- ferent systems has been discussed by (Arndt and Horn-
eral tests of gravity are reviewed by Will (2014). A review berger, 2014).
of the past WEP tests and future proposals is given by
The most accurate WEP test using both quantum and
Sondag and Dittus (2016). Modern tests of the WEP in-
classical objects was reported by Peters et al. (1999).
clude torsion-balance experiments, free-fall experiments
This experiment compared the values for the local accel-
and measurement of relative motions of celestial bodies
eration due to the Earth’s gravity, g, obtained using an
(for example, lunar laser ranging). To quantify viola-
atom interferometer based on a fountain of 133 Cs laser-
tions of WEP, we suppose that the gravitational mass of
cooled atoms and a Michelson-interferometer classical
a body, mg , is not equal to its inertial mass, mI . Then,
gravimeter which used macroscopic glass object, demon-
the acceleration a of a body in a gravitational field g is
m strating agreement to 7 parts in 109 .
given by a = mgI g. To test WEP, one compares the accel-
Fully quantum WEP tests with atomic interferometry
erations a1 , a2 of two falling bodies which differ in their
directly compare the phase shifts of two different types
composition, and measures the “Eötvös” ratio η
of matter waves without the use of classical gravimeters.
a1 − a2 The potential of matter-wave interferometers using quan-
η = 2 . (91)
a1 + a2 tum gases for probing fundamental concepts of quantum
72

mechanics and GR has been discussed by Biedermann from either difference of the inertial and gravitational
et al. (2015); Dimopoulos et al. (2007); Herrmann et al. masses or an additional (unknown) force which depends
(2012); Müntinga et al. (2013). We refer the reader to on the composition of the atomic species and differs for
87
Cronin et al. (2009) for a review of atom interferometry Rb and 39 K atoms. The same apparatus may be used
and its applications. to improve the precision by two orders of magnitudes, to
The sensitivity of atom interferometers to WEP vi- the ppb level.
olations increases linearly with the momentum differ- A scheme to suppress common-mode noise in lasers
ence between the two matter waves emerging from a used for atom interferometry was demonstrated by
beam splitter and quadratically with the time of free Zhou et al. (2015), resulting in measurement of
fall. Sensitivity can be increased by increasing the mo- η(87 Rb,85 Rb) = (2.8 ± 3.0) × 10−8 .
mentum difference or the time in free fall (or both). One of the advantages for using cold atom clouds
Therefore, the space-based experiments promise a break- for gravity tests is the opportunity to perform qualita-
through in sensitivity because of long free-fall times. Cur- tively different WEP tests with well-characterised “test
rent and proposed tests of gravity with atom interferome- masses” with a definite spin for a search of the spin-
try include splitting free-falling BECs in atomic fountains gravity coupling effects. Tarallo et al. (2014) reported
Hartwig et al. (2015); Schlippert et al. (2014), drop tow- such an experimental comparison of the gravitational in-
ers Müntinga et al. (2013), parabolic flights (Seidel et al., teraction for a 88 Sr (boson, I = 0) with that of a 87 Sr
2015), and outer space (Tino et al., 2013; Williams et al., (fermion, I = 9/2). The Eötvös ratio and possible spin-
2016). gravity coupling were constrained at the 10−7 level. Note
In 2010, a preparation and observation of a Bose- that such a test is completely insensitive to the types of
Einstein condensate during free fall in a 146-meter-tall spin-gravity interactions probed in spin-precession exper-
evacuated drop tower of the Center of Applied Space iments such as those of Venema et al. (1992) and Jack-
Technology and Microgravity (ZARM) in Bremen, Ger- son Kimball et al. (2017); see the discussion in Sec. VII.G.
many was reported (van Zoest et al., 2010). The realiza- Duan et al. (2016) reported a test of the universality
tion of an asymmetric Mach-Zehnder interferometer op- of free fall with 87 Rb atoms in different spin orientations.
erated with a Bose-Einstein condensate in extended free They used a Mach-Zehnder-type atom interferometer to
fall at ZARM was reported by Müntinga et al. (2013). alternately measure the free fall acceleration of the atoms
These proof-of-principle experiments demonstrated a fea- in mF = +1 and mF = −1 magnetic sublevels, with the
sibility of coherent matter-wave experiments in micro- resultant Eötvös ratio of η = (0.2 ± 1.2) × 107 . Hartwig
gravity paving the way toward for matter-wave exper- et al. (2015) proposed a long baseline atom interferome-
iments in space. In 2017, the QUANTUS collabora- ter test of EEP with Rb and Yb. With over 10 meters of
tion (Seidel et al., 2015) conducted successful MAIUS free fall, their experiment is estimated to reach 7 × 10−13
1 (Matter-Wave Interferometry in Microgravity) experi- accuracy in the Eötvös ratio. Use of the heavy alkaline
ment aboard a sounding rocket at altitude up to 243 km earth Yb will broaden the scope of atom interferomet-
above the Earth’s surface, well above the Kármán line ric EEP tests in view of EEP violation parametrization
that marks the boundary of outer space (the Interna- based on the dilaton model described by Damour (2012).
tional Space Station’s orbit is about 400 km above the A number of quantum WEP tests in microgravity are
surface of the Earth). About 100 discrete matter-wave being pursued, with the promise of greatly increased
experiments were conducted during the six-minute ex- precision over current quantum tests. This is a goal
perimental phase of this flight. of the QUANTUS collaboration mentioned above. The
Atom-interferometry quantum tests of the universality I.C.E. (Interférometrie atomique á sources Cohérentes
of free fall with cold rubidium 85 Rb and 87 Rb atoms were pour l’Espace - Coherent atom interferometry for space
performed by Fray et al. (2004) and Bonnin et al. (2013) applications) experiment (Geiger et al., 2011) is a com-
at the 10−7 level. The first quantum test of the UFF pact and transportable atom interferometer, designed to
with matter waves of two different atomic species was re- test WEP by comparing the accelerations of free-falling
ported by Schlippert et al. (2014). This experiment com- clouds of ultracold Rb and K atoms inside an airplane in
pared the free-fall accelerations of laser-cooled ensembles free fall. Searching for WEP violation at high-precision
of 87 Rb and 39 K atoms by measuring the gravitationally is the primary science objectives of the SpaceTime Ex-
induced shift in two Mach-Zehnder-type interferometers. plorer and QUantum Equivalence Space Test (STE –
Schlippert et al. (2014) measured the Eötvös ratio QUEST) space mission (Altschul et al., 2015) designed
gRb − gK to measure the Eötvös ratio between matter waves of
ηRb,K = 2 = (0.3 ± 5.4) × 10−7 , (92) two Rb isotopes in a differential atom interferometer at
gRb + gK
the 2 × 10−15 uncertainty level. Although QUEST was
where gRb and gK are free-fall accelerations of the 87 Rb not selected for the European Space Agency M3 Cos-
and 39 K atoms, respectively. A non-zero value of the mic Vision Programme, it demonstrated the potential of
Eötvös ratio would indicate a UFF violation resulting future space-based quantum WEP tests. Other propos-
73

als for quantum atom interferometry space-based WEP eters. The gravity curvature was produced by nearby
tests include Quantum Test of the Equivalence principle source masses along one axis. In the experimental set
and Space Time (QTEST) (Williams et al., 2016) and a up designed by Rosi et al. (2015), three atomic clouds
Quantum WEP test (Q-WEP) (Tino et al., 2013) on the launched in the vertical direction are simultaneously in-
International Space Station. terrogated by the same atom interferometry sequence
probing the gravity field at three equally spaced posi-
tions. Such atomic sensor is capable to measure gravity,
B. Determination of the Newtonian gravitational constant gravity gradient, and curvature along the vertical direc-
tion at the same time, important for geodesy studies and
The Newtonian gravitational constant, G, was the sec- and Earth monitoring applications. The same scheme
ond fundamental constant subject to an absolute mea- may be used for a novel approach to G measurement.
surement, which was first conducted by Cavendish in
1797-98 (Cavendish, 1798). The 2014 CODATA recom-
mended value (Mohr et al., 2016) is G = 6.67408(31) × C. Detection of gravitational waves
10−11 m3 kg−1 s−2 . The relative standard uncertainty in
G of 4.7 × 10−5 is by far the largest of any of the primary The detection of the gravitational waves (GW) by the
fundamental constants, exceeding that of the Boltzmann Advanced LIGO in 2015 (Abbott et al., 2016a,b) ini-
constant, k, by two orders of magnitude. The slow rate of tiated the field of gravitational-wave astronomy. This
progress in reducing the uncertainty in G is a matter of opens a new window on the universe since many of the
concern in the precision-measurement community (An- GW cosmic sources do not have detectable electromag-
derson et al., 2015; Schlamminger, 2014; Schlamminger netic emissions. Theoretical physics implications of the
et al., 2015). Reflecting the isolation of gravity and GR observed binary black-hole mergers and probes of new
from the Standard Model, G is also unique among the physics and cosmology enabled by the detection of the
fundamental constants in having no dependence upon gravitational waves are described by Yunes et al. (2016).
any of the other constants included in the CODATA Once at full sensitivity in 2019, the Advanced LIGO de-
least-squares fit (Mohr et al., 2016). tectors will be able to see inspiralling binaries made up
Until 2014, all experimental determinations of G that of two 1.4 solar-mass neutron stars to a distance of 300
contributed to its CODATA recommended value involved megaparsecs (Mpc, 1 parsec = 3.3 light years) and coa-
measurement of classical forces. Fixler et al. (2007) lescing black-hole systems at the cosmological distance,
performed the first measurement of G using an atom- to the red shifts z = 0.4, significantly increasing the num-
interferometric gravity gradiometer. This demonstrated ber of potentially detectable events. Advanced LIGO at
the gravitational action of a laboratory source mass full capacity will be essentially operating at the quantum
upon an atomic de Broglie wavelength, an intrinsically noise limit. With the Advanced Virgo GW detector in
quantum-mechanical effect [the first experimental obser- Italy coming online in 2017 along with future detectors,
vation of a gravitational shift of a de Broglie wave was the GW signals may be triangulated. There are already
made in 1974 using neutron interferometry in the Earth’s proposals for 10 km and 40 km laser interferometers, the
gravitational field (Colella et al., 1975)]. This experiment Einstein telescope (Sathyaprakash et al., 2012) and the
yielded a value of G consistent with the CODATA recom- Cosmic explorer (Abbott et al., 2017), significantly longer
mended value then, but with a larger uncertainty. Atom- than the Advanced LIGO 4 km arms, and thus able to
interferometric measurements of G certainly offer the measure lower frequencies at smaller fractional sensitiv-
prospect of having systematic uncertainties qualitatively ity.
different from those of classical experiments, and they The detection capability of the laser interferometry
may eventually provide a link between G and other fun- terrestrial detectors is limited to GWs with frequencies
damental constants, as in the example of the dependence above ∼ 10 Hz by the seismic noise and Newtonian noise
of the fine structure constant, α, on the ratio ~/M (87 Rb). (fluctuations of the terrestrial gravity field which creates
Rosi et al. (2014) reported an atom-interferometric deter- a tidal effect on separated test masses) (Saulson, 1984).
mination of G was with a relative standard uncertainty The ability to detect gravitational waves of lower fre-
of 0.015%. Their measurement was included in the least- quencies will significantly increase the number of binary
squares fit that determined 2014 CODATA recommended star mergers from which the gravitational waves may be
value. Past and ongoing G determinations based on atom detected and allow for longer observation of the inspi-
interferometry are reviewed by Rosi (2016). ralling binary stars before the merger. Stochastic gravi-
Use of multiple atomic samples in an interferometer tational waves, i.e. relic GWs from the early evolution of
also enables measurements of higher-order spatial deriva- the universe, from cosmological (and possibly unforeseen)
tives of the gravity field. In 2015, Rosi et al. (2015) re- sources, such as inflation and reheating, a network of cos-
ported the first direct measurement of the gravity-field mic strings, or phase transitions in the early Universe,
curvature based on three conjugated atom interferom- etc., can also be easier to detect at these low frequen-
74

cies, see Dimopoulos et al. (2008b). Proposals for the driving atomic single-photon transitions. Use of single-
detection of the gravitational waves at lower frequencies photon transitions in alkaline-earth atoms (Sr) with long
include space-based laser interferometry detector [Laser lifetimes of the excited state significantly reduces laser
Interferometer Space Antenna (LISA) (Danzmann et al., frequency noise (Graham et al., 2013) in comparison to
2011) and Evolved-LISA (eLISA) (Amaro-Seoane et al., the GW detector proposals based on alkali-metal atoms
2012)] and both terrestrial and space based matter-wave requiring implementation of two-photon transitions (Di-
detectors, using either atom interferometers or atomic mopoulos et al., 2008b). The GWs are detected by mon-
clocks (Chaibi et al., 2016; Chiow et al., 2015; Dimopou- itoring the phase difference between the two interferome-
los et al., 2008b; Graham et al., 2013, 2016a; Hogan et al., ters caused by the variation of the light travel time across
2011; Hogan and Kasevich, 2016; Kolkowitz et al., 2016; the baseline due to a passing GW. As described by Gra-
Yu and Tinto, 2011). ham et al. (2016a), the atom interferometric GW detec-
Terrestrial atom interferometers have been proposed tor essentially compares time kept by the laser and atom
(Chaibi et al., 2016; Dimopoulos et al., 2008b) for GW “clocks.“ A gravitational wave affects the flat-space re-
detection in the 0.3 - 3 Hz frequency band. Chaibi et al. lation between these clocks by a factor proportional to
(2016) proposed a new detection strategy based on a cor- the distance between them, and such change oscillates in
related array of atom interferometers which allows re- time with the frequency of the gravitational wave result-
duction of the Newtonian noise limiting all ground based ing in a detectable signal.
GW detectors below ∼ 10 Hz. The matter-wave laser In previous proposals (Dimopoulos et al., 2008b; Gra-
interferometer gravitation antenna (MIGA) is a hybrid ham et al., 2013), the same laser would drive atom tran-
detector that couples laser and matter-wave interferome- sitions at both ends, requiring the laser to remain colli-
try aimed at both the geophysical studies and sub-hertz mated over the optical path between two satellites, sig-
GWs detection in a low-noise underground laboratory to nificantly restricting the maximum baseline length. In
minimise effects of laboratory vibrations. (Geiger et al., the Hogan and Kasevich (2016) scheme, both satellites
2016). house a master laser and a local-oscillator laser that have
LISA/eLISA are space-based laser interferometric de- sufficient intensity to drive transitions in the local atom
tectors analogous to LIGO, to be composed of three interferometer. The master laser beam interacts with
spacecraft forming either a two- or three-arm Michelson its satellite’s atomic cloud, and then propagates to the
interferometer (Amaro-Seoane et al., 2012; Danzmann second satellite acting now as a reference beam which
et al., 2011), with the GW frequency-detection range does not have to be collimated as it reaches the opposite
from 0.1 mHz to 1 Hz. satellite. A local oscillator in the other satellite is phase
Both atomic clocks and atom interferometry technolo- referenced or phase locked to the incoming reference laser
gies improved tremendously over the past decade, also beam and drives the transitions in this satellite’s atomic
leading to fast development of the gravitational wave de- cloud. An identical scheme is implemented in the re-
tection schemes (Chaibi et al., 2016; Chiow et al., 2015; verse direction. Since much less intensity of the reference
Dimopoulos et al., 2008b; Graham et al., 2013, 2016a; beam is required for phase reference than for atomic exci-
Hogan et al., 2011; Hogan and Kasevich, 2016; Kolkowitz tations, this scheme allows for much larger baseline lead-
et al., 2016) with improved sensitivities, realistic require- ing to enhanced sensitivity, simplified atom optics, and
ments for the underlying technologies, and addressing reduced atomic-source flux requirements. Such a GW
some problems (Bender, 2011; Bender, 2014) of earlier detector scheme with 12 photon recoil atom optics and
proposals. The intrinsic noise sources and sensitivity lim- 6 × 108 m baseline is evaluated to exceed the sensitivity
its of atom-based vs. light-based interferometers for GW of proposed the LISA detector by a factor of ten (Hogan
detection are being clarified (Baker and Thorpe, 2012; and Kasevich, 2016).
Bender, 2011; Bender, 2014). Most importantly, the re- Graham et al. (2016a) proposed an atom interferomet-
striction of a space-based atom interferometry GW de- ric GW detector that can operate in a resonant detec-
tector to a relatively short baseline, ∼ 1000 km, in com- tion mode and can switch between the broadband and
parison with LISA, has been lifted in the most recent narrowband detection modes to increase sensitivity.
2016 proposal (Hogan and Kasevich, 2016). Due to con- Kolkowitz et al. (2016) proposed to use a two-satellite
siderable evolution of the AMO GW detector proposals, scheme sharing ultrastable optical laser light over a sin-
we very briefly describe only the most recent proposals gle baseline, with atomic optical lattice clocks (rather
based on atom interferometers and atomic clocks. than atom interferometers) as sensitive, narrowband de-
Hogan and Kasevich (2016) proposed a space-based tectors of the local frequency of the shared laser light. A
GW detector based on two satellites with light-pulse passing GW induces effective Doppler shifts and the GW
atom interferometers separated by a long baseline (over signal is detected as a differential frequency shift of the
100,000 km), capable of detecting GWs in the 0.1-mHz shared laser light due to the relative velocity of the satel-
to 1-Hz frequency band. The light pulses are sent back lites. Such a scheme can detect GWs with frequencies
and forth across the baseline from alternating directions, ranging from 3 mHz to 10 Hz without loss of sensitiv-
75

ity. The clock scheme may be integrated with an optical was considered in (Hohensee et al., 2013b), letting the
interferometric detector. The next stage of matter-wave limits on equivalence principle violations in antimatter
gravitational detector development is a demonstration of from tests using bound systems of normal matter. We
ground-based prototype systems and characterization of emphasize that any difference between matter and an-
the noise sources. timatter gravity would run into theoretical conceptual
troubles (Karshenboim, 2016).
Furthermore, as discussed in previous sections, there
D. Gravity experiments with antimatter have been numerous and stringent matter-based tests
of the equivalence principle and CP T invariance, and
One of the recent foci of the experimental efforts on these must have a direct bearing on the proposed tests
matter-antimatter comparisons is testing whether anti- with antimatter, especially considering that most of the
matter is affected by gravity in the same way as matter. mass of the antiproton comes from the quark binding
For example, the CERN based GBAR collaboration is energy (i.e., the gluon field). The “true” antimatter
developing an ingenious technique (Pérez et al., 2015) mass/energy content of the antiproton in the form of an-
where they will first create a positive ion consisting of tiquarks can reasonably be assumed to be less than or
an antiproton and two positrons that will be sympathet- about 1%, while the antimatter content of ordinary mat-
ically cooled with Be+ ions and then “gently” photoion- ter due to virtual particles is non-negligible. This implies
ize to produce a cold neutral antihydrogen atom that will that there is a connection between matter-based equiv-
fall under gravity over a known distance before being de- alence principle tests and proposed antihydrogen experi-
tected. The AEgIS experiment at CERN is also aimed at ments. Another compelling, albeit model-dependent, ar-
the direct measurement of the Earth’s gravitational ac- gument (Adelberger et al., 1991) limiting the difference
celeration on antihydrogen but has a completely different between gravitational acceleration for matter and anti-
design (Testera et al., 2015). In the AEgIS expariment, matter to less than a part in 105 or perhaps much bet-
a cold, pulsed beam of antihydrogen will pass through a ter (considering the considerably more sensitive updated
moiré deflectometer (Aghion et al., 2014), coupled to a versions of the torsion pendulum experiment) is based
position-sensitive detector to measure the strength of the on reasonable assumptions of equivalence principle viola-
gravitational interaction between matter and antimatter tions arising from a scalar/vector gravitational coupling
to a precision of 1%. A second goal of the AEgIS exper- and combining data from the exquisitely precise measure-
iment is to carry out spectroscopic measurements on an- ments using torsion pendulums with stringent limits on
tihydrogen atoms. There is a possibility of laser cooling CP T invariance.
of the negative ion of lanthanum (La, Z = 57) (Jordan
et al., 2015). Laser-cooled La− might be used for sympa-
thetic cooling of antiprotons for subsequent antihydrogen E. Other AMO tests of gravity
formation (Kellerbauer et al., 2009).
A method to directly measure the ratio of the gravi- A test of the local Lorentz invariance of post-
tational to inertial mass of antimatter accomplished by Newtonian gravity was performed by (Müller et al., 2008)
searching for the free fall (or rise) of ground-state anti- by monitoring Earth’s gravity with a Mach-Zehnder atom
hydrogen atoms was proposed by ALPHA collaboration interferometer (see Sec.XI). Hohensee et al. (2012) pro-
(Amole et al., 2013). The antihydrogen atoms are re- posed an experimental realization of the gravitational
leased from the trap; the escaping anti-atoms are then Aharonov-Bohm effect: measurement of phase shifts with
detected when they annihilate on the trap wall. an atom interferometer due to a gravitational potential
One should emphasize that the possibility of a differ- U in the absence of a gravitational force. A pair of labo-
ence in gravity between matter and antimatter is already ratory masses will be used as a source of the gravitational
constrained, under some assumptions, at the 1 ppm level potential. A matter-wave interferometry experiment to
by experiments of Gabrielse et al. (1999) which found measure such phase shifts in the absence of a classical
no differences in gravitational red shift of matter and force is currently under construction at the University of
antimatter clocks. Ulmer et al. (2015) interpreted their California, Berkeley.
result for the sidereal variation of the q/m ratios given Testing sub-gravitational forces on atoms from a
by Eq. (49) as a test of the weak equivalence principle for miniature, in-vacuum source mass has been reported by
baryonic antimatter. Following Hughes and Holzscheiter Jaffe et al. (2016).
(1991), they expressed a possible gravitational anomaly Tests of gravity are interconnected with the searches
acting on antimatter with a parameter ag , which modi- for exotic forces. Leefer et al. (2016) have suggested and
fies the effective newtonian gravitational potential U to implemented the use of atomic spectroscopy to search for
give ag U , setting an upper limit of |ag − 1| < 8.7 × 10−7 . Yukawa-type fifth-forces. By studying the behaviour of
The role of the internal kinetic energy of bound systems atomic transition frequencies at varying distances away
of matter in tests of the Einstein equivalence principle from massive bodies (e.g., the Sun, Moon, heavy masses
76

in the laboratory), Leefer et al. (2016) have placed con- derived values of coefficients for Lorentz and CP T viola-
straints on possible non-gravitational interactions of a tion in the Standard Model Extension discussed below.
scalar field with the photon, electron and nucleons. This The listed experiments include searches for Lorentz vio-
work also placed constraints on combinations of interac- lation (LV) in the matter, photon, neutrino, and gravity
tion parameters that cannot otherwise be probed with sectors. The Data Tables for Lorentz and CP T Viola-
traditional anomalous-force measurements. Leefer et al. tion has grown in length by 50% in the past three years
(2016) suggested further measurements to improve on demonstrating large number of new experiments in many
the current level of sensitivity. Such measurements in- sectors.
clude the use of more precise atomic clocks and other sys- This recent interest in tests of Lorentz symmetries is
tems (molecular, highly-charged ionic and nuclear transi- motivated by theoretical developments in quantum grav-
tions), and implementing different experimental geome- ity suggesting that Lorentz symmetry may be violated
tries (e.g., the size of the effect can be increased by up to at some energies, tremendous progress in experimental
four orders of magnitude by measuring atomic transition precision, and development of a theoretical framework to
frequencies first on Earth, then on a space probe headed analyze different classes of experiments. A particular at-
towards the Sun). traction of the LLI tests is a tantalizing possibility of a
positive result: a confirmed measurement of Lorentz vi-
olation would be an unambiguous signal of new physics.
XI. LORENTZ SYMMETRY TESTS The natural energy scale for strong LV induced by quan-
tum gravity is the Planck scale (MPl ∼ 1019 GeV/c2 ),
Local Lorentz invariance (LLI) is one of the founda- which is far beyond the reach of existing observations:
tions of the current laws of physics: the outcome of any even ultra-high energy cosmic rays still fall eight orders-
local non-gravitational experiment is independent of the of-magnitude short of the Planck scale. The good news
velocity and the orientation of the (freely-falling) appa- is that strong LV at the Planck scale may also lead to
ratus. The first test of Lorentz invariance was Michel- tiny but potentially observable low-energy LV. Therefore,
son’s 1881 experiment (Michelson, 1881) aimed at de- high-precision tests of LLI with matter, gravity, or light
tecting the ether (erroneously assumed to be the medium may provide insight into possible new physics and set
for electromagnetic wave propagation). This experiment limits on various theories such as quantum gravity. The
was further improved by Michelson and Morley (1887). bad news is that there are no predictions of the magni-
Michelson and Morley’s apparatus measured the interfer- tude of LV violation at low energies. Lorentz-violating
ence between two beams of light travelling back and forth effects may be suppressed by some power of the ratio R
along two perpendicular paths. This light interferometer between the electroweak scale and the natural (Planck)
was rotated relative to the Earth to test the isotropy of energy scale for strings: R = mew /MP l = 2 × 10−17
the speed of light. (Kostelecký and Potting, 1995) or electron mass to Plank
In the 1960s, the first spectroscopic tests of Lorentz scale 4 × 10−23 (Liberati and Maccione, 2009).
symmetry were performed by Hughes et al. (1960) and Lorentz violation tests are analyzed in the context of
Drever (1961) where they searched for sidereal variation an effective field theory known as the Standard Model ex-
of nuclear magnetic resonance (NMR) lines in 7 Li. The tension (SME). Two approaches are used when construct-
Hughes-Drever tests were inspired by the suggestion of ing such an effective field theory to describe Lorentz vi-
Cocconi and Salpeter (1958) that it might be possible, olations: (1) add renormalizable Lorentz-violating terms
based on Mach’s principle, for inertial mass to acquire a to the Standard Model Lagrangian (Colladay and Kost-
tensor character due to anisotropic distribution of mat- elecký, 1998) and (2) explicitly break Lorentz invariance
ter in the universe. This would cause a particle’s iner- by introducing nonrenormalizable operators (Myers and
tial mass to depend on the orientation of its orbit with Pospelov, 2003).
respect to the matter anisotropy, which in turn would In minimal SME, corresponding to the first approach,
generate energy shifts in atoms and nuclei. Experiments the Standard Model Lagrangian is augmented with every
similar to the Hughes-Drever test have come to be known possible combination of the SM fields that are not term-
as a “clock-comparison tests” in which the frequencies of by-term Lorentz invariant, while maintaining gauge in-
different atomic “clock” transitions are compared as the variance, energy–momentum conservation, and Lorentz
clocks rotate with the Earth. Since these early tests, invariance of the total action (Colladay and Kostelecký,
the field of Lorentz symmetry tests has flourished, en- 1998). Separate violations of LLI are possible for each
compassing almost all fields of physics (Kostelecký and type of particle, making it essential to verify LLI in
Russell, 2011; Liberati and Maccione, 2009; Mattingly, different systems at a high level of precision. Liberati
2005). The Data Tables for Lorentz and CP T Viola- and Maccione (2009) reviewed non-minimal SME exper-
tion, an extraordinary effort by Kostelecký and Russell imental tests, and all current limits are given in 2017
(2017), provides yearly updates of experimental progress edition of the Data Tables for Lorentz and CP T Vi-
of the last decade and gives tables of the measured and olation Kostelecký and Russell (2017). We limit this
77

review to recent AMO tests and proposals. The di- A. Electron sector of the SME
verse set of AMO Lorentz symmetry tests involves ex-
periments with atomic clocks (Wolf et al., 2006), other Testing LLI of the electron motion in an atom has an
precision spectroscopy measurements (Hohensee et al., advantage of testing for new physics in a well understood
2013a), magnetometers (Allmendinger et al., 2014a; Smi- system. In atomic experiments aimed at the LLI tests in
ciklas et al., 2011), electromagnetic cavities (Eisele et al., the electron-photon sector (Hohensee et al., 2013a; Prut-
2009), and quantum-information-trapped-ion technolo- tivarasin et al., 2015), one searches for variations of the
gies (Pruttivarasin et al., 2015). atomic energy levels when the orientation of the elec-
In minimal SME, a general expression for the quadratic tronic wave function is rotated with respect to a stan-
Hermitian Lagrangian density describing a single spin- dard reference frame. Generally, one uses the Sun cen-
1/2 Dirac fermion of mass m (electron, proton, or neu- tered celestial-equatorial frame (SCCEF) for the analysis
tron) in the presence of Lorentz violation is given by of the experiments (Kostelecký and Mewes, 2002), indi-
(Kostelecký and Lane, 1999) cated by the coordinate indexes T , X, Y , and Z. For
example, the cµν tensor has 9 components that need to
1 ←
→ be experimentally determined: parity-even cT T and cJK
L= icψΓν ∂ ν ψ − M c2 ψψ, (93)
2 and parity-odd cT J , where J, K = X, Y, Z. The elements
where ψ is a four-component Dirac spinor, cJK which describe the dependence of the kinetic energy
on the direction of the momentum have a leading or-

→ der time-modulation period related to the sidereal day
f ∂ ν g = f ∂ ν g − g∂ ν f
(12-hr and 24-hr modulation) in the laboratory experi-
, ments described below. The cT J and cT T describe the
dependence of the kinetic energy on the boost of the lab-
1
M = m + aµ γ µ + bµ γ5 γ µ + Hµν σ µν (94) oratory frame and have a leading order time-modulation
2 period related to the sidereal year. The terms cT J are
and proportional to the ratio of the Earth’s orbital velocity
to the speed of light β⊕ ≈ 10−4 ; and the cT T term is
1 suppressed by β⊕ 2
≈ 10−8 , resulting in weaker bound on
Γν = γν +cµν γν +dµν γ5 γν +eν +iγ5 fν + gλµν σλµ . (95)
2 these components of the cµν tensor. The indexes (0,1,
The first terms in the expressions for M and Γν give the 2, 3) are used for the laboratory frame. The most sen-
usual SM Lagrangian. Lorentz violation is quantified by sitive LLI tests for electrons have been conducted with
the parameters aµ , bµ , cµν , dµν , eµ , fµ , gλµν , and Hµν . neutral Dy atoms (Hohensee et al., 2013a) and Ca+ ions
The coefficients in Eq. (94) have dimensions of mass; the (Pruttivarasin et al., 2015) as described below.
coefficients in Eq. (95) are dimensionless. The field oper- Violations of Lorentz invariance in bound electronic
ators in Eqs. (94,95) containing the coefficients cµν , dµν , states result in a perturbation of the Hamiltonian that
and Hµν are even under CP T and the remaining ones can be described by (Hohensee et al., 2013a; Kostelecký
are odd under CP T . The framework of interpreting the and Lane, 1999)
laboratory experiments involving monitoring atomic or   2
nuclear frequencies in terms of the SME coefficients is (0) 2U p 1 (2) (2)
δH = − C0 − 2 c00 − C T , (96)
described in detail by Kostelecký and Lane (1999); Kost- 3c 2me 6me 0 0
elecký and Mewes (2002). Such atomic experiments may
be interpreted as Lorentz-invariance tests for the photon, where p is the momentum of a bound electron. The
electron, and nuclear constituents, such as proton and second term in the parentheses gives the leading order
neutron, with varying sensitivities to different combina- gravitational redshift anomaly in terms of the Newto-
tions of LLI effects. A number of experiments are sen- nian potential U . We refer the reader to Kostelecký and
sitive to either electron or nucleon sectors, with photon Tasson (2011) for the study of the gravitational couplings
contributions appearing in all atomic experiments. AMO of matter in the presence of Lorentz and CP T violation
tests of LLI also include testing isotropy of gravity, a test and the derivation of the relativistic quantum Hamilto-
of the LLI of post-Newtonian gravity was performed by nian from the gravitationally coupled minimal SME. The
(0) (2)
Müller et al. (2008) by monitoring Earths gravity with parameters C0 and C0 are elements of the cµν tensor
a Mach-Zehnder atom interferometer. Expressed within in the laboratory frame introduced by Eq. (95):
the standard model extension, the analysis limits four co- (0)
efficients describing anisotropic gravity at the ppb level C0 = c00 + (2/3)cjj , (97)
(2)
and three others at the 10 ppm level. Using the SME, C0 = cjj + (2/3)c33 , (98)
Müller et al. (2008) explicitly demonstrated how their ex-
(2) (2)
periment actually compares the isotropy of gravity and where j = 1, 2, 3. The C±1 and C±2 do not contribute
electromagnetism. to the energy shift of bound states. The values of the
78

(0) (2)
C0 and C0 in the laboratory frame are the functions nucleus since both Dy isotopes used in the experiment,
162
of the cµν tensor in SCCEF frame and the velocity and Dy and 164 Dy, have nuclear spin I = 0.
orientation of the lab. The Dy experiment used repeated measurements ac-
(2) (2)
The nonrelativistic form of the T0 operator is T0 = quired over nearly two years to obtain constraints on
2 2 eight of the nine elements of the cµν tensor. Hohensee
p − 3pz . Predicting the energy shift due to LV involves
calculating of the expectation value of the above Hamilto- et al. (2013a) tightened the previous limits (Altschul,
nian for the atomic states of interest. The larger the ma- 2006, 2010; Müller et al., 2007) on four of the six parity-
trix elements, the more sensitive is this atomic state. One even components by factors ranging from 2 to 10, limiting
has to take into account that only a transition-energy Lorentz violation for electrons at the level of 10−17 for
shift can be measured, so the difference of the sensitiv- the cJK components. Previous studies used rotating op-
ities of the upper and lower states is important for the tical Fabry-Perot resonators and microwave whispering-
final experimental analysis in terms of the cµν tensor. gallery sapphire resonators (Müller et al., 2007) and high-
The most accurate tests in the electron-photon sector energy astrophysical sources, synchrotron and inverse
can be conducted in atoms or ions with highest possible Compton data (Altschul, 2006, 2010; Müller et al., 2007)
sensitivities which are amenable to high-precision mea- to constrain cµν coefficients for electrons.
surement techniques. Since the operators in Eq. (96) Hohensee et al. (2013a) also improved bounds on gravi-
contain the second power of the momentum operator p, tational redshift anomalies for electrons (Hohensee et al.,
the corresponding matrix elements are expected to be 2011; Vessot et al., 1980) by 2 orders of magnitude, to
large for orbitals with large kinetic energy. This happens 10−8 .
for atomic 4f -electrons localized deep inside the atom in
the area of large (negative) potential and kinetic energy
in some atomic systems. 2. LLI test with calcium ion
We note that the formalism is the same for the LV
Pruttivarasin et al. (2015) performed a test of Lorentz
violation in the nuclei, and the expectation values of the
symmetry using an electronic analogue of a Michelson-
same operators (but for the nuclear states) determine the
Morley experiment using the 2 D5/2 atomic states of
sensitivity. 40
Ca+ ion with anisotropic electron momentum distri-
butions. The experiment involved interfering such states
aligned along different directions. A pair of 40 Ca+ ions
1. LLI tests with dysprosium
was trapped in a linear Paul trap, with a static magnetic
field applied defining the eigenstates of the system. The
A joint test of local Lorentz invariance and the Ein-
direction of this magnetic field changes with respect to
stein equivalence principle for electrons was reported by
the Sun as the Earth rotates, resulting in a rotation of
Hohensee et al. (2013a) using long-term measurements of
the interferometer as illustrated in Fig. 23.
the transition frequency between two nearly degenerate
In the magnetic field, the 3d 2 D5/2 atomic state splits
states of atomic dysprosium.
into six states with the magnetic quantum numbers mJ =
Dy a lanthanide element with partially filled elec- ±1/2, ±3/2, and ±5/2. Using the Hamiltonian given by
tronic f -shell possesses two near-degenerate, low-lying Eq. (96), and calculating the corresponding matrix el-
excited states with significant momentum quadrupole (2)
ements of the T0 operator, the energy shift of these
moments, opposite parity, and leading configurations: 2
D5/2 atomic states induced by the Lorenz violation in
[Xe]4f 10 5d6s, J = 10 (state A) and [Xe]4f 9 5d2 6s, J = 10
the electron-photon sector is given by
(state B). The energy difference between states A and B
can be measured directly by driving an electric-dipole δE  (2)
= 2.16 × 1015 − 7.42 × 1014 m2J C0 . (100)
  
transition with a radio-frequency (rf) field. The average h
shift in the the B → A transition frequency ωrf , properly Since the LV energy shift depends on the magnetic quan-
weighted for transition frequencies for different magnetic tum number, monitoring the energy difference between
sublevels, is given by the mJ = ±1/2 and the mJ = ±5/2 Zeeman substates
    during the Earth rotation probes the cJK components of
δωrf (0) 2U (2)
= 1014 Hz 500 C0 −

c00 + 9.1C 0 , the LV tensor. The frequency difference (in Hz) between
2π 3c2 the LV shifts of the mJ = 5/2 and mJ = 1/2 substates
(99)
of the 3d 2 D5/2 manifold is given by
where U is the Sun’s gravitational potential. The sign
of the frequency shift is opposite for 162 Dy and 164 Dy. 1 (2)
EmJ =5/2 − EmJ =1/2 = −4.45(9) × 1015 Hz × C0 .
  
The uncertainty in the numerical coefficient in front of h
the first term in the square brackets may be large due to (101)
(0)
the compilations in the evaluation of the matrix elements This experiment is not sensitive to the scalar C0 coeffi-
of the p2 operator. There is no LV contribution from the cient of Eq. (96).
79

for another system with a long-lived or ground state that


(2)
has a large hj|T0 |ji matrix element. Dzuba et al. (2016)
carried out a systematic study of this quantity for various
systems and identified general rules for the enhancement
of the reduced matrix elements of the T (2) operator. The
authors identified the ytterbium ion Yb+ to be an ideal
system for future LV tests with high sensitivity, as well
as excellent experimental controllability. The sensitivity
of the 4f 13 6s2 2 F7/2 state of Yb+ to LV is over an order
of magnitude higher than that of the Ca+ 2 D5/2 state.
This state also has an exceptionally long lifetime on the
order of several years (Huntemann et al., 2012), so the
proposed experiment is not limited by spontaneous decay
during a measurement in contrast to the Ca+ case.
FIG. 23 (color online). Rotation of the quantization axis of
the experiment with respect to the Sun as the Earth rotates. Experimental techniques for precision control and ma-
A magnetic field (B) is applied vertically in the laboratory nipulation of Yb+ atomic states are particulary well de-
frame to define the eigenstates of the system. As the Earth ro- veloped owing to atomic clock (Huntemann et al., 2016)
tates with an angular frequency given by ω⊕ = 2π/(23.93 h), and quantum information (Islam et al., 2013) applica-
the orientation of the magnetic field and, consequently, that tions making it an excellent candidate for searches of the
of the electron wave packet (as shown in the inset in terms Lorentz-violation signature.
of probability envelopes) changes with respect to the Sun’s
rest frame (positions at various times UTC are illustrated). Dzuba et al. (2016) estimated that experiments with
The angle χ is the colatitude of the experiment. From Prut- the metastable 4f 13 6s2 2 F7/2 state of Yb+ can reach
tivarasin et al. (2015). sensitivities of 1.5 × 10−23 for the cJK coefficients, over
105 times more stringent than current best limits. More-
over, the projected sensitivity to the cT J coefficients will
The main source of decoherence in this experiment be at the level of 1.5 × 10−19 , below the ratio between
is magnetic field noise, since it also shifts the energies the electroweak and Planck energy scales. Similar sen-
of the Zeeman substates. This problem is resolved by sitivities may potentially be reached for LV tests with
applying quantum-information inspired techniques and highly charged ions (Dzuba et al., 2016), given future de-
creating a two-ion product state that is insensitive to velopment of experimental techniques for these systems
magnetic field fluctuation to first order. The energy dif- (Schmöger et al., 2015). Another interesting future pos-
ference between the two-ion states | ± 5/2, ∓5/2i and sibility is measuring transition energies of rare-earth ions
|±1/2, ∓1/2i was measured for 23 hours, resulting in the doped in crystalline lattices, which can be highly sen-
limit of h × 11 mHz. Pruttivarasin et al. (2015) pointed sitive to the electron SME parameters (Harabati et al.,
out that the experimental results may be interpreted in 2015).
terms of either photon or electron LV violation described Also of note is the unique work of Botermann et al.
via c0µν = cµν + kµν /2, where the first term refers to the (2014), where a clock comparison test was performed at
electron LV and the second term to the photon LV. The relativistic speeds using Li+ in a storage ring.
Ca+ experiment improved the limits to the c0JK coeffi-
cients of the LV-violation in the electron-photon sector
to the 10−18 level. Because 40 Ca+ nucleus has nuclear
spin I = 0, there in no nuclear LV contribution, just as B. Proton and neutron sectors of the SME
in the case of the Dy experiment. The same experiment
can be interpreted as testing anisotropy in the speed of 1. Cs clock experiment
light with the sensitivity similar to that of more recent
work reported by Nagel et al. (2015). Another example of a clock comparison test is the work
of Wolf et al. (2006) who used a cold Cs atomic clock
to test LLI in the matter sector, setting limits on the
3. Future prospects and other experiments tensor Lorentz-violating coefficients for the proton. The
Cs clock, which is also the primary frequency standard
With optimization, both Dy and Ca+ experiments defining the second, operates on the |F = 3i ←→ |F =
could yield significantly improved constraints. An op- 4i hyperfine transition of the 133 Cs 6S1/2 ground state,
timized Dy experiment may reach sensitivities on the or- where F = J + I is the total angular momentum and Cs
der of 9 × 10−20 in one year for the cJK components nuclear spin is I = 7/2. In the magnetic field, F = 3 and
(Hohensee et al., 2013a). F = 4 clock states split into 7 and 9 Zeeman substates
Further significant improvement of LV constraints calls with mF = [−3, 3] and mF = [−4, 4], respectively. The
80

atomic clock operates on the


|F = 3, mF = 0i ↔ |F = 4, mF = 0i (102)
hyperfine transition at 9.2 GHz, which is insensitive to
either Lorentz violation or first-order magnetic field ef-
fects, but the other transitions with δmF 6= 0 are used for
magnetic field characterization. To test Lorentz symme-
try, Wolf et al. (2006) monitored a combination of clock
|F = 3, mF = 3i ←→ |F = 4, mF = 3i, (103)
and
FIG. 24 Change of K-3 He self-compensating SERF comagne-
|F = 3, mF = −3i ←→ |F = 4, mF = −3i (104) tometer signal for 180◦ platform rotation as a function of the
initial platform angle. Figure from Brown et al. (2010).
transitions to form a combined observable

νc = ν+3 + ν−3 − 2ν0 . (105) a linearly polarized laser beam propagating along x, is
given to leading order by:
The ν0 , ν+3 , and ν−3 are frequencies of (102), (103),  
and (104) transitions above. The combined observable is e e γe N e Ωy
Px = P z β − βy + , (106)
used to avoid the dominant noise source - the first order Γrel y γN
Zeeman shift due to the magnetic field fluctuations which
strongly affect the states with mF 6= 0, but cancels for the where βyN and βye describe the phenomenological SME
±mF combination. Since the mF is the same for upper background fields along the y-direction coupling to the
3
and lower states of all transitions, there is no Lorentz- He nucleus (N ) and valence electron (e) of K, respec-
violating tensor component from the electron sector. The tively, Pze is the K electron spin polarization along z, Γrel
133
Cs nucleus has one unpaired proton, the experiment is the relaxation rate for K polarization, γe and γN are
is interpreted in terms of the proton LV parameters of the gyromagnetic ratios for electrons and 3 He nuclei, re-
the cµν tensor, using the Schmidt nuclear model. The spectively, and Ωy is the rotation rate of the apparatus.
Cs clock experiment set the limits for parameters for the A number of steps are taken to eliminate various sources
proton at the 10−21 − 10−25 level. A reanalysis of this of noise and systematic error. For example, the signal de-
experiment is currently in progress (Pihan-Le Bars et al., scribed by Eq. (106) depends explicitly on the rotation
2017) in order to obtain a constraint on the proton cT T rate of the apparatus, so a nonzero Pxe is generated by the
coefficient. gyro-compass effect due to the Earth’s rotation (Heckel
et al., 2008; Venema et al., 1992). To compensate for this
effect, the experiment is mounted on a rotary platform.
2. Comagnetometer experiments Figure 24 shows the change in the comagnetometer signal
for a 180◦ rotation as a function of the initial platform an-
Some of the most stringent clock-comparison tests of gle, demonstrating a significant effect of Earth’s rotation
LLI (Brown et al., 2010; Smiciklas et al., 2011) have been on Pxe . In order to test LLI, the orientation of the ap-
carried out using the self-compensating spin-exchange paratus is alternated between North-South or East-West
relaxation-free (SERF) comagnetometry scheme (Kor- (marked by crosses on the plot of Fig. 24) every 22 s
nack et al., 2005; Kornack and Romalis, 2002), which over the course of many days. Nonzero values of βyN or
was also used for constraining anomalous dipole-dipole βye would lead to sidereal oscillation of the amplitude of
interactions (Vasilakis et al., 2009) as discussed in detail the difference between the North-South or East-West sig-
in Sec. VII. nals. The results of the measurements, carried out over
The experiment of Brown et al. (2010) employs over- 143 days, are consistent with no LLI violation. Combined
lapping ensembles of K and 3 He coupled via spin- with the constraints on electron couplings to SME back-
exchange collisions. The atoms are in the low-magnetic- ground fields from Heckel et al. (2008), the results of this
field SERF regime where broadening of the Zeeman reso- experiment probe neutron couplings to SME background
nances due to spin-exchange collisions is eliminated. The fields at energy scales ∼ 10−25 eV (Brown et al., 2010).
magnetic field along the z-direction is tuned to the com- The closely related experiment of Smiciklas et al.
pensation point where the K-3 He SERF comagnetometer (2011) uses a 21 Ne-Rb-K SERF comagnetometer with
is insensitive to magnetic fields but highly sensitive to a shot-noise-limited sensitivity to LLI violations that is
anomalous interactions that do not scale with the mag- an order of magnitude better than the K-3 He comage-
netic moments. The spin polarization of the K atoms tometer. Additionally, because the nuclear spin of 21 Ne
along the x-direction, probed via optical rotation with is I = 3/2, the 21 Ne-Rb-K comagnetometer is sensitive
81

to tensor anisotropies as well as the vector anisotropies ented orthogonally to one another. Whispering gallery
probed by the K-3 He comagetometer (I = 1/2 for 3 He). mode resonances near 13 GHz were excited and the ap-
A new version of this experiment, performed at the South paratus was rotated with a period of ≈ 100 s. Again
Pole to better control for the gyro-compass effect (Hedges the observable was the spatial-orientation-dependence of
et al., 2015), is expected to improve on these constraints the beat frequency between the signals from the two mi-
by yet another order-of-magnitude. crowave cavities. Compared to the original experiments
A different scheme was used by Allmendinger et al. of Michelson and Morley (1887), this is an improvement
(2014a) to test LV in the neutron sector at a similar level of 17 orders of magnitude. Kosteleckỳ et al. (2016) also
of accuracy by measuring precession of overlapping en- point out that gravitational wave detectors, km-scale
sembles of 3 He and 129 Xe atoms [although note the dis- laser interferometers with exquisite sensitivity, establish
cussion between Romalis et al. (2014) and Allmendinger constraints on certain LV parameters of the SME that are
et al. (2014b) regarding these results]. several orders-of-magnitude more stringent than previous
limits. Even more stringent constraints come from re-
interpretation of existing data: Flambaum and Romalis
C. Quartz oscillators (2017) have noted that by analyzing the Coulomb inter-
actions between the constituent particles of atoms and
Lo et al. (2016) proposed and demonstrated a novel nuclei, comagnetometer experiments testing LLI (Smi-
approach to LLI tests in the matter sector taking ad- ciklas et al., 2011) establish that the speed of light is
vantage of new, compact, and reliable quartz oscillator isotropic to a part in 1028 .
technology. Violations of LLI in the matter and pho-
ton sector of the SME generate anisotropies in particles’
inertial masses and the elastic constants of solids, giv- XII. SEARCH FOR VIOLATIONS OF QUANTUM
ing rise to anisotropies in the resonance frequencies of STATISTICS, SPIN-STATISTICS THEOREM
acoustic modes in solids. Thus the spatial-orientation-
dependence of acoustic resonances can be used to con- The concept of identical particles is unique to quan-
strain LV: the initial experiment of Lo et al. (2016) set tum physics. In contrast to, for example, identical twins
constraints on certain SME parameters some 3 orders- or so-called “standard-candle” supernovae, all electrons,
of-magnitude more stringent than other laboratory tests helium atoms, 85 Rb nuclei, etc., are, as far as we can
and ten times more stringent than astrophysical limits. tell, truly identical to each other. This means that if
we have a wavefunction representing a system containing
identical particles, particle densities should not change
D. Photon sector of the SME upon interchange of two identical particles. As a conse-
quence, the wavefunction should either remain invariant
At the close of this section, we return to the experimen- or change sign under permutation of identical particles.
tal setup that was the basis of the first tests of Lorentz This is the essence of the permutation-symmetry postu-
invariance, the rotating interferometer. Recent exper- late (PSP). The spin-statistics theorem (SST) dictates
iments with rotating optical and microwave resonators which of the two options is realized given the particular
establish some of the most stringent constraints on LV intrinsic spin of the particles. (This connection is non-
in the photon sector (Chen et al., 2016a; Eisele et al., trivial and, one might argue, a-priori unexpected.) The
2009; Herrmann et al., 2009; Hohensee et al., 2010; Müller resulting division of particles into fermions and bosons is
et al., 2007; Nagel et al., 2015). For example, Eisele et al. one of the cornerstones of modern physics.
(2009) searched for a spatial anisotropy of the speed of The SST is proved in the framework of relativistic field
light using two orthogonal standing-wave optical cavities theory using the assumptions of causality and Lorentz
contained in a single block of glass with ultralow ther- invariance in 3 + 1 spacetime dimensions, along with a
mal expansion coefficient. The orthogonal cavities were number of more subtle implicit assumptions enumerated
probed with a laser and rotated nearly 200,000 times over by Wichmann (2001).
the course of 13 months using an air cushion rotation ta- While it is notoriously difficult to build a consistent rel-
ble with low axis wobble, low vibration level, and active ativistic theory incorporating SST and PSP violations (a
stabilization of optical elements. The quantity of interest feat that has not as yet been accomplished, to the best of
in the experiment was the spatial-orientation-dependence our knowledge), it is important to put these properties to
of the beat frequency between the light from the two cav- rigorous tests given their fundamental importance in our
ities, which was found to be invariant at a level below a understanding of Nature. One may think of such tests as
part in 1017 . The experiment of Nagel et al. (2015) im- probing all the assumptions in the SST proof, as well as
proved upon this result by a factor of 10 through the use providing a possible experimental window into theories
of two cryogenic cylindrical copper cavities loaded with that go beyond conventional field theory, for instance,
identical sapphire dielectric crystals whose axes were ori- string theory. For example, plausible theoretical scenar-
82

ios for small spin-statistics violations include excitations et al., 1999), used a selection rule for atomic transitions
of higher dimensions allowing particles to possess wrong- that is closely related to the Landau-Yang theorem
symmetry states in the usual 3-dimensional space while (Landau, 1948; Yang, 1950) in high-energy physics. The
maintaining the correct symmetry in an N -dimensional selection rule states that two collinear, equal-frequency,
space (Greenberg and Mohapatra, 1989). photons cannot be in a state of total angular momentum
Since all our observations so far are consistent with one. An example in high-energy physics is that the
PSP and SST, the experiments should search for small neutral spin-one Z0 boson cannot decay to two photons.
violations of PSP and SST, the effects sometimes referred (According to the Particle Data Group, the branching
to as “violations of quantum statistics.” ratio for this process is limited to < 5.2 × 10−5 , although
A comprehensive review of the literature on the spin- there are additional reasons that suppress such decay.)
statistics connection and related issues such as the Pauli For atoms, the selection rule means that two collinear
exclusion principle and particle indistinguishably, includ- equal-frequency photons cannot stimulate a transition
ing theoretical background and experimental searches, is between atomic states of total angular momentum
given by Curceanu et al. (2012). Here we limit our discus- zero and one. The experiment employed an atomic
sion to examples of recent experiments to give the reader beam of barium optically excited in a power-buildup
a flavor of atomic, molecular, and optical techniques that cavity and resulted in a limit for two photons to be in
are used in this field. wrong (i.e., fermionic) symmetry state of < 4.0 × 10−11 .
The strongest limit on a possible violation of the Pauli Further improvements by several orders of magnitude
exclusion principle for electrons currently comes from the are expected in ongoing experiments using ultra-cold Sr
VIP experiment at Gran Sasso (Marton et al., 2013). atoms (Guzman et al., 2015).
Here strong electric current is flown through a copper
sample and Pauli-forbidden atomic transitions involving As mentioned above, quantum-statistics violation
occupied atomic orbitals are searched for by measuring would be an effect outside of the framework of conven-
x-rays at the anticipated transition energy. The limit on tional field theory, in contrast to most other “exotic”
the probability for two electrons to be in a symmetry- effects discussed in this review. Combined with the ab-
forbidden state is currently < 4.7 · 10−29 with expected sence of a consistent alternative framework, discussion of
improvement in the upgraded VIP2 experiment by fur- such effects often leads to conceptual difficulties, includ-
ther two orders of magnitude (Marton et al., 2017; Shi ing questions like: What is the experiment really testing?,
et al., 2016). How can we compare results from different experiments?,
Molecular spectroscopy has played an important his- etc.
torical role in establishing the experimental basis for the The results of some early experiments were dismissed
PSP and the SST (Curceanu et al., 2012). The gen- as they did not take into account a so-called superselec-
eral idea is that in a molecule containing two identi- tion rule stating that the permutation symmetry of a sys-
cal nuclei, rotational states corresponding to the over- tem of identical particles cannot change in the course of
all molecular wavefunction being symmetric (in the case the system’s evolution. Being truly identical implies that
of half-integer-spin nuclei) or antisymmetric (in the case the particles cannot be distinguished by any measure-
of integer-spin nuclei) are forbidden by quantum statis- ment. In particular, this means that all operators cor-
tics, and so the spectral lines involving these molecular responding to physical observables must commute with
states are absent from the molecular spectrum. A pow- all exchange operators ξ, for example [H, ξ] = 0 for any
erful experimental methodology for testing for statistics Hamiltonian H. This fact is used by Amado and Pri-
violations is to look for such forbidden lines (Tino, 2001). makoff (1980) to derive the aforementioned superselec-
Recent experiments (Cancio Pastor et al., 2015) us- tion rule, which implies:
ing saturated-absorption cavity ring-down spectroscopy
hA|H|Si = 0 , (107)
searched for forbidden rovibrational lines at a 4.25 µm
wavelength in the spectra of the 12 C16 O2 molecule con- where |Si and |Ai are exchange symmetric and antisym-
taining two bosonic oxygen nuclei. They limited the metric states, respectively. The superselection rule pre-
relative probability for the molecule to be in a wrong- vents, for example, the transition between a symmetric
symmetry state at < 3.8 × 10−12 level, significantly im- and antisymmetric state (as was searched for in some
proving on earlier results. An interesting extension is to early experiments purporting to test the SST) based
molecules containing more than two identical nuclei that purely on the fact that the particles are identical and
would allow to probe for more complex permutation sym- not on the PSP or SST. However, it is important here
metries than are allowed for just two identical particles. to note that the superselection rule does not prevent cre-
An experimental test of Bose-Einstein (BE) statistics ation of particles with mixed statistics. The quon al-
and, consequently, the SST as it applies to photons gebra (Greenberg, 1991; Greenberg and Hilborn, 1999),
interacting with atoms was carried out by English et al. for example, takes advantage of this exception by postu-
(2010). The experiment, extending earlier work (DeMille lating creation and annihilation operators which do not
83

obey the usual commutation relations, leading to the cre- • What lies beyond the Standard Model of particles
ation of particle states which are neither symmetric nor ad interactions?
antisymmetric. Another immediate consequence of the
superselection rule is that a description in terms of a • How can general relativity be unified with quantum
wavefunction with a mixed permutation symmetry is not theory? ...
acceptable and a density matrix should be used instead.
These questions are, in a sense, “urgent.” For instance,
A further discussion of these points and related references
dark matter constitutes most of the mass in galaxies in-
can be found in the paper by Elliott et al. (2012).
cluding our own, and so it is likely that a discovery of
the dark-matter composition is “around the corner.”
XIII. CONCLUSION We hope that with this review, we have succeeded in
conveying to the reader our own excitement and antic-
AMO physics has been crucially important in laying ipation of forthcoming paradigm-shifting discoveries in
the foundations of our understanding of the fundamen- fundamental physics with atoms, molecules, and light.
tal laws of nature ever since the advent of precision
spectroscopy in the 19th century. The most remarkable
success is the discovery of the inevitability of quantum ACKNOWLEDGEMENTS
theory and its subsequent spectacular development, in-
cluding firming up such fundamental concepts as indis- We are grateful to Catalina Curceanu, Victor Flam-
tinguishability of identical particles, the spin-statistics baum, Kent Irwin, Mikhail Kozlov, Konrad Lehnert, Hol-
connection, the role of discrete symmetries such as par- ger Müller, Sergey Porsev, Surjeet Rajendran, Michael
ity and time-reversal, entanglement, relativistic quantum Snow, Yevgeny Stadnik, and Alexander Sushkov for help-
mechanics and quantum field theory, and many others. ful discussions. AD acknowledges the support of the
From the early days, AMO physics has been closely con- National Science Foundation under grants PHY-1506424
nected to astronomy and astrophysics, from the discov- and PHY-1607396. DB acknowledges the support of the
ery of new elements in the solar spectrum to determining DFG Koselleck program, the Heising-Simons and Simons
the velocities of stars and measuring the expansion of the Foundations, and the National Science Foundation under
Universe via red shifts of spectral lines. The list of sem- grant PHY-1507160, as well as the European Research
inal fundamental physics discoveries using AMO tech- Council (ERC) under the European Unions Horizon 2020
niques can, of course, be made almost arbitrarily long. research and innovation program (grant agreement No.
Remarkably, two centuries after its birth, the field of 695405). DD acknowledges the support of the National
precision AMO tests of fundamental physics continues Science Foundation under grant PHY-1404146, the Tem-
to be at the forefront of discovery, showing no signs of pleton Foundation, and the Heising-Simons Foundation.
slowing down! Conversely, with collider physics becom- DFJK acknowledges the support of the National Science
ing more and more expensive and potentially reaching Foundation under grant PHY-1307507 and the Heising-
saturation in terms of accessible particle energies and Simons and Simons Foundations. MSS acknowledges the
intensities, AMO physics beautifully complements high- support of the National Science Foundation under grants
energy physics and, in some cases, provides powerful ways PHY-1404156 and PHY-1620687. This research was per-
to indirectly explore potential new phenomena at energy formed in part under the sponsorship of the U.S. Depart-
scales reaching orders of magnitude beyond what can be ment of Commerce, National Institute of Standards and
expected to be directly accessible with accelerators in any Technology and the National Science Foundation via the
foreseeable future. Physics Frontiers Center at the Joint Quantum Institute.
Having powerful AMO tools for fundamental-physics
inquiry is especially important because there are many
basic properties of the Universe that we do not under- Appendix A: Notations, units, and abbreviations
stand:
1. Atomic and molecular properties as encoded in
• What are dark matter and dark energy? spectroscopic notation

• Why is there so much more matter in the Universe Atoms and molecules make wonderful clocks and pre-
than antimatter? cision measurement instruments because their electronic
• Why are the masses of all known particles so much states offer read, write and storage capabilities extending
smaller than the fundamental energy scales such as across many decades of bandwidth. Key properties of a
the grand-unification and the Planck scales? given state can be understood by symmetry considera-
tions, for which an understanding of conventional spec-
• Why do strong interactions appear to respect the troscopic notation is a useful aid. We present a brief sum-
CP symmetry? mary of notation that is germane to most of the specific
84

examples discussed in this paper. The concepts can be There are 140 independent electronic states that are
found on display in the periodic table of the elements con- members of the (A2) configuration. These are differenti-
structed for use in atomic spectroscopy (Dragoset et al., ated by the term and level hierarchies. The ground state
2017). More comprehensive accounts can be found in of Ce has the term and level designation
Martin and Wiese (2002) and Bunker, P. R. et al. (1997);
Schutte, C. J. H. et al. (1997a,b). [Xe] 4f 5d 6s2 1 G◦4 . (A3)

Four properties are encoded in the rightmost expression,


1 ◦
2. Atomic symmetries G4 , of expression (A3):

The conventional periodic table of the elements • The state’s total electronic spin angular momentum
(Dragoset et al., 2017) is laid out in a way that displays S, (in units of ~), which is encoded as 2S + 1 in the
1
the Aufbau principle. As the atomic number Z increases, G superscript. Here the state is a “spin singlet”
electrons are added one by one to atomic electron shells, with S = 0.
n, l. These are labeled by the integer principal quantum • The state’s total electronic orbital angular momen-
number n ≥ 1 and orbital angular momentum quantum tum L, (in units of ~), which is encoded as a capital
number l , (0 ≤ l < n). These two quantum numbers are letter. The string SPDFGHIK expresses the char-
encountered in the nonrelativistic quantum theory of the acter values for 0 ≤ L ≤ 7 . Here the state has
hydrogen atom. The beginning of the nth row of the pe- L = 4.
riodic table marks the start of filling the electron shell
shell n, 0, and the end marks the complete filling of the • The state’s total electronic angular momentum J,
electron shell n, n − 1. (in units of ~), which is shown in the 1 G◦4 subscript.
This representation of atomic structure is only an ap- Here J = 4, consistent with L = 4 and S = 0.
proximate model, but it also defines a zeroth-order basis
of many-electron wavefunctions that can be consistently • The state’s parity under inversion of spatial coor-
improved upon and enlarged by techniques of quantum dinates. This is sometimes shown by ◦ if the parity
many-body theory. Good guidance for rough estimates is odd, the superscript is omitted for even-parity
of energies and transition probabilities is communicated states. Indeed it is redundant, because
P the parity
in a standard notation. This notation expresses, in or- is given by (−1) to the power of k lk , where the
der of descending magnitude of energy: the atomic mean sum runs over all atomic electrons. Here it is odd,
field (configuration), electron-electron interaction (term), which is readily verified since 4f is the only elec-
spin-orbit interaction (electronic level), possible electron- tron shell that makes an odd contribution to the
nucleus interactions (hyperfine level), and projection of sum.
the total angular momentum (Zeeman sublevel).
Of these four indices, only two are exact: parity (to
We illustrate this using the example of the ground state
the extent that electroweak interactions are negligible)
of cerium (Ce, Z = 58). Its electron configuration is con-
and total angular momentum J (in cases where there is
ventionally described as
no nuclear angular momentum or neglecting hyperfine
1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p6 4d10 5s2 5p6 4f 1 5d1 6s2 . (A1) structure). Concerning J, when there is no hyperfine
structure, the application of a weak magnetic field re-
Here s, p, d, f designate l = 0, 1, 2, 3, and the superscripts veals that the ground level has 2J + 1 distinct Zeeman
designate occupation numbers. Thus, starting from the sublevels. As for S and L, there are cases in which it
left, the expression (A1) indicates that there are two elec- is meaningful to consider them as good quantum num-
trons in the n = 1, l = 0 shell, two more in 2, 0 subshell bers. For example, helium was once considered to con-
of the n = 2 shell, six more in 2, 1 subshell and so on up sist of two elements, ortho- and para-helium, because it
to the last closed subshell, 5p. That consolidated list of evinced distinctive singlet and triplet spectra, between
subshells is the same as that for the ground state of Xe, which there seemed to be no connection (Keesom, 1942).
so it is convenient to rewrite the expression (A1) as Now we understand those to be spectra associated with
[Xe] 4f 5d 6s2 , (A2) states that are (predominantly) S = 0 and S = 1, re-
spectively. In many cases of atoms with several valence
where no superscript indicates single occupancy of the electrons, different configurations and terms are strongly
electron shell. mixed, and the dominant configuration and LS term are
This shows that Ce has four electrons outside an listed. Indeed, our example expression, (A3), is a case in
isotropic closed-shell Xe-like core. These electrons de- point! The Ce ground state approximately described by
termine the symmetries of the electronic wavefunction (A3) is, in fact, a mixture of different configuration and
and have predominant influence on the atom’s chemical terms, where the weight of (A3) is about 60% (Kramida
and physical properties. et al., 2016).
85

When the atomic nucleus has no angular momentum,


TABLE III Mathematical symbols used and their meanings.
as is the case for all even-even isotopes in their nuclear
ground state, then a level designation such as expres- Symbol Meaning
sion (A3) identifies 2J + 1 degenerate states, correspond- c speed of light
ing to the distinct values of MJ , the projection of J upon
0 electric constant
some arbitrary quantization axis. When the nucleus has
G Newtonian constant of gravitation
spin I 6= 0, then the total atomic angular momentum is
designated F = I+J. As above, the corresponding quan- h Planck constant, ~ = h/2π
tum numbers are I, J, and F . The magnetic quantum e elementary charge
number of an atomic state, MF , takes one of 2F + 1 α fine structure constant
discrete values. The separate values of F correspond R∞ Rydberg constant
to different relative arrangements of electronic and nu- a0 Bohr radius
clear magnetic and electric moments, whereby they have µN nuclear magneton
slightly different energies. These energy differences were GF Fermi constant
called “hyperfine structure” when they were first inter-
g local acceleration due to the Earth’s gravity
preted by Pauli (1924), because they were a minute detail
me electron mass
of atomic spectra.
mp proton mass
MZ Z-boson mass
3. Molecular symmetries θW weak mixing angle
σi Pauli matrices, i = 1, 2, 3
The molecular term symbols that designate the elec- γµ Dirac matrices, µ = 0, 1, 2, 3
tronic states of a diatomic molecule take the form γ5 Dirac matrix associated with pseudoscalars
σ µν σ µν = 2i (γ µ γ ν − γ ν γ µ )
2S+1 (+/−) s single electron spin
ΛΩ,(g/u) . (A4) S multi-electron atom total spin
p linear momentum
The symbols Λ, S, Ω are analogous to their atomic coun-
E electric field (vector)
terparts L, S, J. Indeed, S designates the same net elec-
tronic spin in both cases. In the body frame of the B magnetic field (vector)
molecule, i.e. a frame in which the internuclear axis is C, P, T charge conjugation, parity, and time-reversal
fixed in space, rotations of all electrons about the in- transformations
ternuclear axis commute with the Hamiltonian, so pro- µ = mp /me proton to electron mass ratio, µ = 1/µ (Sec. II)
jections of electronic angular momentum upon that axis K dimensionless sensitivity factor of an energy level
can be taken to be good quantum numbers. The abso- to α-variation (Sec. II)
lute value of the projection upon this axis of electronic Kµ dimensionless sensitivity factor of an energy level
orbital angular momentum, L, is designated Λ (in units to µ-variation (Sec. II)
of ~). Thus, Λ = 0, 1, 2, 3, . . . , designated respectively mq average mass of light quarks (Sec. II)
by uppercase Greek letters, Σ, Π, ∆, Φ, . . . in analogy ΛQCD QCD energy scale (Sec. II)
with the atomic S, P, D, F, . . . . As atomic L is to Λ, so is κ dimensionless sensitivity factor of an energy level
atomic J to Ω, which is the magnitude of the projection to a variation of Xq = mq /ΛQCD (Sec. II)
of electronic total angular momentum upon the internu- kX dimensionless factor quantifying the spatial
clear axis, again in units of ~. As for J in atoms, Ω is variation of the fundamental constant X (Sec. II)
an integer or half integer. As an example, we consider a QW nuclear weak charge (Sec. IV)
state of the thorium oxide molecule, ThO, that is men-
d electric dipole moment (Sec. V)
tioned in Sec. V.F.2. Its electronic state is labeled there
P dimensionless electrical polarization (Sec. V)
as 3 ∆1 , thus S = 1, Λ = 2 and Ω = 1.
For isolated molecules, only total angular momen- S Schiff moment (Sec. V)
tum is rigorously conserved. Total angular momentum d˜ chromo-EDM (Sec. V)
and parity also depend on the rotational motion of the σ̂ unit vector along spin (Sec. VII)
molecule. The rotational quantum number is designated
J = 0, 1, 2 . . . ; the corresponding inversion symmetry is
(−1)J . For homonuclear diatomic molecules, there is
an additional quantum number associated with inver-
sion with respect to the symmetry plane bisecting the
line connecting the two nuclei. It can be even (German:
86

“gerade”) or odd (“ungerade”) under this transforma-


TABLE IV Abbreviations and their meanings.
tion, which is represented in the term symbol as g or
Abbreviation Meaning u.
AMO atomic, molecular and optical physics Finally, there is a symmetry of the molecular Hamilto-
ALPs axion-like particles
nian under reflection in any plane that contains the inter-
nuclear axis. The electronic wavefunction may be even
APV atomic parity violation
or odd under this transformation, which accounts for the
CPT combined operation CP T
+ or − superscript that is an option in the expression
CPV CP -violation (A4). It is used only for Σ states, the best-known ex-
cEDM chromo-EDM ample being the 3 Σ− g ground state of molecular oxygen,
DFSZ Dine-Fischler-Srednicki-Zhitnitskii O2 .
DM dark matter
EDM electric dipole moment
eEDM electron EDM 4. Units
EEP Einstein equivalence principle
GDM gravitational dipole moment The International System of Units (SI) is used through-
GR general relativity out this paper, unless noted otherwise. Atomic units are
GPS Global Positioning System often used in the source literature. In atomic units, the
values of elementary charge e, the electron mass me , and
GW gravitational wave
the reduced Planck constant ~ have numerical value 1,
HCI highly charged ion
and the electric constant 0 has numerical value 1/(4π).
ISL inverse-square law The conversion between SI and atomic units for com-
KSVZ Kim-Shifman-Vainshtein-Zakharov monly used quantities, including formulas and numerical
LHC Large Hadron Collider values, is given, for example, in Table XXXVII of Mohr
LLI local Lorentz invariance et al. (2016), p. 62. For example, atomic unit of electric
LPI local position invariance field is Eat = e/(4π0 a20 ).
LV Lorentz symmetry violation
MWDM Moody-Wilczek-Dobrescu-Mocioiu
MQM magnetic quadrupole moment 5. Symbols and abbreviations
NAM nuclear anapole moment
NIST National Institute of Standards and Technology The common mathematical symbols and abbreviations
which appear throughout the review are listed in Tables
NMR nuclear magnetic resonance
III and IV for convenience. Chapter-specific notations
n.r. nonrelativistic
are given under the chapter headings. The designations
PSP permutation-symmetry postulate specific to a single subtopic and used only briefly are not
QCD quantum chromodynamics tabulated below, but are defined the first time they are
SLI semileptonic interaction introduced. CODATA and Particle Data Group designa-
SM Standard Model tions are adopted in the review for common quantities.
SME Standard Model extention Every effort is made to use notations and abbreviations
SMt Schiff moment which most commonly appear in the literature.
SQUID Superconducting QUantum Interference Device
SCCEF Sun centered celestial-equatorial frame
SST spin-statistics theorem REFERENCES
SUSY supersymmetry
Aad, G, T. Abajyan, B. Abbott, J. Abdallah, S. Ab-
T,PV simultaneous T - and P -violation del Khalek, A. A. Abdelalim, O. Abdinov, R. Aben, B. Abi,
TV T -violating but P -conserving M. Abolins, O. S. AbouZeid, H. Abramowicz, H. Abreu,
VULF virialized ultralight field and et al. (ATLAS Collaboration) (2012), “Observation of
a new particle in the search for the Standard Model Higgs
WEP weak equivalence principle
boson with the ATLAS detector at the LHC,” Phys. Lett.
WIMP weakly-interacting massive particle B 716, 1.
UFF universality of free fall Abbott, B P, R. Abbott, T. D. Abbott, M. R. Abernathy,
F. Acernese, K. Ackley, C. Adams, T. Adams, P. Addesso,
R. X. Adhikari, et al. (LIGO Scientific Collaboration, and
Virgo Collaboration) (2016a), “GW151226: Observation of
gravitational waves from a 22-solar-mass binary black hole
coalescence,” Phys. Rev. Lett. 116, 241103.
87

Abbott, B P, R. Abbott, T. D. Abbott, M. R. Abernathy, Ahmadi, M, B. X. R. Alves, C. J. Baker, W. Bertsche, E. But-


F. Acernese, K. Ackley, C. Adams, T. Adams, P. Ad- ler, A. Capra, C. Carruth, C. L. Cesar, M. Charlton, S. Co-
desso, R. X. Adhikari, et al. (LIGO Scientific Collaboration, hen, et al. (2017), “Observation of the hyperfine spectrum
and Virgo Collaboration) (2016b), “Observation of gravita- of antihydrogen,” Nature 548, 66.
tional waves from a binary black hole merger,” Phys. Rev. Ahmadi, M, M. Baquero-Ruiz, W. Bertsche, E. Butler,
Lett. 116, 061102. A. Capra, C. Carruth, C. L. Cesar, M. Charlton, A. E.
Abbott, B P, R. Abbott, T. D. Abbott, M. R. Abernathy, Charman, S. Eriksson, et al. (2016), “An improved limit on
K. Ackley, C. Adams, P. Addesso, R. X. Adhikari, V. B. the charge of antihydrogen from stochastic acceleration,”
Adya, C. Affeldt, and et al. (2017), “Exploring the sen- Nature (London) 529, 373.
sitivity of next generation gravitational wave detectors,” Aidala, C, S. Bass, D. Hasch, and G. Mallot (2013), “The
Class. Quantum Gravity 34, 044001. spin structure of the nucleon,” Rev. Mod. Phys. 85, 655.
Abbott, L F, and P. Sikivie (1983), “A cosmological bound Aldaihan, S, D. E. Krause, J. C. Long, and W. M. Snow
on the invisible axion,” Phys. Lett. B 120, 133. (2017), “Calculations of the dominant long-range, spin-
Abe, M, G. Gopakumar, M. Hada, B. P. Das, H. Tate- independent contributions to the interaction energy be-
waki, and D. Mukherjee (2014), “Application of relativistic tween two nonrelativistic Dirac fermions from double-boson
coupled-cluster theory to the effective electric field in YbF,” exchange of spin-0 and spin-1 bosons with spin-dependent
Phys. Rev. A 90, 22501. couplings,” Phys. Rev. D 95, 096005.
Abele, H, T. Jenke, H. Leeb, and J. Schmiedmayer (2010), Allen, B, W. G. Anderson, P. R. Brady, D. A. Brown, and
“Ramsey’s method of separated oscillating fields and its ap- J. D. E. Creighton (2012), “FINDCHIRP: an algorithm for
plication to gravitationally induced quantum phase shifts,” detection of gravitational waves from inspiraling compact
Phys. Rev. D 81, 065019. binaries,” Phys. Rev. D 85, 122006.
Abrahamyan, S, Z. Ahmed, H. Albataineh, K. Aniol, D. Arm- Allen, S W, R. W. Schmidt, A. C. Fabian, and H. Ebeling
strong, W. Armstrong, T. Averett, B. Babineau, A. Barbi- (2003), “Cosmological constraints from the local X-ray lu-
eri, V. Bellini, et al. (2012), “Measurement of the neutron minosity function of the most X-ray-luminous galaxy clus-
radius of 208 Pb through parity violation in electron scat- ters,” Mon. Not. R. Astron Soc. 342, 287.
tering,” Phys. Rev. Lett. 108, 112502. Allmendinger, F, W. Heil, S. Karpuk, W. Kilian, A. Scharth,
Ackerman, Lotty, Matthew R. Buckley, Sean M. Carroll, and U. Schmidt, A. Schnabel, Y. Sobolev, and K. Tull-
Marc Kamionkowski (2009), “Dark matter and dark radi- ney (2014a), “New limit on Lorentz-invariance- and CPT-
ation,” Phys. Rev. D 79, 023519. violating neutron spin interactions using a free-spin-
Acosta, V, M. P. Ledbetter, S. M. Rochester, D. Budker, D. F. precession 3 He - 129 Xe comagnetometer,” Phys. Rev. Lett.
Jackson Kimball, D. C. Hovde, W. Gawlik, S. Pustelny, 112, 110801.
J. Zachorowski, and V. V. Yashchuk (2006), “Nonlinear Allmendinger, F, U. Schmidt, W. Heil, S. Karpuk, A. Scharth,
magneto-optical rotation with frequency-modulated light Yu. Sobolev, and K. Tullney (2014b), “Allmendinger et al.
in the geophysical field range,” Phys. Rev. A 73, 053404. reply:,” Phys. Rev. Lett. 113, 188902.
Adam, R, P. A. R. Ade, N. Aghanim, Y. Akrami, M. I. R. Allred, J C, R. N. Lyman, T. W. Kornack, and M. V. Romalis
Alves, F. Argüeso, M. Arnaud, F. Arroja, M. Ashdown, (2002), “High-sensitivity atomic magnetometer unaffected
and J. Aumont et al. (Planck Collaboration) (2016), by spin-exchange relaxation,” Phys. Rev. Lett. 89, 130801.
“Planck 2015 results,” Astron. Astrophys. 594, A1–A28. Alonso, A M, B. S. Cooper, A. Deller, S. D. Hogan, and D. B.
Adelberger, E G, J. H. Gundlach, B. R. Heckel, S. Hoedl, and Cassidy (2015), “Controlling positronium annihilation with
S. Schlamminger (2009), “Torsion balance experiments: A electric fields,” Phys. Rev. Lett. 115, 183401.
low-energy frontier of particle physics,” Prog. Part. Nucl. Altmann, R K, S. Galtier, L. S. Dreissen, and K. S. E. Eikema
Phys. 62, 102. (2016), “High-precision ramsey-comb spectroscopy at deep
Adelberger, E G, B. R. Heckel, S. Hoedl, C. D. Hoyle, D. J. ultraviolet wavelengths,” Phys. Rev. Lett. 117, 173201.
Kapner, and A. Upadhye (2007), “Particle-physics impli- Altschul, B (2006), “Limits on Lorentz violation from syn-
cations of a recent test of the gravitational inverse-square chrotron and inverse compton sources,” Phys. Rev. Lett.
law,” Phys. Rev. Lett. 98, 131104. 96, 201101.
Adelberger, E G, B. R. Heckel, and A. E. Nelson (2003), Altschul, B (2010), “Laboratory bounds on electron Lorentz
“Tests of the gravitational inverse-square law,” Annu. Rev. violation,” Phys. Rev. D 82, 016002.
Nucl. Part. Sci. 53, 77. Altschul, B, Q. G. Bailey, L. Blanchet, K. Bongs, P. Bouyer,
Adelberger, E G, B. R. Heckel, C. W. Stubbs, and Y. Su L. Cacciapuoti, S. Capozziello, N. Gaaloul, D. Giulini,
(1991), “Does antimatter fall with the same acceleration as J. Hartwig, et al. (2015), “Quantum tests of the Einstein
ordinary matter?” Phys. Rev. Lett. 66, 850. equivalence principle with the STE-QUEST space mis-
Adkins, Gregory S, Minji Kim, Christian Parsons, and sion,” Adv. Space Research 55, 501.
Richard N. Fell (2015), “Three-photon-annihilation contri- Amado, R D, and H. Primakoff (1980), “Comments on testing
butions to positronium energies at order mα7 ,” Phys. Rev. the Pauli principle,” Phys. Rev. C 22, 1338.
Lett. 115, 233401. Amaro, P, S. Schlesser, M. Guerra, E.-O. Le Bigot, J.-M.
Aghion, S, O. Ahlén, C. Amsler, A. Ariga, T. Ariga, A. S. Isac, P. Travers, J. P. Santos, C. I. Szabo, A. Gumberidze,
Belov, K. Berggren, G. Bonomi, P. Bräunig, J. Bremer, and P. Indelicato (2012), “Absolute measurement of the
et al. (2014), “A moiré deflectometer for antimatter,” Na- relativistic magnetic dipole transition energy in heliumlike
ture Commun. 5, 4538. argon,” Phys. Rev. Lett. 109, 043005.
Ahmadi, M, B. X. R. Alves, C. J. Baker, W. Bertsche, E. But- Amaro-Seoane, P, S. Aoudia, S. Babak, P. Binétruy, E. Berti,
ler, A. Capra, C. Carruth, C. L. Cesar, M. Charlton, S. Co- A. Bohé, C. Caprini, M. Colpi, N. J. Cornish, K. Danz-
hen, et al. (2017), “Observation of the 1s − 2s transition in mann, et al. (2012), “Low-frequency gravitational-wave
trapped antihydrogen,” Nature (London) 541, 506. science with eLISA/NGO,” Class. Quantum Gravity 29,
88

124016. Antypas, D, and D. S. Elliott (2014), “Measurement of weak


Amole, C, M. D. Ashkezari, M. Baquero-Ruiz, W. Bertsche, optical transition moments through two-pathway coherent
E. Butler, A. Capra, C. L. Cesar, M. Charlton, S. Eriksson, control,” Can. J. Chem. 92, 144.
J. Fajans, et al. (2012), “Resonant quantum transitions in Aoyama, T, M. Hayakawa, T. Kinoshita, and M. Nio (2008),
trapped antihydrogen atoms,” Nature (London) 483, 439. “Revised value of the eighth-order QED contribution to the
Amole, C, M. D. Ashkezari, M. Baquero-Ruiz, W. Bertsche, anomalous magnetic moment of the electron,” Phys. Rev.
E. Butler, A. Capra, C. L. Cesar, M. Charlton, S. Eriks- D 77, 053012.
son, J. Fajans, et al. (2014), “An experimental limit on the Appelquist, Thomas, Bogdan A. Dobrescu, and Adam R.
charge of antihydrogen,” Nature Commun. 5, 3955. Hopper (2003), “Nonexotic neutral gauge bosons,” Phys.
Amole, C, M. D. Ashkezari, M. Baquero-Ruiz, W. Bertsche, Rev. D 68, 035012.
E. Butler, A. Capra, C. L. Cesar, M. Charlton, S. Eriks- Arkani-Hamed, N, S. Dimopoulos, and G. R. Dvali (1998),
son, et al. (ALPHA Collaboration) (2013), “Description “The hierarchy problem and new dimensions at a millime-
and first application of a new technique to measure the ter,” Phys. Lett. B 429, 263.
gravitational mass of antihydrogen,” Nature Commun. 4, Arkani-Hamed, N, D. P. Finkbeiner, T. R. Slatyer, and
1785. N. Weiner (2009a), “A theory of dark matter,” Phys. Rev.
Anderson, J D, G. Schubert, V. Trimble, and M. R. Feldman D 79, 015014.
(2015), “Measurements of Newton’s gravitational constant Arkani-Hamed, Nima, Hsin-Chia Cheng, Markus A. Luty,
and the length of day,” Europhys. Lett. 110, 10002. and Shinji Mukohyama (2004), “Ghost condensation and
Anderson, W G, P. R. Brady, J. D. E. Creighton, and E. E. a consistent infrared modification of gravity,” J. High En-
Flanagan (2001), “Excess power statistic for detection of ergy Phys. 5, 074.
burst sources of gravitational radiation,” Phys. Rev. D 63, Arkani-Hamed, Nima, Hsin-Chia Cheng, Markus A. Luty,
042003. and Jesse Thaler (2005), “Universal dynamics of sponta-
Andreas, S (2012), “Update on hidden sectors with dark forces neous Lorentz violation and a new spin-dependent inverse-
and dark matter,” in Proceedings of the 8th Patras Work- square law force,” J. High Energy Phys. 7, 029.
shop on Axions, WIMPs and WISPs, DESY-PROC-2012- Arkani-Hamed, Nima, and Savas Dimopoulos (2005), “Super-
04, arxiv:1211.5160. symmetric unification without low energy supersymmetry
Andresen, G B, M. D. Ashkezari, M. Baquero-Ruiz, and signatures for fine-tuning at the LHC,” J. High Energy
W. Bertsche, P. D. Bowe, E. Butler, C. L. Cesar, M. Charl- Phys. 2005, 73.
ton, A. Deller, et al. (2011), “Confinement of antihydrogen Arkani-Hamed, Nima, Savas Dimopoulos, and Gia Dvali
for 1,000 seconds,” Nature Phys. 7, 558. (1999), “Phenomenology, astrophysics, and cosmology of
Androic, D, D. S. Armstrong, A. Asaturyan, T. Averett, theories with submillimeter dimensions and TeV scale
J. Balewski, J. Beaufait, R. S. Beminiwattha, J. Benesch, quantum gravity,” Phys. Rev. D 59, 086004.
F. Benmokhtar, J. Birchall, et al. (2013), “First determi- Arkani-Hamed, Nima, Douglas P. Finkbeiner, Tracy R.
nation of the weak charge of the proton,” Phys. Rev. Lett. Slatyer, and Neal Weiner (2009b), “A theory of dark mat-
111, 141803. ter,” Phys. Rev. D 79, 15014.
Ansel’m, A A (1982), “Possible new long-range interaction Arndt, M, and K. Hornberger (2014), “Testing the limits
and methods for detecting it,” Pis’ma Zh. Eksp. Teor. Fiz. of quantum mechanical superpositions,” Nature Phys. 10,
36, 46 [JETP Lett. 36, 46 (1982)]. 271.
Anthony et al., P L (1996), “Deep inelastic scattering of po- Arnison, G, A. Astbury, B. Aubert, C. Bacci, G. Bauer,
larized electrons by polarized 3 He and the study of the A. Bézaguet, R. Böck, T. J V Bowcock, M. Calvetti,
neutron spin structure,” Phys. Rev. D 54, 6620. T. Carroll, et al. (1983a), “Experimental observation of
Antognini, Aldo, François Nez, Karsten Schuhmann, Fer- isolated large transverse
√ energy electrons with associated
nando D. Amaro, François Biraben, João M. R. Cardoso, missing energy at s = 540 Gev,” Phys. Lett. B 122, 103.
Daniel S. Covita, Andreas Dax, Satish Dhawan, Marc Arnison, G, A. Astbury, B. Aubert, C. Bacci, G. Bauer,
Diepold, et al. (2013), “Proton structure from the measure- A. Bézaguet, R. Böck, T. J. V. Bowcock, M. Calvetti,
ment of 2s-2p transition frequencies of muonic hydrogen,” P. Catz, et al. (1983b), “Experimental observation of lep-
Science 339, 417. ton pairs of invariant mass around 95 Gev/c2 at the CERN
Antoniadis, I, S. Baessler, M. Büchner, V. V. Fedorov, SPS collider,” Phys. Lett. B 126, 398.
S. Hoedl, A. Lambrecht, V. V. Nesvizhevsky, G. Pig- Arrington, John, and Ingo Sick (2015), “Evaluation of the
nol, K. V. Protasov, S. Reynaud, and Yu. Sobolev proton charge radius from electronproton scattering,” J.
(2011), “Short-range fundamental forces,” Comptes Ren- Phys. Chem. Ref. Dat. 44, 031204.
dus Physique 12, 755. Artemyev, A N, V. M. Shabaev, I. I. Tupitsyn, G. Plu-
Antoniadis, I, S. Dimopoulos, and G. Dvali (1998), nien, and V. A. Yerokhin (2007), “QED calculation of the
“Millimetre-range forces in superstring theories with weak- 2p3/2 − 2p1/2 transition energy in boronlike argon,” Phys.
scale compactification,” Nucl. Phys. B 516, 70. Rev. Lett. 98, 173004.
Antypas, D, and D. S. Elliott (2011), “Measurements of the Artemyev, A N, V. M. Shabaev, V. A. Yerokhin, G. Plunien,
static polarizability of the 8s 2 S1/2 state of atomic cesium,” and G. Soff (2005), “QED calculation of the n = 1 and
Phys. Rev. A 83, 062511. n = 2 energy levels in He-like ions,” Phys. Rev. A 71,
Antypas, D, and D. S. Elliott (2013a), “Measurement of a 062104.
weak transition moment using two-pathway coherent con- Arvanitaki, A, S. Dimopoulos, S. Dubovsky, N. Kaloper, and
trol,” Phys. Rev. A 87, 42505. J. March-Russell (2010), “String axiverse,” Phys. Rev. D
Antypas, D, and D. S. Elliott (2013b), “Measurement of the 81, 123530.
radial matrix elements of the 6s 2 S1/2 → 7p 2 PJ transitions Arvanitaki, A, and A. A. Geraci (2014), “Resonantly de-
in atomic cesium,” Phys. Rev. A 88, 52516. tecting axion-mediated forces with nuclear magnetic reso-
89

nance,” Phys. Rev. Lett. 113, 161801. Banks, T, and A. Zaks (1982), “On the phase structure of
Arvanitaki, A, J. Huang, and K. Van Tilburg (2015), “Search- vector-like gauge theories with massless fermions,” Nucl.
ing for dilaton dark matter with atomic clocks,” Phys. Rev. Phys. B 196, 189.
D 91, 015015. Barbieri, R, M. Cerdonio, G. Fiorentini, and S. Vitale (1989),
Ashby, N, T. P. Heavner, S. R. Jefferts, T. E. Parker, A. G. “Axion to magnon conversion. A scheme for the detection
Radnaev, and Y. O. Dudin (2007), “Testing local po- of galactic axions,” Phys. Lett. B 226, 357.
sition invariance with four cesium-fountain primary fre- Barger, V, T. Han, C. Kao, and R.-J. Zhang (1999), “As-
quency standards and four NIST hydrogen masers,” Phys. trophysical constraints on large extra dimensions,” Phys.
Rev. Lett. 98, 070802. Lett. B 461, 34.
Aspelmeyer, M, T. J. Kippenberg, and F. Marquardt (2014), Barkov, L M, and M. S. Zolotorev (1978), “Measuring optical
“Cavity optomechanics,” Rev. Mod. Phys. 86, 1391. activity of bismuth vapour,” Pisma Zh. Eksp. Teor. Fiz 28,
Asztalos, S, E. Daw, H. Peng, L. J. Rosenberg, C. Hagmann, 544 [JETP Lett. 28, 503 (1978)].
D. Kinion, W. Stoeffl, K. van Bibber, P. Sikivie, N. S. Sul- Baron, J, W. C. Campbell, D. DeMille, J. M. Doyle,
livan, et al. (2001), “Large-scale microwave cavity search G. Gabrielse, Y. V. Gurevich, P. W. Hess, N. R. Hutzler,
for dark-matter axions,” Phys. Rev. D 64, 092003. E. Kirilov, I. Kozyryev, et al. (2014), “Order of magnitude
Asztalos, S J, G. Carosi, C. Hagmann, D. Kinion, K. van smaller limit on the electric dipole moment of the electron.”
Bibber, M. Hotz, L. J. Rosenberg, G. Rybka, J. Hoskins, Science 343, 269.
J. Hwang, P. Sikivie, D. B. Tanner, R. Bradley, and Barr, S M (1992a), “Measurable T- and P-odd electron-
J. Clarke (2010), “Squid-based microwave cavity search for nucleon interactions from Higgs-boson exchange,” Phys.
dark-matter axions,” Phys. Rev. Lett. 104, 041301. Rev. Lett. 68, 1822.
Aubin, S, J. A. Behr, R. Collister, V. V. Flambaum, Barr, S M (1993), “A review of CP violation in atoms,” Int.
E. Gomez, G. Gwinner, K. P. Jackson, D. Melconian, L. A. J. Mod. Phys. A 8, 209.
Orozco, M. R. Pearson, et al. (2013), “Atomic parity non- Barr, S M, and A. Zee (1990), “Electric dipole moment of
conservation: the francium anapole project of the FrPNC the electron and of the neutron,” Phys. Rev. Lett. 65, 21.
collaboration at TRIUMF,” Hyperfine Interact. 214, 163. Barr, Stephen M (1992b), “T-and P-odd electron-nucleon in-
Auerbach, N, V. V. Flambaum, and V. Spevak (1996), “Col- teractions and the electric dipole moments of large atoms,”
lective T- and P-odd electromagnetic moments in nuclei Phys. Rev. D 45, 4148.
with octupole deformations,” Phys. Rev. Lett. 76, 4316. Barranco, J, and A. Bernal (2011), “Self-gravitating system
Avelino, P P, C. J. A. P. Martins, N. J. Nunes, and K. A. made of axions,” Phys. Rev. D 83, 043525.
Olive (2006), “Reconstructing the dark energy equation of Bassi, A, K. Lochan, S. Satin, T. P. Singh, and H. Ulbricht
state with varying couplings,” Phys. Rev. D 74, 083508. (2013), “Models of wave-function collapse, underlying the-
Bagdonaite, J, M. Daprà, P. Jansen, H. L. Bethlem, ories, and experimental tests,” Rev. Mod. Phys. 85, 471.
W. Ubachs, S. Muller, C. Henkel, and K. M. Menten Beck, B R, J. A. Becker, P. Beiersdorfer, G. V. Brown, K. J.
(2013a), “Robust constraint on a drifting proton-to- Moody, J. B. Wilhelmy, F. S. Porter, C. A. Kilbourne, and
electron mass ratio at z = 0.89 from methanol observation R. L. Kelley (2007), “Energy splitting of the ground-state
at three radio telescopes,” Phys. Rev. Lett. 111, 231101. doublet in the nucleus 229 Th,” Phys. Rev. Lett. 98, 142501.
Bagdonaite, J, P. Jansen, C. Henkel, H. L. Bethlem, K. M. Beck, B R, C.Y. Wu, P. Beiersdorfer, G. V. Brown, J. A.
Menten, and W. Ubachs (2013b), “A stringent limit on a Becker, K. J. Moody, J. B. Wilhelmy, F. S. Porter, C. A.
drifting proton-to-electron mass ratio from alcohol in the Kilbourne, and R. L. Kelley (2009), “Improved value for
early universe,” Science 339, 46. the energy splitting of the ground-state doublet in the nu-
Bagdonaite, J, E. J. Salumbides, S. P. Preval, M. A. Barstow, cleus 229mTh,” LLNL-PROC-415170.
J. D. Barrow, M. T. Murphy, and W. Ubachs (2014), Beiersdorfer, P (2010), “Testing QED and atomic-nuclear in-
“Limits on a gravitational field dependence of the proton- teractions with high-Z ions,” J. Phys. B 43, 074032.
electron mass ratio from H2 in white dwarf stars,” Phys. Beiersdorfer, P, E. Träbert, G. V. Brown, J. Clementson,
Rev. Lett. 113, 123002. D. B. Thorn, M. H. Chen, K. T. Cheng, and J. Sapirstein
Bailin, David, and Alex Love (1987), “Kaluza-Klein theo- (2014), “Hyperfine splitting of the 2s1/2 and 2p1/2 levels
ries,” Rep. Prog. Phys. 50, 1087. in Li- and Be-like ions of 14159 Pr,” Phys. Rev. Lett. 112,
Baker, C A, D. D. Doyle, P. Geltenbort, K. Green, M. G. D. 233003.
van der Grinten, P. G. Harris, P. Iaydjiev, S. N. Ivanov, Beiersdorfer, P, S. B. Utter, K. L. Wong, J. R. Crespo López-
D. J. R. May, J. M. Pendlebury, J. D. Richardson, Urrutia, J. A. Britten, H. Chen, C. L. Harris, R. S. Thoe,
D. Shiers, and K. F. Smith (2006), “Improved experimen- D. B. Thorn, E. Träbert, M. G. H. Gustavsson, C. Forssén,
tal limit on the electric dipole moment of the neutron,” and A.-M. Mårtensson-Pendrill (2001), “Hyperfine struc-
Phys. Rev. Lett. 97, 131801. ture of hydrogenlike thallium isotopes,” Phys. Rev. A 64,
Baker, J G, and J. I. Thorpe (2012), “Comparison of Atom 032506.
Interferometers and Light Interferometers as Space-Based Beloy, K, A. Borschevsky, P. Schwerdtfeger, and V. V.
Gravitational Wave Detectors,” Phys. Rev. Lett. 108 (21), Flambaum (2010), “Enhanced sensitivity to the time vari-
211101. ation of the fine-structure constant and mp /me in diatomic
Balazs, C, G. White, and J. Yue (2017), “Effective field the- molecules: A closer examination of silicon monobromide,”
ory, electric dipole moments and electroweak baryogene- Phys. Rev. A 82, 022106.
sis,” J. High Energy Phys. 3, 30. Bender, P L (2011), “Comment on “Atomic gravitational
Ban, Shufang, Jacek Dobaczewski, Jonathan Engel, and wave interferometric sensor”,” Phys. Rev. D 84, 028101.
A. Shukla (2010), “Fully self-consistent calculations of nu- Bender, P L (2014), “Comparison of atom interferometry with
clear Schiff moments,” Phys. Rev. C 82, 15501. laser interferometry for gravitational wave observations in
90

space,” Phys. Rev. D 89, 062004. 542, A118.


Bennett, G W, B. Bousquet, H. N. Brown, G. Bunce, R. M. Bergé, J, P. Touboul, M. Rodrigues, and for the
Carey, P. Cushman, G. T. Danby, P. T. Debevec, M. Deile, MICROSCOPE team (2015), “Status of MICROSCOPE,
et al. (2006), “Final report of the E821 muon anomalous a mission to test the equivalence principle in space,” in
magnetic moment measurement at BNL,” Phys. Rev. D 73, Journal of Physics Conference Series, Vol. 610, p. 012009.
072003. Bergström, L, P. Ullio, and J. H. Buckley (1998), “Observ-
Bennett, S C, and C. E. Wieman (1999), “Measurement of ability of γ rays from dark matter neutralino annihilations
the 6S → 7S transition polarizability in atomic cesium and in the milky way halo,” Astroparticle Physics 9, 137.
an improved test of the standard model,” Phys. Rev. Lett. Bernauer, J C, P. Achenbach, C. Ayerbe Gayoso, R. Böhm,
82, 2484. D. Bosnar, L. Debenjak, M. O. Distler, L. Doria, A. Esser,
Berengut, J C, Dmitry Budker, Cedric Delaunay, Victor V. H. Fonvieille, et al. (A1 Collaboration) (2010), “High-
Flambaum, Claudia Frugiuele, Elina Fuchs, Christophe precision determination of the electric and magnetic form
Grojean, Roni Harnik, Roee Ozeri, Gilad Perez, et al. factors of the proton,” Phys. Rev. Lett. 105, 242001.
(2017), “Probing new light force-mediators by isotope shift Bernreuther, Werner, and Mahiko Suzuki (1991), “The elec-
spectroscopy,” arXiv:1704.05068. tric dipole moment of the electron,” Rev. Mod. Phys. 63,
Berengut, J C, V. A. Dzuba, and V. V. Flambaum (2010), 313.
“Enhanced laboratory sensitivity to variation of the fine- Bertone, G, Ed. (2013), Particle Dark Matter: Observations,
structure constant using highly-charged ions,” Phys. Rev. Models and Searches (Cambridge University, Cambridge,
Lett. 105, 120801. UK).
Berengut, J C, V. A. Dzuba, and V. V. Flambaum (2011a), Bertone, G, D. Hooper, and J. Silk (2005), “Particle dark
“Transitions in Zr, Hf, Ta, W, Re, Hg, Ac and U ions with matter: Evidence, candidates and constraints,” Phys. Rep.
high sensitivity to variation of the fine structure constant,” 405, 279.
Phys. Rev. A 84, 054501. Bezerra, V B, G. L. Klimchitskaya, V. M. Mostepanenko, and
Berengut, J C, V. A. Dzuba, V. V. Flambaum, J. A. King, C. Romero (2011), “Constraints on non-Newtonian grav-
M. G. Kozlov, M. T. Murphy, and J. K. Webb (2011b), ity from measuring the Casimir force in a configuration
“Atomic transition frequencies, isotope shifts, and sensi- with nanoscale rectangular corrugations,” Phys. Rev. D 83,
tivity to variation of the fine structure constant for studies 075004.
of quasar absorption spectra,” Astrophys. Space Sci. Proc. van Bibber, K, and G. Carosi (2013), “Status of the ADMX
22, 9. and ADMX-HF experiments,” in Proceedings of the 8th
Berengut, J C, V. A. Dzuba, V. V. Flambaum, and Patras Workshop on Axions, WIMPs and WISPs, DESY-
A. Ong (2011c), “Hole transitions in multiply-charged ions PROC-2012-04, arXiv:1304.7803.
for precision laser spectroscopy and searching for alpha- Biedermann, G W, X. Wu, L. Deslauriers, S. Roy, C. Ma-
variation,” Phys. Rev. Lett. 106, 210802. hadeswaraswamy, and M. A. Kasevich (2015), “Testing
Berengut, J C, V. A. Dzuba, V. V. Flambaum, and A. Ong gravity with cold-atom interferometers,” Phys. Rev. A 91,
(2012a), “Highly charged ions with E1, M1, and E2 transi- 033629.
tions within laser range,” Phys. Rev. A 86, 022517. Biesheuvel, J, J.-P. Karr, L. Hilico, K. S. E. Eikema,
Berengut, J C, V. A. Dzuba, V. V. Flambaum, and A. Ong W. Ubachs, and J. C. J. Koelemeij (2016), “Probing QED
(2012b), “Optical transitions in highly-charged californium and fundamental constants through laser spectroscopy
ions with high sensitivity to variation of the fine-structure of vibrational transitions in HD+ ,” Nature Commun. 7,
constant,” Phys. Rev. Lett. 109, 070802. 10385.
Berengut, J C, V. A. Dzuba, V. V. Flambaum, and S. G. BIPM, (2014), SI Brochure: The International System of
Porsev (2009), “Proposed experimental method to deter- Units (SI) [8th edition, 2006; updated in 2014], Avail-
mine α sensitivity of splitting between ground and 7.6 eV able online: http://www.bipm.org/en/publications/si-
isomeric states in 229 Th,” Phys. Rev. Lett. 102, 210801. brochure.
Berengut, J C, and V. V. Flambaum (2012), “Manifesta- Bishof, Michael, Richard H. Parker, Kevin G. Bailey, John P.
tions of a spatial variation of fundamental constants on Greene, Roy J. Holt, Mukut R. Kalita, Wolfgang Ko-
atomic clocks, Oklo, meteorites, and cosmological phenom- rsch, Nathan D. Lemke, Zheng-Tian Lu, Peter Mueller,
ena,” Europhys. Lett. 97, 20006. Thomas P. O’Connor, Jaideep T. Singh, and Matthew R.
Berengut, J C, V. V. Flambaum, J. A. King, S. J. Curran, Dietrich (2016), “Improved limit on the 225 Ra electric
and J. K. Webb (2011d), “Is there further evidence for spa- dipole moment,” Phys. Rev. C 94, 025501.
tial variation of fundamental constants?” Phys. Rev. D 83, Bjorken, James D, and S. D. Drell (1964), Relativistic Quan-
123506. tum Mechanics (McGraw-Hill).
Berengut, J C, V. V. Flambaum, and A. Ong (2013a), “Test- Blatt, J M, and V. F. Weisskopf (1979), Theoretical Nuclear
ing spatial α-variation with optical atomic clocks based on Physics (Springer, New York).
highly charged ions,” in European Physical Journal Web of Blatt, S, A. D. Ludlow, G. K. Campbell, J. W. Thomsen,
Conferences, Vol. 57, p. 2001. T. Zelevinsky, M. M. Boyd, J. Ye, X. Baillard, M. Fouché,
Berengut, J C, V. V. Flambaum, A. Ong, J. K. Webb, J. D. R. Le Targat, A. Brusch, P. Lemonde, M. Takamoto, F.-L.
Barrow, M. A. Barstow, S. P. Preval, and J. B. Holberg Hong, H. Katori, and V. V. Flambaum (2008), “New limits
(2013b), “Limits on the dependence of the fine-structure on coupling of fundamental constants to gravity using 87 Sr
constant on gravitational potential from white-dwarf spec- optical lattice clocks,” Phys. Rev. Lett. 100, 140801.
tra,” Phys. Rev. Lett. 111, 010801. Bloch, I, J. Dalibard, and S. Nascimbène (2012), “Quantum
Berengut, J C, E. M. Kava, and V. V. Flambaum (2012c), simulations with ultracold quantum gases,” Nat. Phys. 8,
“Is there a spatial gradient in values of the fine-structure 267.
constant? A reanalysis of the results,” Astron. Astrophys.
91

Bloch, I, J. Dalibard, and W. Zwerger (2008), “Many-body to a precise measurement of the 6,7 Li D2 lines [Phys. Rev.
physics with ultracold gases,” Rev. Mod. Phys. 80, 885. A 87, 032504 (2013)],” Phys. Rev. A 88, 069902.
Bluhm, R, V. A. Kostelecký, and N. Russell (1997), “Testing Brown, R C, S. Wu, J. V. Porto, C. J. Sansonetti, C. E.
CPT with anomalous magnetic moments,” Phys. Rev. Lett. Simien, S. M. Brewer, J. N. Tan, and J. D. Gillaspy
79, 1432. (2013b), “Quantum interference and light polarization ef-
Blundell, S A, W. R. Johnson, and J. Sapirstein (1990), fects in unresolvable atomic lines: Application to a pre-
“High-accuracy calculation of the 6S1/2 to 7S1/2 parity- cise measurement of the 6,7 Li D2 lines,” Phys. Rev. A 87,
nonconserving transition in atomic cesium and implications 032504.
for the standard model,” Phys. Rev. Lett. 65, 1411. Brubaker, B M, L. Zhong, Y. V. Gurevich, S. B. Cahn, S. K.
Bonnin, A, N. Zahzam, Y. Bidel, and A. Bresson (2013), Lamoreaux, M. Simanovskaia, J. R. Root, S. M. Lewis,
“Simultaneous dual-species matter-wave accelerometer,” S. Al Kenany, K. M. Backes, et al. (2016), “First results
Phys. Rev. A 88, 043615. from a microwave cavity axion search at 24 micro-eV,”
Botermann, Benjamin, Dennis Bing, Christopher Geppert, arXiv:1610.02580.
Gerald Gwinner, Theodor W Hänsch, Gerhard Huber, Bruhns, H, J. Braun, K. Kubiček, J. R. Crespo López-Urrutia,
Sergei Karpuk, Andreas Krieger, Thomas Kühl, Wilfried and J. Ullrich (2007), “Testing QED screening and two-
Nörtershäuser, et al. (2014), “Test of time dilation using loop contributions with He-like ions,” Phys. Rev. Lett. 99,
stored li+ ions as clocks at relativistic speed,” Phys. Rev. 113001.
Lett. 113, 120405. Bubin, Sergiy, Oleg V. Prezhdo, and Kálmán Varga (2013),
Bouchendira, R, P. Cladé, S. Guellati-Khélifa, F. Nez, and “Instability of tripositronium,” Phys. Rev. A 87, 054501.
F. Biraben (2011), “New determination of the fine structure Buckley, Matthew R, and Michael J Ramsey-Musolf (2012),
constant and test of the quantum electrodynamics,” Phys. “Precision probes of a leptophobic boson,” Phys. Lett. B
Rev. Lett. 106, 080801. 712 (3), 261–265.
Bouchiat, C, and C. A. Piketty (1983), “Parity violation in Budker, D, and A. Derevianko (2015), “A data archive for
atomic cesium and alternatives to the standard model of storing precision measurements,” Physics Today 68.
electroweak interaction,” Phys. Lett. 128B, 73. Budker, D, W. Gawlik, D. F. Kimball, S. M. Rochester,
Bouchiat, C, and C. A. Piketty (1991), “Nuclear spin depen- V. V. Yashchuk, and A. Weis (2002), “Resonant nonlin-
dent atomic parity violation, nuclear anapole moments and ear magneto-optical effects in atoms,” Rev. Mod. Phys. 74,
the hadronic axial neutral current,” Z. Phys. C 49, 91. 1153.
Bouchiat, M A, and C. Bouchiat (1974), “Parity violation Budker, D, P. W. Graham, M. Ledbetter, S. Rajendran, and
induced by weak neutral currents in atomic physics. I,” J. A. O. Sushkov (2014), “Proposal for a cosmic axion spin
Phys. 35, 899. precession experiment (CASPEr),” Phys. Rev. X 4, 021030.
Bouchiat, M A, and C. Bouchiat (1975), “Parity violation Budker, D, and D. F. Jackson Kimball, Eds. (2013), Optical
induced by weak neutral currents in atomic physics. II,” Magnetometry (Cambridge University, Cambridge, UK).
Journal de Physique 36, 493. Budker, D, and M. Romalis (2007), “Optical magnetometry,”
Bouchiat, M-A, and C. Bouchiat (1997), “Parity violation in Nature Phys. 3, 227.
atoms,” Rep. Prog. Phys. 60, 1351. Budker, Dmitry (1999), “Parity Nonconservation in Atoms,”
Bouchiat, M A, J. Guena, L. Pottier, and L. Hunter (1984), in Physics Beyond the Standard Model, Proceedings of the
“New observation of a parity violation in cesium,” Phys. Fifth Intrnational WEIN Symposium, edited by P. Herczeg,
Lett. B 134, 463. C. M. Hoffman, and H. V. Klapdor-Kleingrothaus (World
Bouchiat, Marie-Anne (2007), “Linear Stark shift in dressed Scientific) p. 418.
atoms as a signal to measure a nuclear anapole moment Budker, Dmitry, Derek F. Kimball, and David P. DeMille
with a cold-atom fountain or interferometer,” Phys. Rev. (2008), Atomic physics: an exploration through problems
Lett. 98, 43003. and solutions (Oxford University, USA).
Braaten, E, A. Mohapatra, and H. Zhang (2016), “Dense Bulatowicz, M, R. Griffith, M. Larsen, J. Mirijanian, C. B.
axion stars,” Phys. Rev. Lett. 117, 121801. Fu, E. Smith, W. M. Snow, H. Yan, and T. G. Walker
Bradley, R, J. Clarke, D. Kinion, L. J. Rosenberg, K. van (2013), “Laboratory search for a long-range T-odd, P-odd
Bibber, S. Matsuki, M. Mück, and P. Sikivie (2003), “Mi- interaction from axionlike particles using dual-species nu-
crowave cavity searches for dark-matter axions,” Rev. Mod. clear magnetic resonance with polarized 129 Xe and 131 Xe
Phys. 75, 777. gas,” Phys. Rev. Lett. 111, 102001.
Brans, C, and R. H. Dicke (1961), “Mach’s principle and a Bunker, P. R.,, Schutte, C. J. H., Hougen, J. T., Mills, I. M.,
relativistic theory of gravitation,” Phys. Rev. 124, 925. Watson, J. K. G., and Winnewisser, B. P. (1997), “No-
Brown, B A, G. F. Bertsch, L. M. Robledo, M. V. Romalis, tations and conventions in molecular spectroscopy: Part
and V. Zelevinsky (2016), “Nuclear matrix elements for 3. Permutation and permutation-inversion symmetry nota-
tests of fundamental symmetries,” arXiv:1604.08187. tion (IUPAC Recommendations 1997),” Pure Appl. Chem.
Brown, B A, A. Derevianko, and V. V. Flambaum (2009), 69, 1651.
“Calculations of the neutron skin and its effect in atomic Cacciapuoti, L, N. Dimarcq, G. Santarelli, P. Laurent,
parity violation,” Phys. Rev. C 79, 035501. P. Lemonde, A. Clairon, P. Berthoud, A. Jornod, F. Reina,
Brown, J M, S. J. Smullin, T. W. Kornack, and M. V. Ro- S. Feltham, and C. Salomon (2007), “Atomic clock en-
malis (2010), “New limit on Lorentz- and CPT-violating semble in space: Scientific objectives and mission status,”
neutron spin interactions,” Phys. Rev. Lett. 105, 151604. Nucl. Phys. B Proc. Suppl. 166, 303.
Brown, R C, S. Wu, J. V. Porto, C. J. Sansonetti, C. E. Cairncross, W B, D. N. Gresh, M. Grau, K. C. Cossel, T. S.
Simien, S. M. Brewer, J. N. Tan, and J. D. Gillaspy Roussy, Y. Ni, Y. Zhou, J. Ye, and E. A. Cornell (2017),
(2013a), “Erratum: Quantum interference and light po- “A precision measurement of the electron’s electric dipole
larization effects in unresolvable atomic lines: Application moment using trapped molecular ions,” arXiv:1704.07928.
92

Campbell, C J, A. G. Radnaev, A. Kuzmich, V. A. Dzuba, J. N. Tan, J. A. Kimpton, E. Takacs, and K. Makonyi


V. V. Flambaum, and A. Derevianko (2012), “A single- (2012), “Testing three-body quantum electrodynamics with
ion nuclear clock for metrology at the 19th decimal place,” trapped Ti20+ ions: Evidence for a Z-dependent diver-
Phys. Rev. Lett. 108, 120802. gence between experiment and calculation,” Phys. Rev.
Cancio Pastor, P, I. Galli, G. Giusfredi, D. Mazzotti, and Lett. 109, 153001.
P. De Natale (2015), “Testing the validity of Bose-Einstein Charlton, M, and J. W. Humberston (2000), Positron
statistics in molecules,” Phys. Rev. A 92, 063820. Physics, by M. Charlton and J. W. Humberston,
Carlson, C E (2015), “The proton radius puzzle,” Progress in pp. 464. ISBN 0521415500. Cambridge, UK: Cambridge
Particle and Nuclear Physics 82, 59. University Press, December 2000.
Carroll, S M, and G. B. Field (1994), “Consequences of prop- Chatrchyan, S, V. Khachatryan, A. M. Sirunyan, A. Tu-
agating torsion in connection-dynamic theories of gravity,” masyan, W. Adam, E. Aguilo, T. Bergauer, M. Dragicevic,
Phys. Rev. D 50, 3867. J. Erö, C. Fabjan, M. Friedl, R. Frühwirth, and V. M.
Cartan, É (1922), “On a generalization of the riemann cur- Ghete (2012), “Observation of a new boson at a mass of
vature concept and space with torsion,” C. R. Acad. Sci. 125 GeV with the CMS experiment at the LHC,” Phys.
(Paris) 174, 593. Lett. B 716, 30.
Cartan, É (1923), “On varieties of affine connections and the Chaudhuri, S, P. W. Graham, K. Irwin, J. Mardon, S. Rajen-
theory of general relativity (part I),” Ann. Ec. Norm. Sup. dran, and Y. Zhao (2015), “Radio for hidden-photon dark
40, 325. matter detection,” Phys. Rev. D 92, 075012.
Cartan, É (1924), “On varieties of affine connections and the Chen, Q, E. Magoulakis, and S. Schiller (2016a), “High-
theory of general relativity (part I),” Ann. Ec. Norm. Sup. sensitivity crossed-resonator laser apparatus for improved
41, 1. tests of lorentz invariance and of space-time fluctuations,”
Cartan, É (1925), “On varieties of affine connections and the Phys. Rev. D 93, 022003.
theory of general relativity (part II),” Ann. Ec. Norm. Sup. Chen, Y-J, W. K. Tham, D. E. Krause, D. López, E. Fis-
42, 17. chbach, and R. S. Decca (2016b), “Stronger limits on hy-
Casalbuoni, R, S. de Curtis, D. Dominici, and R. Gatto pothetical Yukawa interactions in the 30–8000 nm range,”
(1999), “Bounds on new physics from the new data on par- Phys. Rev. Lett. 116, 221102.
ity violation in atomic cesium,” Phys. Lett. B 460, 135. Chengalur, J N, and N. Kanekar (2003), “Constraining the
Casola, Eolo Di, Stefano Liberati, and Sebastiano variation of fundamental constants using 18cm OH lines,”
Sonego (2015), “Nonequivalence of equivalence principles,” Phys. Rev. Lett. 91, 241302.
arXiv:1310.7426v2. Chikashige, Y, R. Mohapatra, and R. Peccei (1981), “Are
Cassidy, D B, T. H. Hisakado, H. W. K. Tom, and A. P. there real goldstone bosons associated with broken lepton
Mills (2012a), “Efficient production of rydberg positron- number?” Phys. Lett. 98B, 265.
ium,” Phys. Rev. Lett. 108, 043401. Chin, C, and V. V. Flambaum (2006), “Enhanced sensitivity
Cassidy, D B, T. H. Hisakado, H. W. K. Tom, and A. P. Mills to fundamental constants in ultracold atomic and molecular
(2012b), “Optical spectroscopy of molecular positronium,” systems near Feshbach resonances,” Phys. Rev. Lett. 96,
Phys. Rev. Lett. 108, 133402. 230801.
Cassidy, D B, T. H. Hisakado, H. W. K. Tom, and A. P. Chin, C, V. V. Flambaum, and M. G. Kozlov (2009), “Ultra-
Mills (2012c), “Positronium hyperfine interval measured cold molecules: new probes on the variation of fundamental
via saturated absorption spectroscopy,” Phys. Rev. Lett. constants,” New J. Phys. 11, 055048.
109, 073401. Chiow, S-W, J. Williams, and N. Yu (2015), “Laser-ranging
Cassidy, D B, and A. P. Mills (2007), “The production of long-baseline differential atom interferometers for space,”
molecular positronium,” Nature 449, 195. Phys. Rev. A 92, 063613.
Cavendish, H (1798), “Experiments to determine the density Cho, D, K. Sangster, and E. A. Hinds (1989), “Tenfold im-
of the Earth. By Henry Cavendish, Esq. F. R. S. and A. S.” provement of limits on T violation in thallium fluoride,”
Philosophical Transactions of the Royal Society of London Phys. Rev. Lett. 63 (23), 2559–2562.
Series I 88, 469. Cho, D, K Sangster, and EA Hinds (1991), “Search for time-
Chaibi, W, R. Geiger, B. Canuel, A. Bertoldi, A. Landragin, reversal-symmetry violation in thallium fluoride using a jet
and P. Bouyer (2016), “Low frequency gravitational wave source,” Phys. Rev. A 44, 2783.
detection with ground-based atom interferometer arrays,” Choi, J, and D. S. Elliott (2016), “Measurement scheme and
Phys. Rev. D 93, 021101. analysis for weak ground-state-hyperfine-transition mo-
Chand, H, R. Srianand, P. Petitjean, and B. Aracil (2004), ments through two-pathway coherent control,” Phys. Rev.
“Probing the cosmological variation of the fine-structure A 93, 1.
constant: Results based on VLT-UVES sample,” Astron. Chu, P-H, A. Dennis, C. B. Fu, H. Gao, R. Khatiwada,
Astrophys. 417, 853. G. Laskaris, K. Li, E. Smith, W. M. Snow, H. Yan, and
Chang, S, C. Hagmann, and P. Sikivie (1998), “Studies of W. Zheng (2013), “Laboratory search for spin-dependent
the motion and decay of axion walls bounded by strings,” short-range force from axionlike particles using optically
Phys. Rev. D 59, 023505. polarized 3 He gas,” Phys. Rev. D 87, 011105(R).
Chantler, C T, M. N. Kinnane, J. D. Gillaspy, L. T. Hudson, Chu, P-H, E. Weisman, C.-Y. Liu, and J. C. Long (2015),
A. T. Payne, L. F. Smale, A. Henins, J. M. Pomeroy, J. A. “Search for exotic short-range interactions using paramag-
Kimpton, E. Takacs, and K. Makonyi (2013), “Chantler et netic insulators,” Phys. Rev. D 91, 102006.
al. Reply:,” Phys. Rev. Lett. 110, 159302. Chu, S (1998), “Nobel lecture: The manipulation of neutral
Chantler, C T, M. N. Kinnane, J. D. Gillaspy, L. T. Hud- particles,” Rev. Mod. Phys. 70, 685.
son, A. T. Payne, L. F. Smale, A. Henins, J. M. Pomeroy, Chupp, Timothy, and Michael Ramsey-Musolf (2015), “Elec-
tric dipole moments: A global analysis,” Phys. Rev. C 91,
93

035502. Damour, T (2012), “Theoretical aspects of the equivalence


Cirigliano, Vincenzo, Yingchuan Li, Stefano Profumo, and principle,” Class. Quantum Gravity 29, 184001.
Michael J Ramsey-Musolf (2010), “MSSM baryogenesis Damour, T, F. Piazza, and G. Veneziano (2002), “Runaway
and electric dipole moments: an update on the phe- dilaton and equivalence principle violations,” Phys. Rev.
nomenology,” J. High Energy Phys. 2010, 1. Lett. 89, 081601.
Clowe, D, M. Bradač, A. H. Gonzalez, M. Markevitch, S. W. Damour, T, and A. M. Polyakov (1994), “The string dilation
Randall, C. Jones, and D. Zaritsky (2006), “A direct em- and a least coupling principle,” Nucl. Phys. B 423, 532.
pirical proof of the existence of dark matter,” Astrophys. Danzmann, K, T. Prince, P. Binetruy, P. Bender, S. Buch-
J. Lett. 648, L109. man, J. Centrella, M. Cerdonio, N. Cornish, A. Cruise,
Cocconi, Giuseppe, and E Salpeter (1958), “A search for C. Cutler, and et al. (2011), “LISA: Unveiling a hidden
anisotropy of inertia,” Il Nuovo Cimento 10, 646. universe,” Assessment Study Report ESA/SRE 3, 2.
Cohen-Tannoudji, C N (1998), “Nobel lecture: Manipulating Davis, R L (1985), “Goldstone bosons in string models of
atoms with photons,” Rev. Mod. Phys. 70, 707. galaxy formation,” Phys. Rev. D 32, 3172.
Colella, R, A. W. Overhauser, and S. A. Werner (1975), “Ob- Davoudiasl, Hooman, Hye Sung Lee, and William J. Mar-
servation of gravitationally induced quantum interference,” ciano (2014), “Muon g-2, rare kaon decays, and parity vio-
Phys. Rev. Lett. 34, 1472. lation from dark bosons,” Phys. Rev. D 89, 1, 1402.3620.
Coleman, S (1985), “Q-balls,” Nucl. Phys. B 262, 263. de Nijs, A J, W. Ubachs, and H. L. Bethlem (2014), “Ramsey-
Colladay, D, and V. A. Kostelecký (1998), “Lorentz-violating type microwave spectroscopy on CO (a3 Π),” J. Mol. Spec-
extension of the standard model,” Phys. Rev. D 58, 116002. trosc. 300, 79.
Commins, E D, and P. H. Bucksbaum (1983), Weak Interac- Decca, R S, D. López, H. B. Chan, E. Fischbach, D. E.
tions of Leptons and Quarks (Cambridge University, Cam- Krause, and C. R. Jamell (2005), “Constraining new forces
bridge - London - New York - New Rochelle - Melbourne - in the Casimir regime using the isoelectronic technique,”
Sydney). Phys. Rev. Lett. 94, 240401.
Commins, E D, and D. P. DeMille (2009), “The electric dipole Delaunay, C, C. Frugiuele, E. Fuchs, and Y. Soreq (2017),
moment of the electron,” in Lepton Dipole Moments, edited “Probing new spin-independent interactions through pre-
by B. L. Roberts and W. J. Marciano (World Scientific). cision spectroscopy in atoms with few electrons,” ArXiv
Commins, Eugene D (2012), “Electron spin and its history,” e-prints arXiv:1709.02817 [hep-ph].
Annu. Rev. Nucl. Part. Sci. 62, 133. Delaunay, C, R. Ozeri, G. Perez, and Y. Soreq (2016), “Prob-
Commins, Eugene D, J. D. Jackson, and David P. DeMille ing the atomic Higgs force,” arXiv:1601.05087.
(2007), “The electric dipole moment of the electron: An DeMille, D, F. Bay, S. Bickman, D. Kawall, L. Hunter,
intuitive explanation for the evasion of Schiff’s theorem,” D. Krause, S. Maxwell, and K. Ulmer (2001), “Search for
Am. J. Phys. 75, 532. the electric dipole moment of the electron using metastable
Conti, R, P. Bucksbaum, S. Chu, E. Commins, and L. Hunter PbO,” in Art and Symmetry in Experimental Physics, Vol.
(1979), “Preliminary observation of parity nonconservation 596 (American Institute of Physics, Melville, NY) p. 72.
in atomic thallium,” Phys. Rev. Lett. 42, 343. DeMille, D, D. Budker, N. Derr, and E. Deveney (1999),
Conti, R S, and I. B. Khriplovich (1992), “New limits on “Search for exchange-antisymmetric two-photon states,”
T-odd, P-even interactions,” Phys. Rev. Lett. 68, 3262. Phys. Rev. Lett. 83, 3978.
Cornell, E A, and C. E. Wieman (2002), “Nobel lecture: DeMille, D, S. B. Cahn, D. Murphree, D. A. Rahmlow, and
Bose-Einstein condensation in a dilute gas, the first 70 years M. G. Kozlov (2008a), “Using molecules to measure nu-
and some recent experiments,” Rev. Mod. Phys. 74, 875. clear spin-dependent parity violation,” Phys. Rev. Lett.
Crewther, R J, P. Di Vecchia, G. Veneziano, and E. Witten 100, 23003.
(1979), “Chiral estimate of the electric dipole moment of DeMille, D, S. Sainis, J. Sage, T. Bergeman, S. Kotochigova,
the neutron in quantum chromodynamics,” Phys. Lett. B and E. Tiesinga (2008b), “Enhanced sensitivity to varia-
88, 123. tion of me /mp in molecular spectra,” Phys. Rev. Lett. 100,
Cronin, A D, J. Schmiedmayer, and D. E. Pritchard (2009), 043202.
“Optics and interferometry with atoms and molecules,” DeMille, David (1995), “Parity Nonconservation in the
Rev. Mod. Phys. 81, 1051. 6s 21 S0 → 6s5d 3 D1 Transition in Atomic Ytterbium,”
Cullen, S, and M. Perelstein (1999), “SN1987A constraints Phys. Rev. Lett. 74, 4165.
on large compact dimensions,” Phys. Rev. Lett. 83, 268. Denis, Malika, and Timo Fleig (2016), “In search of discrete
Curceanu, C, J. D. Gillaspy, and R. C. Hilborn (2012), “Re- symmetry violations beyond the standard model: Thorium
source letter SS-1: The spin-statistics connection,” Am. J. monoxide reloaded,” J. Chem. Phys. 145, 214307.
Phys. 80, 561. Derbin, A V, A. I. Egorov, I. A. Mitropol’sky, V. N. Mura-
Cvetic, M, and P. Langacker (1996), “Implications of Abelian tova, D. A. Semenov, and E. V. Unzhakov (2009), “Search
extended gauge structures from string models,” Phys. Rev. for resonant absorption of solar axions emitted in M1 tran-
D 54, 3570. sition in 57 Fe nuclei,” Eur. Phys. J. C 62, 755.
Czarnecki, Andrzej, and Robert Szafron (2016), “Light-by- Derevianko, A (2000), “Reconciliation of the measurement
light scattering in the Lamb shift and the bound electron of parity-nonconservation in Cs with the standard model,”
g factor,” Phys. Rev. A 94, 060501(R). Phys. Rev. Lett. 85, 1618.
Dahia, F, and A. S. Lemos (2016a), “Constraints on extra Derevianko, A, V. A. Dzuba, and V. V. Flambaum (2012),
dimensions from atomic spectroscopy,” Phys. Rev. D 94, “Highly charged ions as a basis of optical atomic clockwork
084033. of exceptional accuracy,” Phys. Rev. Lett. 109, 180801.
Dahia, F, and A. S. Lemos (2016b), “Is the proton radius Derevianko, A, and M. Pospelov (2014), “Hunting for topo-
puzzle evidence of extra dimensions?” Euro. Phys. J. C 76, logical dark matter with atomic clocks,” Nature Phys. 10,
435. 933.
94

Derevianko, Andrei (2016), “Detecting dark matter waves Drake, G W F, Mark M. Cassar, and Razvan A. Nistor
with a network of precision measurement tools,” (2002), “Ground-state energies for helium, h− , and ps− ,”
arXiv:1605.09717. Phys. Rev. A 65, 054501.
Derevianko, Andrei, and S. G. Porsev (2007), “Theoretical Drake, G W F, and Z.-C. Yan (2008), “High-precision spec-
overview of atomic parity violation: Recent developments troscopy as a test of quantum electrodynamics in light
and challenges,” Eur. Phys. J. A 32, 517. atomic systems,” Can. J. Phys. 86, 45.
Derevianko, Andrei, and Sergey G. Porsev (2002), “Reevalu- Drever, R W P (1961), “A search for anisotropy of inertial
ation of the role of nuclear uncertainties in experiments on mass using a free precession technique,” Philosophical Mag-
atomic parity violation with isotopic chains,” Phys. Rev. A azine 6, 683.
65, 52115. Duan, Xiao-Chun, Xiao-Bing Deng, Min-Kang Zhou,
Desplanques, Bertrand, John F. Donoghue, and Barry R. Ke Zhang, Wen-Jie Xu, Feng Xiong, Yao-Yao Xu, Cheng-
Holstein (1980), “Unified treatment of the parity violating Gang Shao, Jun Luo, and Zhong-Kun Hu (2016), “Test
nuclear force,” Annals of Physics 124, 449. of the universality of free fall with atoms in different spin
Deutsch, Martin (1951), “Evidence for the formation of orientations,” Phys. Rev. Lett. 117, 023001.
positronium in gases,” Phys. Rev. 82, 455. Dubielzig, T, M. Niemann, A.-G. Paschke, M. Carsjens,
Di Piazza, A, C. Müller, K. Z. Hatsagortsyan, and C. H. M. Kohnen, and C. Ospelkaus (2013), “Quantum logic
Keitel (2012), “Extremely high-intensity laser interactions enabled test of discrete symmetries,” in APS Division of
with fundamental quantum systems,” Rev. Mod. Phys. 84, Atomic, Molecular and Optical Physics Meeting Abstracts.
1177. Duffy, L D, and P. Sikivie (2008), “Caustic ring model of the
Dickenson, G D, M. L. Niu, E. J. Salumbides, J. Komasa, MilkyWay halo,” Phys. Rev. D 78, 063508.
K. S. E. Eikema, K. Pachucki, and W. Ubachs (2013), Duffy, Leanne D, and Karl van Bibber (2009), “Axions as
“Fundamental vibration of molecular hydrogen,” Phys. dark matter particles,” New J. Phys. 11, 105008.
Rev. Lett. 110, 193601. Dufour, G, D. B. Cassidy, P. Crivelli, P. Debu, A. Lambrecht,
Dimopoulos, S, and H. Georgi (1981), “Softly broken super- V. V. Nesvizhevsky, S. Reynaud, A. Y. Voronin, and T. E.
symmetry and SU(5),” Nucl. Phys. B 193, 150. Wall (2015), “Prospects for studies of the free fall and grav-
Dimopoulos, S, and A. A. Geraci (2003), “Probing submi- itational quantum states of antimatter,” Advances in High
cron forces by interferometry of Bose-Einstein condensed Energy Physics 2015, 379642.
atoms,” Phys. Rev. D 68, 124021. Dzuba, V A, J. C. Berengut, V. V. Flambaum, and
Dimopoulos, S, P. W. Graham, J. M. Hogan, and M. A. B. Roberts (2012a), “Revisiting parity nonconservation in
Kasevich (2007), “Testing general relativity with atom in- cesium,” Phys. Rev. Lett. 109, 203003.
terferometry,” Phys. Rev. Lett. 98, 111102. Dzuba, V A, A. Derevianko, and V. V. Flambaum (2012b),
Dimopoulos, S, P. W. Graham, J. M. Hogan, and M. A. “Ion clock and search for the variation of the fine structure
Kasevich (2008a), “General relativistic effects in atom in- constant using optical transitions in Nd13+ and Sm15+ ,”
terferometry,” Phys. Rev. D 78, 042003. Phys. Rev. A 86, 054502.
Dimopoulos, S, P. W. Graham, J. M. Hogan, M. A. Kasevich, Dzuba, V A, A. Derevianko, and V. V. Flambaum
and S. Rajendran (2008b), “Atomic gravitational wave in- (2013), “Erratum: High-precision atomic clocks with highly
terferometric sensor,” Phys. Rev. D 78, 122002. charged ions: Nuclear-spin-zero f12 -shell ions [Phys. Rev. A
Dine, M, and W. Fischler (1983), “The not-so-harmless ax- 86, 054501 (2012)],” Phys. Rev. A 87, 029906.
ion,” Phys. Lett. B 120, 137. Dzuba, V A, and V. V. Flambaum (2009a), “Atomic calcula-
Dine, M, W. Fischler, and M. Srednicki (1981), “A simple tions and search for variation of the fine-structure constant
solution to the strong CP problem with a harmless axion,” in quasar absorption spectra,” Can. J. Phys. 87, 15.
Phys. Lett. 104B, 199. Dzuba, V A, and V. V. Flambaum (2009b), “Calculation
Dine, Michael, and Alexander Kusenko (2003), “Origin of the of the (T,P)-odd electric dipole moment of thallium and
matter-antimatter asymmetry,” Rev. Mod. Phys. 76, 1. cesium,” Phys. Rev. A 80, 062509.
Dinh, T H, A. Dunning, V. A. Dzuba, and V. V. Flambaum Dzuba, V A, V. V. Flambaum, and J. S. M. Ginges (2001a),
(2009), “Sensitivity of hyperfine structure to nuclear radius “Calculations of parity-nonconserving s − d amplitudes in
and quark mass variation,” Phys. Rev. A 79, 054102. Cs, Fr, Ba+ , and Ra+ ,” Phys. Rev. A 63 (6), 62101.
DiSciacca, J, M. Marshall, K. Marable, G. Gabrielse, S. Et- Dzuba, V A, V V Flambaum, J S M Ginges, and M G Kozlov
tenauer, E. Tardiff, R. Kalra, D. W. Fitzakerley, M. C. (2002), “Electric dipole moments of Hg, Xe, Rn, Ra, Pu,
George, E. A. Hessels, et al. (2013), “One-particle mea- and TlF induced by the nuclear Schiff moment and limits
surement of the antiproton magnetic moment,” Phys. Rev. on time-reversal violating interactions,” Phys. Rev. A 66,
Lett. 110, 130801. 12111.
Dobrescu, Bogdan A (2005), “Massless gauge bosons other Dzuba, V A, V. V. Flambaum, and Hidetoshi Katori (2015a),
than the photon,” Phys. Rev. Lett. 94, 151802. “Optical clock sensitive to variation of the fine structure
Dobrescu, Bogdan A, and Irina Mocioiu (2006), “Spin- constant based on the Ho14+ ion,” Phys. Rev. A 91, 022119.
dependent macroscopic forces from new particle exchange,” Dzuba, V A, V. V. Flambaum, and I. B. Khriplovich (1986),
J. High Energy Phys. 11, 005. “Enhancement of P- and T-nonconserving effects in rare-
Dolgov, A D, and V. A. Novikov (2012), “CPT, Lorentz earth atoms,” Z. Phys. D 1, 243.
invariance, mass differences, and charge non-conservation,” Dzuba, V A, V. V. Flambaum, M. S. Safronova, S. G. Porsev,
JETP Letters 95, 594. T. Pruttivarasin, M. A. Hohensee, and H. Häffner (2016),
Dragoset, R A, A. Musgrove, C. W. Clark, W. C. Martin, and “Strongly enhanced effects of Lorentz symmetry violation
K. Olsen (2017), “Periodic table of the elements,” NIST in entangled Yb+ ions,” Nature Physics 12, 465–468.
Special Publication 699 . Dzuba, V A, V. V. Flambaum, and O. P. Sushkov (1989),
“Summation of the high orders of perturbation theory for
95

the parity nonconserving E1-amplitude of the 6S-7S tran- Fayet, P (2004), “Light spin-1/2 or spin-0 dark matter parti-
sition in the caesium atom,” Phys. Lett. A 141, 147. cles,” Phys. Rev. D 70, 23514.
Dzuba, V A, V. V. Flambaum, and O. P. Sushkov (1995), Fayet, Pierre (2007), “U-boson production in e+ e− annihila-
“Calculation of energy levels, E1 transition amplitudes, and tions, ψ and γ decays, and light dark matter,” Phys. Rev.
parity violation in francium,” Phys. Rev. A 51, 3454. D 75, 115017.
Dzuba, V A, V. V. Flambaum, and J. K. Webb (1999a), “Cal- Fedorov, V V, I. A. Kuznetsov, and V. V. Voronin (2013),
culations of the relativistic effects in many-electron atoms “A search for nEDM and new constraints on short-range
and space-time variation of fundamental constants,” Phys. “pseudo-magnetic” interaction using neutron optics of non-
Rev. A 59, 230. centrosymmetric crystal,” Nucl. Instrum. Methods Phys.
Dzuba, V A, V. V. Flambaum, and J. K. Webb (1999b), Research Sec. B 309, 237.
“Space-time variation of physical constants and relativistic Feng, J L (2010), “Dark matter candidates from particle
corrections in atoms,” Phys. Rev. Lett. 82, 888. physics and methods of detection,” Annu. Rev. Astron. As-
Dzuba, V A, C. Harabati, W. R Johnson, and M. S. Safronova trophys. 48, 495.
(2001b), “Breit correction to the parity-nonconservation Feng, Jonathan L (2013), “Naturalness and the status of su-
amplitude in cesium,” Phys. Rev. A 63, 44103. persymmetry,” Annu. Rev. Nucl. Part. Sci. 63, 351.
Dzuba, V A, M. S. Safronova, U. I. Safronova, and V. V. Ferrell, S J, A. Cingöz, A. Lapierre, A.-T. Nguyen, N. Leefer,
Flambaum (2015b), “Actinide ions for testing the spatial D. Budker, V. V. Flambaum, S. K. Lamoreaux, and
α−variation hypothesis,” Phys. Rev. A 92, 060502. J. R. Torgerson (2007), “Investigation of the gravitational-
Eckel, S, P. Hamilton, E. Kirilov, H. W. Smith, and D. De- potential dependence of the fine-structure constant using
Mille (2013), “Search for the electron electric dipole mo- atomic dysprosium,” Phys. Rev. A 76, 062104.
ment using Ω-doublet levels in PbO,” Phys. Rev. A 87, Ficek, Filip, Derek F. Jackson Kimball, Mikhail G. Ko-
052130. zlov, Nathan Leefer, Szymon Pustelny, and Dmitry Bud-
Edwards, N H, S. J. Phipp, P. E. G. Baird, and S. Nakayama ker (2017), “Constraints on exotic spin-dependent inter-
(1995), “Precise measurement of parity nonconserving op- actions between electrons from helium fine-structure spec-
tical rotation in atomic thallium,” Phys. Rev. Lett. 74, troscopy,” Phys. Rev. A 95, 032505.
2654. Firestone, R B, and V. S. Shirley, Eds. (1998), Table of Iso-
Eides, M I, and V. A. Shelyuto (2017), “One More Hard topes (Wiley-VCH).
Three-Loop Correction to Parapositronium Energy Lev- Fischbach, E, and D. E. Krause (1999), “Constraints on light
els,” ArXiv e-prints arXiv:1705.09166 [hep-ph]. pseudoscalars implied by tests of the gravitational inverse-
Eides, Michael I, Howard Grotch, and Valery A Shelyuto square law,” Phys. Rev. Lett. 83, 3593.
(2001), “Theory of light hydrogenlike atoms,” Phys. Rep. Fischbach, E, D. Sudarsky, A. Szafer, C. Talmadge, and S. H.
342, 63. Aronson (1986), “Reanalysis of the Eötvös experiment,”
Eills, J, J. W. Blanchard, L. Bougas, M. G. Kozlov, Phys. Rev. Lett. 56, 3.
A. Pines, and D. Budker (2017), “Measuring molec- Fischbach, E, and C. L. Talmadge (2012), The Search for
ular parity nonconservation using nuclear magnetic res- non-Newtonian Gravity (Springer, New York).
onance spectroscopy,” ArXiv e-prints arXiv:1707.01759 Fischler, Willy, Sonia Paban, and Scott Thomas (1992),
[physics.chem-ph]. “Bounds on microscopic physics from P and T violation
Einstein, A (1916), “The basis of the general theory of rela- in atoms and molecules,” Phys. Lett. B 289, 373.
tivity,” Annalen der Physik 354, 769. Fitzakerley, D W, M. C. George, E. A. Hessels, T. D. G.
Eisele, C, A. Y. Nevsky, and S. Schiller (2009), “Laboratory Skinner, C. H. Storry, M. Weel, G. Gabrielse, C. D. Ham-
test of the isotropy of light propagation at the 10−17 level,” ley, N. Jones, K. Marable, et al. (ATRAP Collaboration)
Phys. Rev. Lett. 103, 090401. (2016), “Electron-cooled accumulation of 4 × 109 positrons
Elliott, S R, B. H. LaRoque, V. M. Gehman, M. F. Kidd, and for production and storage of antihydrogen atoms,” J.
M. Chen (2012), “An improved limit on Pauli-exclusion- Phys. B 49, 064001.
principle forbidden atomic transitions,” Foundations of Fixler, J B, G. T. Foster, J. M. McGuirk, and M. A. Ka-
Physics 42, 1015. sevich (2007), “Atom interferometer measurement of the
Engel, J, D. Seckel, and A. C. Hayes (1990), “Emission Newtonian constant of gravity,” Science 315, 74.
and detectability of hadronic axions from SN1987A,” Phys. Flambaum, V, S. Lambert, and M. Pospelov (2009), “Scalar-
Rev. Lett. 65, 960. tensor theories with pseudoscalar couplings,” Phys. Rev. D
Engel, J, and P. Vogel (1989), “Spin-dependent cross sections 80, 105021.
of weakly interacting massive particles on nuclei,” Phys. Flambaum, V V (2006), “Enhanced effect of temporal varia-
Rev. D 40, 3132. tion of the fine-structure constant in diatomic molecules,”
Engel, Jonathan, Michael J. Ramsey-Musolf, and U. van Phys. Rev. A 73, 034101.
Kolck (2013), “Electric dipole moments of nucleons, nu- Flambaum, V V (2016), “Enhancing the Effect of Lorentz
clei, and atoms: The Standard Model and beyond,” Prog. Invariance and Einstein’s Equivalence Principle Violation
Part. Nucl. Phys. 71, 21. in Nuclei and Atoms,” Phys. Rev. Lett. 117 (7), 072501.
English, D, V. V. Yashchuk, and D. Budker (2010), “Spectro- Flambaum, V V, D. DeMille, and M. G. Kozlov (2014),
scopic test of Bose-Einstein statistics for photons,” Phys. “Time-reversal symmetry violation in molecules induced by
Rev. Lett. 104, 253604. nuclear magnetic quadrupole moments,” Phys. Rev. Lett.
Epp, S W (2013), “Comment on “Testing three-body quan- 113, 103003.
tum electrodynamics with trapped Ti20+ ions: Evidence Flambaum, V V, V. A. Dzuba, and C. Harabati (2017), “Ef-
for a Z-dependent divergence between experiment and cal- fect of nuclear quadrupole moment on parity nonconserva-
culation”,” Phys. Rev. Lett. 110, 159301. tion in atoms,” arXiv:1704.08809.
96

Flambaum, V V, and J. S. M. Ginges (2002), “Nuclear Schiff Frieman, J A, M. S. Turner, and D. Huterer (2008), “Dark
moment and time-invariance violation in atoms,” Phys. energy and the accelerating universe,” Annu. Rev. Astron.
Rev. A 65, 032113. Astrophys. 46, 385.
Flambaum, V V, A. A. Gribakina, G. F. Gribakin, and Frolov, A (2005), “Positron annihilation in the positronium
M. G. Kozlov (1994), “Structure of compound states in the negative ion ps,” Phys. Lett. A 342, 430.
chaotic spectrum of the Ce atom: Localization properties, Frolov, A M, and D. M. Wardlaw (2008), “Stability of the
matrix elements, and enhancement of weak perturbations,” five-body bi-positronium ion Ps2 e− ,” Phys. Lett. A 372,
Phys. Rev. A 50, 267. 6721.
Flambaum, V V, and I. B. Khriplovich (1980), “P-odd nu- Fujii, Y (1971), “Dilaton and possible non-Newtonian grav-
clear forces - a source of parity violation in atoms,” Zh. ity,” Nature (London) 234, 5.
Eksp. Teor. Fiz. 79, 1656 [Sov. Phys. JETP 52, 835 (1980)]. Gabrielse, G, S. F. Hoogerheide, J. Dorr, and E. Novit-
Flambaum, V V, and I. B. Khriplovich (1985a), “New bounds ski (2014), “Precise Matter and Antimatter Tests of the
on the electric dipole moment of the electron and on T-odd Standard Model with e− , e+ , p, p and H,,” in Funda-
electron-nucleon coupling,” Zh. Eksp. Theor. Fiz. 89, 1505 mental Physics in Particle Traps, Springer Tracts in Mod-
[JETP 62, 872 (1985)]. ern Physics, Vol. 256, edited by W. Quint and M. Vogel
Flambaum, V V, and I. B. Khriplovich (1985b), “On the (Springer, Berlin - Heidelberg) p. 1.
enhancement of parity nonconserving effects in diatomic Gabrielse, G, A. Khabbaz, D. S. Hall, C. Heimann, H. Kali-
molecules,” Phys. Lett. A 110, 121. nowsky, and W. Jhe (1999), “Precision mass spectroscopy
Flambaum, V V, I. B. Khriplovich, and O. P. Sushkov (1984), of the antiproton and proton using simultaneously trapped
“Nuclear anapole moments,” Phys. Lett. 146B, 367. particles,” Phys. Rev. Lett. 82, 3198.
Flambaum, V V, and M. G. Kozlov (2007), “Enhanced sen- Gacesa, M, and R. Côté (2014), “Photoassociation of ultra-
sitivity to the time variation of the fine-structure constant cold molecules near a feshbach resonance as a probe of the
and mp /me in diatomic molecules,” Phys. Rev. Lett. 99, electron-proton mass ratio variation,” J. Mol. Spectrosc.
150801. 300, 124.
Flambaum, V V, D. B. Leinweber, A. W. Thomas, and R. D. Gambini, R, and J. Pullin (2003), “Discrete quantum grav-
Young (2004), “Limits on variations of the quark masses, ity,” Int. J. Mod. Phys. D 12, 1775.
qcd scale, and fine structure constant,” Phys. Rev. D 69, Garbacz, P (2014), “Spinspin coupling in the hd molecule
115006. determined from 1 H and 2 H NMR experiments in the gas-
Flambaum, V V, and M. V. Romalis (2017), “Limits on phase,” Chem. Phys. 443, 1.
lorentz invariance violation from coulomb interactions in Gato-Rivera, B (2015), “Constraining extra space dimen-
nuclei and atoms,” Phys. Rev. Lett. 118, 142501. sions using precision molecular spectroscopy,” in Journal
Flambaum, V V, Y. V. Stadnik, M. G. Kozlov, and of Physics Conference Series, Vol. 626, p. 012052.
A. N. Petrov (2013), “Enhanced effects of temporal varia- Gavela, M B, P. Hernández, J. Orloff, and O. Pène (1994),
tion of the fundamental constants in 2 Π1/2 -term diatomic “Standard model CP-violation and baryon asymmetry,”
molecules: 207 Pb19 F,” Phys. Rev. A 88, 052124. Mod. Phys. Lett. A 9, 795.
Flambaum, V V, and A. F. Tedesco (2006), “Dependence of Geiger, R, L. Amand, A. Bertoldi, B. Canuel, W. Chaibi,
nuclear magnetic moments on quark masses and limits on C. Danquigny, I. Dutta, B. Fang, S. Gaffet, J. Gillot,
temporal variation of fundamental constants from atomic D. Holleville, A. Landragin, M. Merzougui, I. Riou,
clock experiments,” Phys. Rev. C 73, 055501. D. Savoie, and P. Bouyer (2016), “Matter-wave laser inter-
Fortier, T M, N. Ashby, J. C. Bergquist, M. J. Delaney, S. A. ferometric gravitation antenna (MIGA): New perspectives
Diddams, T. P. Heavner, L. Hollberg, W. M. Itano, S. R. for fundamental physics and geosciences,” E3S Web of Con-
Jefferts, K. Kim, et al. (2007), “Precision atomic spec- ferences 4, 01004.
troscopy for improved limits on variation of the fine struc- Geiger, R, V. Ménoret, G. Stern, N. Zahzam, P. Cheinet,
ture constant and local position invariance,” Phys. Rev. B. Battelier, A. Villing, F. Moron, M. Lours, Y. Bidel,
Lett. 98, 070801. A. Bresson, A. Landragin, and P. Bouyer (2011), “Detect-
Fortson, E N, Y. Pang, and L. Wilets (1990), “Nuclear- ing inertial effects with airborne matter-wave interferome-
structure effects in atomic parity nonconservation,” Phys. try,” Nature Commun. 2, 474.
Rev. Lett. 65, 2857. Gelmini, G, S. Nussinov, and T. Yanagida (1983), “Does
Fortson, N, P. Sandars, and S. Barr (2003), “The search for nature like Nambu-Goldstone bosons?” Nucl. Phys. B 219,
a permanent electric dipole moment,” Phys. Today 56, 33. 31.
Fortson, Norval (1993), “Possibility of measuring parity non- Gelmini, G, and M. Roncadelli (1981), “Left-handed neu-
conservation with a single trapped atomic ion,” Phys. Rev. trino mass scale and spontaneously broken lepton number,”
Lett. 70, 2383. Phys. Lett. 99B, 411.
Fray, S, C. A. Diez, T. W. Hänsch, and M. Weitz (2004), Gemmel, C, W. Heil, S. Karpuk, K. Lenz, Yu. Sobolev,
“Atomic interferometer with amplitude gratings of light K. Tullney, M. Burghoff, W. Kilian, S. Knappe-Grüneberg,
and its applications to atom based tests of the equivalence W. Müller, A. Schnabel, F. Seifert, L. Trahms, and
principle,” Phys. Rev. Lett. 93, 240404. U. Schmidt (2010), “Limit on Lorentz and CPT violation
Freese, Katherine, Mariangela Lisanti, and Christopher Sav- of the bound neutron using a free precession 3 He/129 Xe
age (2013), “Colloquium: Annual modulation of dark mat- comagnetometer,” Phys. Rev. D 82, 111901(R).
ter,” Rev. Mod. Phys. 85, 1561. Georgescu, I M, S. Ashhab, and F. Nori (2014), “Quantum
Freivogel, B (2010), “Anthropic explanation of the dark mat- simulation,” Rev. Mod. Phys. 86, 153.
ter abundance,” Journal of Cosmology and Astroparticle Georgi, Howard (2007), “Unparticle physics,” Phys. Rev.
Physics 2010, 021. Lett. 98, 221601.
97

Geraci, A, and H. Goldman (2015), “Sensing short range accelerometers,” Phys. Rev. D 93, 075029.
forces with a nanosphere matter-wave interferometer,” Graham, P W, David E. Kaplan, and Surjeet Rajen-
Phys. Rev. D 92, 062002. dran (2015b), “Cosmological relaxation of the electroweak
Geraci, A A, S. B. Papp, and J. Kitching (2010), “Short- scale,” Phys. Rev. Lett. 115, 221801.
range force detection using optically cooled levitated mi- Graham, P W, and S. Rajendran (2011), “Axion dark matter
crospheres,” Phys. Rev. Lett. 105, 101101. detection with cold molecules,” Phys. Rev. D 84, 055013.
Geraci, A A, S. J. Smullin, D. M. Weld, J. Chiaverini, Graham, P W, and S. Rajendran (2013), “New observables
and A. Kapitulnik (2008), “Improved constraints on non- for direct detection of axion dark matter,” Phys. Rev. D
Newtonian forces at 10 microns,” Phys. Rev. D 78, 022002. 88, 035023.
Geraci, Andrew A, and Andrei Derevianko (2016), “Sensitiv- Graner, B, Y. Chen, E. G. Lindahl, and B. R. Heckel (2016),
ity of Atom Interferometry to Ultralight Scalar Field Dark “Reduced limit on the permanent electric dipole moment
Matter,” Phys. Rev. Lett. 117, 261301. of 199 Hg199,” Phys. Rev. Lett. 116, 1.
Gharibnejad, H, and A. Derevianko (2015), “Dark forces and Gray, C G, G. Karl, and V. A. Novikov (2010), “Magnetic
atomic electric dipole moments,” Phys. Rev. D 91, 035007. multipolar contact fields: The anapole and related mo-
Gillaspy, J D (2014), “Precision spectroscopy of trapped ments,” Am. J. Phys. 78, 936.
highly charged heavy elements: pushing the limits of theory Greenberg, O W (1991), “Particles with small violations of
and experiment,” Phys. Src. 89, 114004. Fermi or Bose statistics,” Phys. Rev. D 43, 4111.
Ginges, J S M, and V. V. Flambaum (2004), “Violations of Greenberg, O W (2002), “CPT violation implies violation of
fundamental symmetries in atoms and tests of unification lorentz invariance,” Phys. Rev. Lett. 89, 231602.
theories of elementary particles,” Phys. Rep. 397, 63. Greenberg, O W, and Robert C. Hilborn (1999), “Quon
Giudice, Gian Francesco, and Andrea Romanino (2006), statistics for composite systems and a limit on the violation
“Electric dipole moments in split supersymmetry,” Phys. of the Pauli principle for nucleons and quarks,” Phys. Rev.
Lett. B 634, 307. Lett. 83, 4460.
Glauber, R J (2006), “Nobel lecture: One hundred years of Greenberg, O W, and R. N. Mohapatra (1989), “Phenomenol-
light quanta,” Rev. Mod. Phys. 78, 1267. ogy of small violations of Fermi and Bose statistics,” Phys.
Glenday, A G, C. E. Cramer, D. F. Phillips, and R. L. Rev. D 39, 2032.
Walsworth (2008), “Limits on anomalous spin-spin cou- Greiner, Walter (2000), Relativistic Quantum Mechanics
plings between neutrons,” Phys. Rev. Lett. 101, 261801. Wave Equations, 3rd ed. (Springer, Berlin).
Godun, R M, P. B. R. Nisbet-Jones, J. M. Jones, S. A. King, Gresh, D, W. Cairncross, K. Cossel, M. Grau, K. B. Ng,
L. A. M. Johnson, H. S. Margolis, K. Szymaniec, S. N. Y. Zhou, Y. Ni, J. Ye, and E. Cornell (2016), “Progress of
Lea, K. Bongs, and P. Gill (2014), “Frequency ratio of two the JILA electron EDM experiment,” in APS Division of
optical clock transitions in 171 Yb+ and constraints on the Atomic, Molecular and Optical Physics Meeting Abstracts.
time variation of fundamental constants,” Phys. Rev. Lett. Griffith, W C, M. D. Swallows, T. H. Loftus, M. V. Romalis,
113, 210801. B. R. Heckel, and E. N. Fortson (2009), “Improved limit
Goldman, N, G. Juzeliūnas, P. Öhberg, and I. B. Spielman on the permanent electric dipole moment of 199 Hg,” Phys.
(2014), “Light-induced gauge fields for ultracold atoms,” Rev. Lett. 102, 101601.
Rep. Prog. Phys. 77, 126401. Guéna, J, M. Abgrall, D. Rovera, P. Rosenbusch, M. E. To-
Gomes Ferreiraa, F A, P. C. Maltab, L. P. R. Ospedalc, and bar, P. Laurent, A. Clairon, and S. Bize (2012), “Improved
J. A. Helayël-Netod (2015), “Topologically massive spin-1 tests of local position invariance using 87 Rb and 133 Cs foun-
particles and spin-dependent potentials,” Eur. Phys. J. C tains,” Phys. Rev. Lett. 109, 080801.
75, 238. Guigue, M, D. Jullien, A. K. Petukhov, and G. Pignol (2015),
Gomez, E, S Aubin, G D Sprouse, L A Orozco, and “Constraining short-range spin-dependent forces with po-
D P DeMille (2007), “Measurement method for the nu- larized 3 He,” Phys. Rev. D 92, 114001.
clear anapole moment of laser-trapped alkali-metal atoms,” Gumberidze, A, T. Stöhlker, D. Banaś, K. Beckert, P. Beller,
Phys. Rev. A 75, 33418. H. F. Beyer, F. Bosch, S. Hagmann, C. Kozhuharov,
Gomez, E, L A Orozco, and G D Sprouse (2006), “Spec- D. Liesen, F. Nolden, X. Ma, P. H. Mokler, M. Steck,
troscopy with trapped francium: advances and perspectives D. Sierpowski, and S. Tashenov (2005), “Quantum elec-
for weak interaction studies,” Rep. Prog. Phys. 69, 79. trodynamics in strong electric fields: The ground-state
Gorenstein, P, and W. Tucker (2014), “Astronomical signa- Lamb shift in hydrogenlike uranium,” Phys. Rev. Lett. 94,
tures of dark matter,” Advances in High Energy Physics 223001.
2014, 878203. Gundlach, J H (2005), “Laboratory tests of gravity,” New J.
Graham, P W, J. M. Hogan, M. A. Kasevich, and S. Rajen- Phys. 7, 205.
dran (2013), “New method for gravitational wave detection Guzman, J, T. Inaki, and A. Penaflor (2015), “Search for
with atomic sensors,” Phys. Rev. Lett. 110, 171102. violations of Bose-Einstein statistics using ultra-cold Sr
Graham, P W, J. M. Hogan, M. A. Kasevich, and S. Ra- atoms,” in APS Division of Atomic, Molecular and Optical
jendran (2016a), “Resonant mode for gravitational wave Physics Meeting Abstracts.
detectors based on atom interferometry,” Phys. Rev. D 94, Haber, H E, G. L. Kane, and T. Sterling (1979), “The fermion
104022. mass scale and possible effects of Higgs bosons on experi-
Graham, P W, I. G. Irastorza, S. K. Lamoreaux, A. Lindner, mental observables,” Nucl. Phys. B 161, 493.
and K. A. van Bibber (2015a), “Experimental searches for Hall, J L (2006), “Nobel lecture: Defining and measuring
the axion and axion-like particles,” Annu. Rev. Nucl. Part. optical frequencies,” Rev. Mod. Phys. 78, 1279.
Sci. 65, 485. Hall, L J, and D. Smith (1999), “Cosmological constraints
Graham, P W, D. E. Kaplan, J. Mardon, S. Rajendran, and on theories with large extra dimensions,” Phys. Rev. D 60,
W. A. Terrano (2016b), “Dark matter direct detection with 085008.
98

Hamilton, P, M. Jaffe, P. Haslinger, Q. Simmons, H. Müller, Heinzen, D J, and D. J. Wineland (1990), “Quantum-limited
and J. Khoury (2015), “Atom-interferometry constraints on cooling and detection of radio-frequency oscillations by
dark energy,” Science 349, 849. laser-cooled ions,” Phys. Rev. A 42, 2977.
Hammond, R T (1995), “Upper limit on the torsion coupling Henkel, C, K. M. Menten, M. T. Murphy, N. Jethava, V. V.
constant,” Phys. Rev. D 52, 6918. Flambaum, J. A. Braatz, S. Muller, J. Ott, and R. Q. Mao
Hanneke, D, S. Fogwell, and G. Gabrielse (2008), “New mea- (2009), “The density, the cosmic microwave background,
surement of the electron magnetic moment and the fine and the proton-to-electron mass ratio in a cloud at redshift
structure constant,” Phys. Rev. Lett. 100, 120801. 0.9,” Astron. Astrophys. 500, 725.
Hänsch, T W (2006), “Nobel lecture: Passion for precision,” Herrmann, S, H. Dittus, and C. (for the QUANTUS and
Rev. Mod. Phys. 78, 1297. PRIMUS teams) Lämmerzahl (2012), “Testing the equiv-
Happer, W, Y.-Y. Jau, and T. Walker (2010), Optically alence principle with atomic interferometry,” Class. and
Pumped Atoms (Wiley-VCH). Quantum Gravity 29, 184003.
Happer, William (1972), “Optical Pumping,” Rev. Mod. Herrmann, S, A. Senger, K. Möhle, M. Nagel, E. V. Ko-
Phys. 44, 169. valchuk, and A. Peters (2009), “Rotating optical cavity
Harabati, C, VA Dzuba, VV Flambaum, and MA Hohensee experiment testing lorentz invariance at the 10−17 level,”
(2015), “Effects of lorentz-symmetry violation on the spec- Phys. Rev. D 80, 105011.
tra of rare-earth ions in a crystal field,” Phys. Rev. A 92, Hill, R J (2017), “Review of experimental and theoretical sta-
040101. tus of the proton radius puzzle,” in European Physical Jour-
Hari Dass, N D (1976), “Experimental tests for some quantum nal Web of Conferences, Vol. 137, p. 01023.
effects in gravitation,” Phys. Rev. Lett. 36, 393. Hoedl, S A, F. Fleischer, E. G. Adelberger, and B. R. Heckel
Haroche, S (2013), “Nobel lecture: Controlling photons in (2011), “Improved constraints on an axion-mediated force,”
a box and exploring the quantum to classical boundary,” Phys. Rev. Lett. 106, 041801.
Rev. Mod. Phys. 85, 1083. Hoekstra, S (2016), Private communication.
Harrick, N J, R. G. Barnes, P. J. Bray, and N. F. Ramsey Hogan, C J (2000), “Why the universe is just so,” Rev. Mod.
(1953), “Nuclear radiofrequency spectra of D2 and H2 in Phys. 72, 1149.
intermediate and strong magnetic fields,” Phys. Rev. 90, Hogan, J M, D. M. S. Johnson, S. Dickerson, T. Kovachy,
260. A. Sugarbaker, S.-W. Chiow, P. W. Graham, M. A. Kase-
Hartwig, J, S. Abend, C. Schubert, D. Schlippert, H. Ahlers, vich, B. Saif, S. Rajendran, P. Bouyer, B. D. Seery, L. Fein-
K. Posso-Trujillo, N. Gaaloul, W. Ertmer, and E. M. Rasel berg, and R. Keski-Kuha (2011), “An atomic gravita-
(2015), “Testing the universality of free fall with rubidium tional wave interferometric sensor in low earth orbit (AGIS-
and ytterbium in a very large baseline atom interferome- LEO),” Gen. Relativ. Gravit. 43, 1953.
ter,” New J. Phys. 17, 035011. Hogan, J M, and M. A. Kasevich (2016), “Atom-
Haxton, W C, and K. Y. Lee (1991), “Red-giant evolution, interferometric gravitational-wave detection using hetero-
metallicity, and new bounds on hadronic axions,” Phys. dyne laser links,” Phys. Rev. A 94, 033632.
Rev. Lett. 66, 2557. Hohensee, M A, S. Chu, A. Peters, and H. Müller (2011),
Haxton, W C, and C. E. Wieman (2001), “Atomic parity non- “Equivalence principle and gravitational redshift,” Phys.
conservation and nuclear anapole moments,” Annu. Rev. Rev. Lett. 106, 151102.
Nucl. Part. Sci. 51, 261. Hohensee, M A, B. Estey, P. Hamilton, A. Zeilinger, and
Haxton, Wick C, and Barry R. Holstein (2013), “Hadronic H. Müller (2012), “Force-free gravitational redshift: pro-
parity violation,” Prog. Part. Nucl. Phys. 71, 185. posed gravitational Aharonov-Bohm experiment,” Phys.
Hayes, A C, J. L. Friar, and P. Möller (2008), “Splitting sen- Rev. Lett. 108, 230404.
sitivity of the ground and 7.6 eV isomeric states of 229 Th,” Hohensee, M A, N. Leefer, D. Budker, C. Harabati, V. A.
Phys. Rev. C 78, 024311. Dzuba, and V. V. Flambaum (2013a), “Limits on vio-
Heavner, T P, E. A. Donley, F. Levi, G. Costanzo, T. E. lations of Lorentz symmetry and the Einstein equivalence
Parker, J. H. Shirley, N. Ashby, S. Barlow, and S. R. principle using radio-frequency spectroscopy of atomic dys-
Jefferts (2014), “First accuracy evaluation of NIST-F2,” prosium,” Phys. Rev. Lett. 111, 050401.
Metrologia 51, 174. Hohensee, M A, H. Müller, and R. B. Wiringa (2013b),
Heckel, B (2016), “http://online.kitp.ucsb.edu/online/nuclear- “Equivalence principle and bound kinetic energy,” Phys.
c16/heckel/options.html,” . Rev. Lett. 111, 151102.
Heckel, B R, E. G. Adelberger, C. E. Cramer, T. S. Cook, Hohensee, Michael A, Paul L. Stanwix, Michael E. Tobar,
S. Schlamminger, and U. Schmidt (2008), “Preferred-frame Stephen R. Parker, David F. Phillips, and Ronald L.
and CP-violation tests with polarized electrons,” Phys. Walsworth (2010), “Improved constraints on isotropic shift
Rev. D 78, 092006. and anisotropies of the speed of light using rotating cryo-
Hedges, Morgan, Marc Smiciklas, and Michael Romalis genic sapphire oscillators,” Phys. Rev. D 82, 076001.
(2015), “South pole lorentz invariance test,” in APS April Holdom, Bob (1986), “Two U(1)’s and  charge shifts,” Phys.
Meeting Abstracts, Vol. 1, p. 6003. Lett. B 166, 196.
Hees, A, J. Guéna, M. Abgrall, S. Bize, and P. Wolf (2016), Hoogeveen, F (1990), “The standard model prediction for the
“Searching for an oscillating massive scalar field as a dark electric dipole moment of the electron,” Nucl. Phys. B 341,
matter candidate using atomic hyperfine frequency com- 322.
parisons,” Phys. Rev. Lett. 117, 061301. Hopkinson, D A, and P. E. G. Baird (2002), “An interfero-
Hehl, Friederich W, Paul von der Heyde, G. David Kerlick, metric test of time reversal invariance in atoms,” J. Phys.
and James M. Nester (1976), “General relativity with spin B 35, 1307.
and torsion: Foundations and prospects,” Rev. Mod. Phys. Hori, M, H. Aghai-Khozani, A. Sótér, D. Barna, A. Dax,
48, 393. R. Hayano, T. Kobayashi, Y. Murakami, K. Todoroki,
99

H. Yamada, D. Horváth, and L. Venturelli (2016), “Buffer- and C. Monroe (2013), “Emergence and frustration of mag-
gas cooling of antiprotonic helium to 1.5 to 1.7 K, and netism with variable-range interactions in a quantum sim-
antiproton-to-electron mass ratio,” Science 354, 610. ulator,” Science 340, 583.
Hori, M, A. Sótér, D. Barna, A. Dax, R. Hayano, S. Friedre- Ivanov, A N, and W. M. Snow (2017), “Parity-even and time-
ich, B. Juhász, T. Pask, E. Widmann, D. Horváth, L. Ven- reversal-odd neutron optical potential in spinning matter
turelli, and N. Zurlo (2011), “Two-photon laser spec- induced by gravitational torsion,” Phys. Lett. B 764, 186.
troscopy of antiprotonic helium and the antiproton-to- Iwazaki, A (2015), “Axion stars and fast radio bursts,” Phys.
electron mass ratio,” Nature (London) 475, 484. Rev. D 91, 023008.
Hori, M, A. Sótér, and V. I. Korobov (2014), “Proposed Jackson, J D (1999), Classical Electrodynamics, 3rd ed. (John
method for laser spectroscopy of pionic helium atoms to de- Willey & Sons, New York).
termine the charged-pion mass,” Phys. Rev. A 89, 042515. Jackson Kimball, D F (2015), “Nuclear spin content and con-
Hornberger, K, S. Gerlich, P. Haslinger, S. Nimmrichter, and straints on exotic spin-dependent couplings,” New J. Phys.
M. Arndt (2012), “Colloquium: Quantum interference of 17, 073008.
clusters and molecules,” Rev. Mod. Phys. 84, 157. Jackson Kimball, D F, A. Boyd, and D. Budker (2010), “Con-
Huber, Stephan J, Maxim Pospelov, and Adam Ritz (2007), straints on anomalous spin-spin interactions from spin-
“Electric dipole moment constraints on minimal elec- exchange collisions,” Phys. Rev. A 82, 062714.
troweak baryogenesis,” Phys. Rev. D 75, 36006. Jackson Kimball, D F, J. Dudley, Y. Li, D. Patel,
Hudson, E R, H. J. Lewandowski, B. C. Sawyer, and J. Ye and J. Valdez (2017), “Constraints on long-range spin-
(2006), “Cold molecule spectroscopy for constraining the gravity and monopole-dipole couplings of the proton,”
evolution of the fine structure constant,” Phys. Rev. Lett. arXiv:1707.00745.
96, 143004. Jackson Kimball, D F, J. Dudley, Y. Li, S. Thulasi,
Hudson, J J, D. M. Kara, I. J. Smallman, B. E. Sauer, M. R. S. Pustelny, D. Budker, and M. Zolotorev (2016a), “Mag-
Tarbutt, and E. A. Hinds (2011), “Improved measurement netic shielding and exotic spin-dependent interactions,”
of the shape of the electron,” Nature (London) 473, 493. Phys. Rev. D 94, 082005.
Hughes, R J, and M. H. Holzscheiter (1991), “Constraints on Jackson Kimball, D F, I. Lacey, J. Valdez, J. Swiatlowski,
the gravitational properties of antiprotons and positrons C. Rios, R. Peregrina-Ramirez, C. Montcrieffe, J. Kremer,
from cyclotron-frequency measurements,” Phys. Rev. Lett. J. Dudley, and C. Sanchez (2013a), “A dual-isotope rubid-
66, 854. ium comagnetometer to search for anomalous long-range
Hughes, V W, H. G. Robinson, and V. Beltran-Lopez (1960), spin-mass (spin-gravity) couplings of the proton,” Annalen
“Upper limit for the anisotropy of inertial mass from nu- der Physik 525, 514.
clear resonance experiments,” Phys. Rev. Lett. 4, 342. Jackson Kimball, D F, S. K. Lamoreaux, and T. E. Chupp
Huntemann, N, B. Lipphardt, C. Tamm, V. Gerginov, (2013b), “Tests of fundamental physics with optical magne-
S. Weyers, and E. Peik (2014), “Improved limit on a tem- tometers” in Optical Magnetometry, edited by D. Budker
poral variation of mp /me from comparisons of Yb+ and Cs and D. F. Jackson Kimball (Cambridge University, Cam-
atomic clocks,” Phys. Rev. Lett. 113, 210802. bridge, UK).
Huntemann, N, M. Okhapkin, B. Lipphardt, S. Weyers, Jackson Kimball, D F, A. O. Sushkov, and D. Budker
C. Tamm, and E. Peik (2012), “High-accuracy optical clock (2016b), “Precessing ferromagnetic needle magnetometer,”
based on the octupole transition in 171 Yb+ ,” Phys. Rev. Phys. Rev. Lett. 116, 190801.
Lett. 108, 090801. Jaeckel, J, and J. Redondo (2013), “Resonant to broadband
Huntemann, N, C. Sanner, B. Lipphardt, C. Tamm, and searches for cold dark matter consisting of weakly interact-
E. Peik (2016), “Single-ion atomic clock with 3 × 10−18 ing slim particles,” Phys. Rev. D 88, 115002.
systematic uncertainty,” Phys. Rev. Lett. 116, 063001. Jaffe, M, P. Haslinger, V. Xu, P. Hamilton, A. Upadhye,
Hunter, L, J. Gordon, S. Peck, D. Ang, and Lin J.-F. (2013), B. Elder, J. Khoury, and H. Müller (2016), “Testing sub-
“Using the Earth as a polarized electron source to search gravitational forces on atoms from a miniature, in-vacuum
for long-range spin-spin interactions,” Science 339, 928. source mass,” arXiv:1612.05171.
Hunter, L R, and D. G. Ang (2014), “Using geoelectrons to Jain, P, and S. Mandal (2006), “Evading the astrophysi-
search for velocity-dependent spin-spin interactions,” Phys. cal limits on light pseudoscalars,” International Journal of
Rev. Lett. 112, 091803. Modern Physics D 15, 2095.
Hylleraas, Egil A, and Aadne Ore (1947), “Binding energy Janka, H T, W. Keil, G. Raffelt, and D. Seckel (1996), “Nu-
of the positronium molecule,” Phys. Rev. 71, 493. cleon spin fluctuations and the supernova emission of neu-
Inoue, T, S. Ando, T. Aoki, H. Arikawa, S. Ezure, K. Harada, trinos and axions,” Phys. Rev. Lett. 76, 2621.
T. Hayamizu, T. Ishikawa, M. Itoh, K. Kato, et al. (2015), Jansen, P, H. L. Bethlem, and W. Ubachs (2014), “Perspec-
“Experimental search for the electron electric dipole mo- tive: Tipping the scales: Search for drifting constants from
ment with laser cooled francium atoms,” Hyperfine Inter- molecular spectra,” J. Chem. Phys. 140, 010901.
act. 231, 157. Jansen, P, I. Kleiner, C. Meng, R. M. Lees, M. H. M. Janssen,
Ishida, A (2015), “New precise measurement of the hyperfine W. Ubachs, and H. L. Bethlem (2013), “Prospects for high-
splitting of positronium,” J. Phys. Chem. Ref. Data 44, resolution microwave spectroscopy of methanol in a Stark-
031212. deflected molecular beam,” Mol. Phys. 111, 1923.
Ishida, A, T. Namba, S. Asai, T. Kobayashi, H. Saito, Jeet, J, C. Schneider, S. T. Sullivan, W. G. Rellergert,
M. Yoshida, K. Tanaka, and A. Yamamoto (2014), “New S. Mirzadeh, A. Cassanho, H. P. Jenssen, E. V. Tkalya,
precision measurement of hyperfine splitting of positron- and E. R. Hudson (2015), “Results of a direct search us-
ium,” Phys. Lett. B 734, 338. ing synchrotron radiation for the low-energy 229 Th nuclear
Islam, R, C. Senko, W. C. Campbell, S. Korenblit, J. Smith, isomeric transition,” Phys. Rev. Lett. 114, 253001.
A. Lee, E. E. Edwards, C.-C. J. Wang, J. K. Freericks,
100

Jenke, T, G. Cronenberg, J. Burgdörfer, L. A. Chizhova, Karshenboim, S G (2010a), “Constraints on a long-range


P. Geltenbort, A. N. Ivanov, T. Lauer, T. Lins, S. Rot- spin-dependent interaction from precision atomic physics,”
ter, H. Saul, U. Schmidt, and H. Abele (2014), “Gravity Phys. Rev. D 82, 113013.
resonance spectroscopy constrains dark energy and dark Karshenboim, S G (2010b), “Constraints on a long-range spin-
matter scenarios,” Phys. Rev. Lett. 112, 151105. independent interaction from precision atomic physics,”
Jentschura, U D (2011), “Lamb shift in muonic hydrogenII. Phys. Rev. D 82, 073003.
Analysis of the discrepancy of theory and experiment,” An- Karshenboim, S G (2010c), “Precision physics of simple atoms
nals of Physics 326, 516. and constraints on a light boson with ultraweak coupling,”
Johnson, W, M. Safronova, and U. Safronova (2003), “Com- Phys. Rev. Lett. 104, 220406.
bined effect of coherent Z exchange and the hyperfine in- Karshenboim, S G (2013), “Recent progress in determination
teraction in the atomic parity-nonconserving interaction,” of fundamental constants and fundamental physics at low
Phys. Rev. A 67, 062106. energies,” Ann. Phys. 525, 472.
Johnson, W R, I. Bednyakov, and G. Soff (2001), “Vacuum- Karshenboim, Savely G (2005), “Precision physics of simple
polarization corrections to the parity-nonconserving 6s − atoms: QED tests, nuclear structure and fundamental con-
7s transition amplitude in 133 Cs,” Phys. Rev. Lett. 87, stants,” Phys. Rep. 422, 1, 0509010.
233001. Karshenboim, Savely G (2016), “Positronium, antihydrogen,
Johnson, Walter R (2007), Atomic Structure Theory - Lec- light, and the equivalence principle,” J. Phys. B 49, 144001.
tures on Atomic Physics (Springer-Verlag, Berlin). Kawall, D (2011), “Searching for the electron EDM in a stor-
Jones, A C L, T. H. Hisakado, H. J. Goldman, H. W. K. Tom, age ring,” J. Phys. Conf. Ser. 295, 12031.
A. P. Mills, and D. B. Cassidy (2014), “Doppler-corrected Keesom, W H (1942), Helium (Elsevier).
balmer spectroscopy of rydberg positronium,” Phys. Rev. Kellerbauer, A, C. Canali, A. Fischer, and U. Warring (2009),
A 90, 012503. “Ultracold antiprotons by indirect laser cooling,” Hyperfine
Jordan, E, G. Cerchiari, S. Fritzsche, and A. Kellerbauer Interact. 194, 77.
(2015), “High-resolution spectroscopy on the laser-cooling Kellerbauer, Alban (2015), “Why antimatter matters,” Euro-
candidate La− ,” Phys. Rev. Lett. 115, 113001. pean Review 23, 45.
Joyce, A, B. Jain, J. Khoury, and M. Trodden (2015), “Be- Ketterle, W (2002), “Nobel lecture: When atoms behave as
yond the cosmological standard model,” Phys. Rep. 568, waves: Bose-Einstein condensation and the atom laser,”
1. Rev. Mod. Phys. 74, 1131.
Kahn, Y, B. R. Safdi, and J. Thaler (2016), “Broadband Khoury, J (2013), “Chameleon field theories,” Class. Quan-
and resonant approaches to axion dark matter detection,” tum Gravity 30, 214004.
Phys. Rev. Lett. 117, 141801. Khoury, J, and A. Weltman (2004a), “Chameleon cosmol-
Kajita, M, G. Gopakumar, M. Abe, and M. Hada (2014), ogy,” Phys. Rev. D 69, 044026.
“Characterizing of variation in the proton-to-electron mass Khoury, J, and A. Weltman (2004b), “Chameleon fields:
ratio via precise measurements of molecular vibrational Awaiting surprises for tests of gravity in space,” Phys. Rev.
transition frequencies,” J. Mol. Spectrosc. 300, 99. Lett. 93, 171104.
Kakhidze, A I, and I. V. Kolokolov (1991), “Antiferromag- Khriplovich, I B (1991), Parity Nonconservation in Atomic
netic axion detector,” Sov. Phys. JETP 72, 598. Phenomena (Gordon and Breach, Philadelphia).
Kamiya, Y, K. Itagaki, M. Tani, G. N. Kim, and S. Ko- Khriplovich, I B, and S. K. Lamoureaux (1997), CP Violation
mamiya (2015), “Constraints on new gravitylike forces in Without Strangeness (Springer, Berlin).
the nanometer range,” Phys. Rev. Lett. 114, 161101. Khriplovich, I B, and A R Zhitnitsky (1982), “What is
Kane, Gordon L (2002), “TASI lectures: weak scale super- the value of the neutron electric dipole moment in the
symmetry - a top-motivated-bottom-up approach,” hep- Kobayashi-Maskawa model?” Phys. Lett. B 109, 490.
ph/0202185. Kim, J (1979), “Weak-interaction singlet and strong CP in-
Kanekar, N (2011), “Constraining changes in the proton- variance,” Phys. Rev. Lett. 43, 103.
electron mass ratio with inversion and rotational lines,” Kim, J E, and G. Carosi (2010), “Axions and the strong CP
Astrophys. J. Lett. 728, L12. problem,” Rev. Mod. Phys. 82, 557.
Kanekar, N, W. Ubachs, K. M. Menten, J. Bagdonaite, Kim, W-J, A. O. Sushkov, D. A. R. Dalvit, and S. K. Lamore-
A. Brunthaler, C. Henkel, S. Muller, H. L. Bethlem, and aux (2010), “Surface contact potential patches and Casimir
M. Daprà (2015), “Constraints on changes in the proton- force measurements,” Phys. Rev. A 81, 022505.
electron mass ratio using methanol lines,” Mon. Not. R. King, J A, J. K. Webb, M. T. Murphy, V. V. Flambaum,
Astron. Soc. 448, L104. R. F. Carswell, M. B. Bainbridge, M. R. Wilczynska, and
Kapner, D J, T. S. Cook, E. G. Adelberger, J. H. Gundlach, F. E. Koch (2012), “Spatial variation in the fine-structure
B. R. Heckel, C. D. Hoyle, and H. E. Swanson (2007), constant - new results from VLT/UVES,” Mon. Not. R.
“Tests of the gravitational inverse-square law below the Astron. Soc. 422, 3370.
dark-energy length scale,” Phys. Rev. Lett. 98, 021101. Klimchitskaya, G L, and V. M. Mostepanenko (2015), “Im-
Kara, D M, I. J. Smallman, J. J. Hudson, B. E. Sauer, proved constraints on the coupling constants of axion-like
M. R. Tarbutt, and E. A. Hinds (2012), “Measurement of particles to nucleons from recent Casimir-less experiment,”
the electron’s electric dipole moment using YbF molecules: Eur. Phys. J. C 75, 164.
methods and data analysis,” New J. Phys. 14, 103051. Klinkenberg, P F A (1952), “Tables of nuclear shell struc-
Karr, J-P (2014), “H2 + and HD: Candidates for a molecular ture,” Rev. Mod. Phys. 24, 63.
clock,” J. Mol. Spectrosc. 300, 37. Kobzarev, I Yu, and L. B. Okun (1962), “Gravitational inter-
Karshenboim, S G (2005), “Precision physics of simple atoms: action of fermions,” 43, 1904 [Sov. Phys. JETP 16, 1343
QED tests, nuclear structure and fundamental constants,” (1963)].
Physics Reports 422, 1–63, hep-ph/0509010.
101

Kolb, E W, and I. I. Tkachev (1993), “Axion miniclusters Kozlov, M G, and S. A. Levshakov (2013), “Microwave and
and Bose stars,” Phys. Rev. Lett. 71, 3051. submillimeter molecular transitions and their dependence
Kolkowitz, S, I. Pikovski, N. Langellier, M. D. Lukin, R. L. on fundamental constants,” Annalen der Physik 525, 452.
Walsworth, and J. Ye (2016), “Gravitational wave detec- Kozlov, M G, and S. G. Porsev (1989), “The possibility to
tion with optical lattice atomic clocks,” Phys. Rev. D 94, study the break of time-reversal invariance in atoms,” Phys.
124043. Lett. A 142, 233.
Komatsu, E, K. M. Smith, J. Dunkley, C. L. Bennett, B. Gold, Kozlov, M G, S. G Porsev, and I. I Tupitsyn (2001), “High
G. Hinshaw, N. Jarosik, D. Larson, M. R. Nolta, L. Page, accuracy calculation of 6S → 7S parity nonconserving am-
et al. (2011), “Seven-year Wilkinson microwave anisotropy plitude in Cs,” Phys. Rev. Lett. 86 (15), 3260.
probe,” Astrophys. J. Suppl. Ser. 192, 18. Kozyryev, I, and N. R. Hutzler (2017), “Precision measure-
Kominis, I K, T. W. Kornack, J. C. Allred, and M. V. Ro- ment of time-reversal symmetry violation with laser-cooled
malis (2003), “A subfemtotesla multichannel atomic mag- polyatomic molecules,” arXiv1705.11020.
netometer,” Nature (London) 422, 596. Kramida, A, Yu. Ralchenko, J. Reader, and NIST ASD Team
Kornack, T W, R. K. Ghosh, and M. V. Romalis (2005), “Nu- (2016), NIST Atomic Spectra Database (version 5.4), [On-
clear spin gyroscope based on an atomic comagnetometer,” line]. Available: http://physics.nist.gov/asd. National In-
Phys. Rev. Lett. 95, 230801. stitute of Standards and Technology, Gaithersburg, MD.
Kornack, T W, and M. V. Romalis (2002), “Dynamics of two Krauss, L, J. Moody, F. Wilczek, and D. E. Morris (1985),
overlapping spin ensembles interacting by spin exchange,” “Calculations for cosmic axion detection,” Phys. Rev. Lett.
Phys. Rev. Lett. 89, 253002. 55, 1797.
Kostelecký, V A, and C. D. Lane (1999), “Constraints Kubiček, K, J. Braun, H. Bruhns, J. R. Crespo López-Urrutia,
on Lorentz violation from clock-comparison experiments,” P. H. Mokler, and J. Ullrich (2012), “High-precision laser-
Phys. Rev. D 60, 116010. assisted absolute determination of X-ray diffraction an-
Kostelecký, V A, and M. Mewes (2002), “Signals for Lorentz gles,” Rev. Sci. Instrum. 83, 013102.
violation in electrodynamics,” Phys. Rev. D 66, 056005. Kubiček, K, P. H. Mokler, V. Mäckel, J. Ullrich, and J. R. C.
Kostelecký, V A, and R. Potting (1995), “CPT, strings, and López-Urrutia (2014), “Transition energy measurements
meson factories,” Phys. Rev. D 51, 3923. in hydrogenlike and heliumlike ions strongly supporting
Kostelecký, V A, and N. Russell (2011), “Data tables for bound-state QED calculations,” Phys. Rev. A 90, 032508.
Lorentz and CPT violation,” Rev. Mod. Phys. 83, 11. Kuchiev, M Yu, and V. Flambaum (2002), “QED radia-
Kostelecký, V A, and N. Russell (2017), “Data tables for tive corrections to parity nonconservation in heavy atoms,”
Lorentz and CPT violation,” arXiv:0801.0287v10. Phys. Rev. Lett. 89 (28), 283002.
Kostelecký, V A, and J. D. Tasson (2011), “Matter-gravity Kuchler, F, E. Babcock, M. Burghoff, T. Chupp, S. De-
couplings and Lorentz violation,” Phys. Rev. D 83, 016013. genkolb, I. Fan, P. Fierlinger, F. Gong, E. Kraegeloh,
Kosteleckỳ, V Alan, Adrian C Melissinos, and Matthew W. Kilian, et al. (2016), “A new search for the atomic EDM
Mewes (2016), “Searching for photon-sector lorentz vio- of 129 Xe at FRM-II,” Hyperfine Interact. 237, 1.
lation using gravitational-wave detectors,” Phys. Lett. B Kuroda, K, and N. Mio (1989), “Test of a composition-
761, 1. dependent force by a free-fall interferometer,” Phys. Rev.
Kostelecký, V Alan, and Arnaldo J. Vargas (2015), “Lorentz Lett. 62, 1941.
and CPT tests with hydrogen, antihydrogen, and related Kuroda, N, S. Ulmer, D. J. Murtagh, S. van Gorp, Y. Na-
systems,” Phys. Rev. D 92, 056002. gata, M. Diermaier, S. Federmann, M. Leali, C. Malbrunot,
Kotler, S, N. Akerman, N. Navon, Y. Glickman, and R. Ozeri V. Mascagna, et al. (2014), “A source of antihydrogen for
(2014), “Measurement of the magnetic interaction between in-flight hyperfine spectroscopy,” Nature Commun. 5, 3089.
two electrons,” Nature (London) 510, 376. Kusenko, A, and P. J. Steinhardt (2001), “Q-ball candi-
Kotler, S, R. Ozeri, and D. F. Jackson Kimball (2015), “Con- dates for self-interacting dark matter,” Phys. Rev. Lett.
straints on exotic dipole-dipole couplings between electrons 87, 141301.
at the micrometer scale,” Phys. Rev. Lett. 115, 081801. Lamm, H (2017), “P −state Positronium for Precision
Kozhedub, Y S, O. V. Andreev, V. M. Shabaev, I. I. Tupitsyn, Physics: an Ultrafine Splitting at α6 ,” ArXiv e-prints
C. Brandau, C. Kozhuharov, G. Plunien, and T. Stöhlker arXiv:1705.10431 [physics.atom-ph].
(2008), “Nuclear deformation effect on the binding energies Lamoreaux, S K (1989), “Optical pumping technique for mea-
in heavy ions,” Phys. Rev. A 77, 032501. suring small nuclear quadrupole shifts in 1 S0 atoms and
Kozlov, A, V. A. Dzuba, and V. V. Flambaum (2013), testing spatial anisotropy,” Nucl. Inst. and Meth. in Phys.
“Prospects of biulding optical atomic clocks using Er I or Research A 284, 43.
Er III,” Phys. Rev. A 88, 032509. Lamoreaux, S K (1997), “Demonstration of the Casimir force
Kozlov, M G (2009), “Λ -doublet spectra of diatomic radicals in the 0.6 to 6 µm range,” Phys. Rev. Lett. 78, 5.
and their dependence on fundamental constants,” Phys. Lamoreaux, S K (2004), “The Casimir force: background,
Rev. A 80, 022118. experiments, and applications,” Rep. Prog. Phys. 68, 201.
Kozlov, M G, V. A. Korol, J. C. Berengut, V. A. Dzuba, and Lamoreaux, S K (2012), “The Casimir force and related ef-
V. V. Flambaum (2004), “Space-time variation of the fine- fects: The status of the finite temperature correction and
structure constant and evolution of isotope abundances,” limits on new long-range forces,” Ann. Rev. Nucl. Part. Sci.
Phys. Rev. A 70, 062108. 62, 37.
Kozlov, M G, L. N. Labzovskii, and A. O. Mitruschenkov Landau, L D (1948), Dokl. Akad. Nauk SSSR 60, 207.
(1991), “Parity nonconservation in diatomic molecules is a Langacker, Paul (2009), “The physics of heavy Z 0 gauge
strong constant magnetic field,” Sov. Phys. JETP 73, 415. bosons,” Rev. Mod. Phys. 81, 1199.
Kozlov, M G, and L. N. Labzowsky (1995), “Parity violation Laporte, O (1924), “Die struktur des eisenspektrums (I),” Z.
effects in diatomics,” J. Phys. B 28, 1933. Physik 23, 135.
102

Le Dall, Matthias, Maxim Pospelov, and Adam Ritz (2015), precise single redshift bound to ∆α/α,” Astron. Astrophys.
“Sensitivity to light weakly-coupled new physics at the pre- 449, 879.
cision frontier,” Phys. Rev. D 92, 16010. Levshakov, S A, F. Combes, F. Boone, I. I. Agafonova,
Le Targat, R, L. Lorini, Y. Le Coq, M. Zawada, J. Guéna, D. Reimers, and M. G. Kozlov (2012), “An upper limit
M. Abgrall, M. Gurov, P. Rosenbusch, D. G. Rovera, to the variation in the fundamental constants at redshift z
B. Nagórny, et al. (2013), “Experimental realization of an = 5.2,” Astron. Astrophys. 540, L9.
optical second with strontium lattice clocks,” Nature Com- Levshakov, S A, M. G. Kozlov, and D. Reimers (2011a),
mun. 4, 2109. “Methanol as a tracer of fundamental constants,” Astro-
Leanhardt, A E, J. L. Bohn, H. Loh, P. Maletinsky, E. R. phys. J. 738, 26.
Meyer, L. C. Sinclair, R. P. Stutz, and E. A. Cornell (2011), Levshakov, S A, A. V. Lapinov, C. Henkel, P. Molaro,
“High-resolution spectroscopy on trapped molecular ions in D. Reimers, M. G. Kozlov, and I. I. Agafonova (2010a),
rotating electric fields: A new approach for measuring the “Searching for chameleon-like scalar fields with the ammo-
electron electric dipole moment,” J. Mol. Spectrosc. 270, nia method. II. Mapping of cold molecular cores in NH3
1. and HC3 N lines,” Astron. Astrophys. 524, A32.
Ledbetter, M P, S. Pustelny, D. Budker, M. V. Romalis, J. W. Levshakov, S A, P. Molaro, M. G. Kozlov, A. V. Lapinov,
Blanchard, and A. Pines (2012), “Liquid-state nuclear spin C. Henkel, D. Reimersi, T. Sakai, and I. I. Agafonova
comagnetometers,” Phys. Rev. Lett. 108, 243001. (2011b), “Searching for chameleon-like scalar fields,” As-
Ledbetter, M P, M. V. Romalis, and D. F. Jackson Kimball trophysics and Space Science Proceedings 22, 103.
(2013), “Constraints on short-range spin-dependent inter- Levshakov, S A, P. Molaro, A. V. Lapinov, D. Reimers,
actions from scalar spin-spin coupling in deuterated molec- C. Henkel, and T. Sakai (2010b), “Searching for
ular hydrogen,” Phys. Rev. Lett. 110, 040402. chameleon-like scalar fields with the ammonia method,”
Lee, T D, and C. N. Yang (1956), “Question of parity con- Astron. Astrophys. 512, A44.
servation in weak interactions,” Phys. Rev. 104, 254. Levshakov, S A, P. Molaro, and D. Reimers (2010c), “Search-
Leefer, N, L. Bougas, D. Antypas, and D. Budker (2014), ing for spatial variations of α2 /µ in the milky way,” Astron.
“Towards a new measurement of parity violation in dys- Astrophys. 516, A113.
prosium,” arXiv:1412.1245. Levshakov, S A, D. Reimers, C. Henkel, B. Winkel,
Leefer, N, A Gerhardus, D Budker, VV Flambaum, and A. Mignano, M. Centurión, and P. Molaro (2013), “Lim-
YV Stadnik (2016), “Search for the effect of massive bodies its on the spatial variations of the electron-to-proton mass
on atomic spectra and constraints on Yukawa-type inter- ratio in the galactic plane,” Astron. Astrophys. 559, A91.
actions of scalar particles,” Phys. Rev. Lett. 117, 271601. Levshakov, S A, D. Reimers, M. G. Kozlov, S. G. Porsev, and
Leefer, N, C. T. M. Weber, A. Cingöz, J. R. Torgerson, and P. Molaro (2008), “A new approach for testing variations
D. Budker (2013), “New limits on variation of the fine- of fundamental constants over cosmic epochs using fir fine-
structure constant using atomic dysprosium,” Phys. Rev. structure lines,” Astron. Astrophys. 479, 719.
Lett. 111, 060801. Lewenstein, M, A. Sanpera, V. Ahufinger, B. Damski, A. Sen,
Leggett, A J (1978), “Macroscopic effect of P- and T- and U. Sen (2007), “Ultracold atomic gases in optical lat-
nonconserving interactions in ferroelectrics: A possible ex- tices: mimicking condensed matter physics and beyond,”
periment?” Phys. Rev. Lett. 41, 586. Adv. Phys. 56, 243.
Lehnert, R, W. M. Snow, and H. Yan (2014), “A first ex- Lewis, A D, D. A. Buote, and J. T. Stocke (2003), “Chandra
perimental limit on in-matter torsion from neutron spin observations of a2029: The dark matter profile down to
rotation in liquid 4 He,” Phys. Lett. B 730, 353. below 0.01 rvir in an unusually relaxed cluster,” Astrophys.
Lehnert, R, W. M. Snow, and H. Yan (2015), “Corrigendum J. 586, 135.
to: “A first experimental limit on in-matter torsion from Liao, Yi, and Ji-Yuan Liu (2007), “Long-range electron spin-
neutron spin rotation in liquid 4 He” [Phys. Lett. B 730 spin interactions from unparticle exchange,” Phys. Rev.
(2014) 353],” Phys. Lett. B 744, 415. Lett. 99, 191804.
Leitner, J, and S. Okubo (1964), “Parity, charge conjugation, Liberati, S, and L. Maccione (2009), “Lorentz violation: Mo-
and time reversal in the gravitational interaction,” Phys. tivation and new constraints,” Annu. Rev. Nucl. Part. Sci.
Rev. 136, B1542. 59, 245.
Lense, J, and H. Thirring (1918), “On the influence of the Linde, A D (1988), “Inflation and axion cosmology,” Phys.
proper rotation of a central body on the motion of the Lett. B 201, 437.
planets and the moon, according to Einstein’s theory of Linder, E V (2008), “The dynamics of quintessence, the
gravitation,” Phys. Z. 19, 156. quintessence of dynamics,” Gen. Rel. Grav. 40, 329.
Leslie, T M, E. Weisman, R. Khatiwada, and J. C. Long Lintz, M, J. Guena, and M. A. Bouchiat (2007), “Pump-
(2014), “Prospects for electron spin-dependent short-range probe measurement of atomic parity violation in cesium
force experiments with rare earth iron garnet test masses,” with a precision of 2.6%,” Europ. Phys. J. A 32, 525.
Phys. Rev. D 89, 114022. Liu, Yu-Sheng, David McKeen, and Gerald A Miller (2016),
Letokhov, V S (1975), “On difference of energy levels of left “Electrophobic scalar boson and muonic puzzles,” Phys.
and right molecules due to weak interactions,” Phys. Lett. Rev. Lett. 117, 101801.
A 53, 275. Lo, Anthony, Philipp Haslinger, Eli Mizrachi, Loı̈c Anderegg,
Levi, F, D. Calonico, C. E. Calosso, A. Godone, S. Mical- Holger Müller, Michael Hohensee, Maxim Goryachev, and
izio, and G. A. Costanzo (2014), “Accuracy evaluation of Michael E Tobar (2016), “Acoustic tests of lorentz symme-
ITCsF2: a nitrogen cooled caesium fountain,” Metrologia try using quartz oscillators,” Phys. Rev. X 6, 011018.
51, 270. Loh, Huanqian, Kevin C. Cossel, M. C. Grau, K.-K. Ni, Ed-
Levshakov, S A, M. Centurión, P. Molaro, S. D’Odorico, mund R. Meyer, John L. Bohn, Jun Ye, and Eric A. Cor-
D. Reimers, R. Quast, and M. Pollmann (2006), “Most nell (2013), “Precision spectroscopy of polarized molecules
103

in an ion trap,” Science 342, 1220. optical lattice,” New J. Phys. 17, 055004.
Lüders, Gerhart, and Bruno Zumino (1957), “Some conse- Meekhof, D M, P A Vetter, P K Majumder, S K Lamore-
quences of TCP-invariance,” Phys. Rev. 106, 385. aux, and E N Fortson (1993), “High-precision measure-
Ludlow, A D, M. M. Boyd, J. Ye, E. Peik, and P. O. Schmidt ment of parity nonconserving optical rotation in atomic
(2015), “Optical atomic clocks,” Rev. Mod. Phys. 87, 637. lead,” Phys. Rev. Lett. 71, 3442.
Lykken, J, and S. Nandi (2000), “Asymmetrical large extra Meyer, Edmund, and John Bohn (2008), “Prospects for an
dimensions,” Phys. Lett. B 485, 224. electron electric-dipole moment search in metastable ThO
Mäckel, V, R. Klawitter, G. Brenner, J. R. Crespo López- and ThF+ ,” Phys. Rev. A 78, 010502.
Urrutia, and J. Ullrich (2011), “Laser spectroscopy on for- Meyer, Edmund, John Bohn, and Michael Deskevich (2006),
bidden transitions in trapped highly charged Ar13+ ions,” “Candidate molecular ions for an electron electric dipole
Phys. Rev. Lett. 107, 143002. moment experiment,” Phys. Rev. A 73 (6), 062108.
Macpherson, M, K. Zetie, R. B. Warrington, D. N. Stacey, Michelson, Albert A (1881), “The relative motion of the Earth
and J. Hoare (1991), “Precise measurement of parity non- and the luminiferous ether,” Am. J. Sci. 22, 120.
conserving optical rotation at 876 nm in atomic bismuth,” Michelson, Albert A, and Abraham Morley (1887), “On the
Phys. Rev. Lett. 67, 27847. relative motion of the Earth and the luminiferous ether,”
Majorovits, B, J. Redondo, et al. (2016), “Madmax: A new American Journal of Science 34, 333.
dark matter axion search using a dielectric haloscope,” Michishio, K, T. Kanai, S. Kuma, T. Azuma, K. Wada,
arXiv:1611.04549. I. Mochizuki, T. Hyodo, A. Yagishita, and Y. Nagashima
Mantry, S, M. Pitschmann, and M. J. Ramsey-Musolf (2014), (2016), “Observation of a shape resonance of the positron-
“Distinguishing axions from generic light scalars using elec- ium negative ion,” Nature Commun. 7, 11060.
tric dipole moment and fifth-force experiments,” Phys. Rev. Milgrom, M (1983), “A modification of the Newtonian dy-
D 90, 054016. namics as a possible alternative to the hidden mass hy-
Marciano, W J (1995), in Precision Tests of the Standard pothesis,” Astrophys. J. 270, 365.
Electroweak Model, edited by P. Langacker (World Scien- Mills, Jr, A P (2014), “Optical spectroscopy of atomic and
tific, Singapore) p. 170. molecular positronium,” in Journal of Physics Conference
Marciano, W J, and A I Sanda (1978), “Parity violation in Series, Vol. 488, p. 012001.
atoms induced by radiative corrections,” Phys. Rev. D 17, Mills, Allen P (1981), “Observation of the positronium nega-
3055. tive ion,” Phys. Rev. Lett. 46, 717.
Martin, W C, and W. L. Wiese (2002), Atomic, Molec- Milstein, A I, O. P Sushkov, and I. S Terekhov (2002),
ular, and Optical Physics Handbook (version 2.2). “Radiative corrections and parity nonconservation in heavy
[Online]. Available at https://www.nist.gov/pml/atomic- atoms,” Phys. Rev. A 89, 22108.
spectroscopy-compendium-basic-ideas-notation-data-and- Milstein, A I, O. P. Sushkov, and I. S. Terekhov (2003),
formulas. National Institute of Standards and Technology, “Calculation of radiative corrections to the effect of parity
Gaithersburg, MD. nonconservation in heavy atoms,” Phys. Rev. A 67, 62103.
Martins, C J A P (2015), “Fundamental cosmology in the E- Miyazaki, A, T. Yamazaki, T. Suehara, T. Namba, S. Asai,
ELT era: the status and future role of tests of fundamental T. Kobayashi, H. Saito, Y. Tatematsu, I. Ogawa, and
coupling stability,” Gen. Relativ. Gravit. 47, 1843. T. Idehara (2015), “First millimeter-wave spectroscopy of
Marton, J, S. Bartalucci, A. Bassi, M. Bazzi, S. Bertolucci, ground-state positronium,” Progress of Theoretical and Ex-
C. Berucci, M. Bragadireanu, M. Cargnelli, A. Clozza, perimental Physics 2015, 011C01.
C. Curceanu, et al. (2017), “VIP-2 at LNGS: An exper- Mohr, P J, D. B. Newell, and B. N. Taylor (2016), “CODATA
iment on the validity of the Pauli exclusion principle for recommended values of the fundamental physical constants:
electrons,” arXiv:1703.01615. 2014*,” Rev. Mod. Phys. 88, 035009.
Marton, J, S. Bartalucci, S. Bertolucci, C. Berucci, M. Bra- Mohr, P J, B. N. Taylor, and D. B. Newell (2012), “CODATA
gadireanu, M. Cargnelli, C. Curceanu (Petrascu, S. Di recommended values of the fundamental physical constants:
Matteo, J.-P. Egger, C. Guaraldo, et al. (2013), “Testing 2010,” Rev. Mod. Phys. 84, 1527.
the Pauli exclusion principle for electrons,” in Journal of Moody, J E, and F. Wilczek (1984), “New macroscopic
Physics Conference Series, Vol. 447, p. 012070. forces?” Phys. Rev. D 30, 130.
Massó, E, and J. Redondo (2005), “Evading astrophysical Mooser, A, S. Ulmer, K. Blaum, K. Franke, H. Kracke, C. Lei-
constraints on axion-like particles,” Journal of Cosmology teritz, W. Quint, C. C. Rodegheri, C. Smorra, and J. Walz
and Astroparticle Physics 2005, 15. (2014), “Direct high-precision measurement of the mag-
Masuda, M, and M. Sasaki (2009), “Limits on nonstandard netic moment of the proton,” Nature (London) 509, 596.
forces in the submicrometer range,” Phys. Rev. Lett. 102, Morgan, T A, and A. Peres (1962), “Direct test for the strong
171101. equivalence principle,” Phys. Rev. Lett. 9, 79.
Mathavan, Sreekanth C, Artem Zapara, Quinten Esajas, Mukhamedjanov, T N, and O. P. Sushkov (2005), “Suggested
and Steven Hoekstra (2016), “Deceleration of a supersonic search for 207 Pb nuclear Schiff moment in PbTiO3 ferro-
beam of SrF molecules to 120 ms−1 ,” ChemPhysChem 17, electric,” Phys. Rev. A 72, 34501.
3709. Müller, H, S.-W. Chiow, S. Herrmann, S. Chu, and K.-Y.
Matloob, R, and H. Falinejad (2001), “Casimir force between Chung (2008), “Atom-Interferometry Tests of the Isotropy
two dielectric slabs,” Phys. Rev. A 64, 042102. of Post-Newtonian Gravity,” Phys. Rev. Let. 100 (3),
Mattingly, D (2005), “Modern tests of Lorentz invariance,” 031101.
Living Reviews in Relativity 8, 5. Müller, H, A. Peters, and S. Chu (2010), “A precision mea-
McGuyer, B H, M. McDonald, G. Z. Iwata, M. G. Tarallo, surement of the gravitational redshift by the interference
A. T. Grier, F. Apfelbeck, and T. Zelevinsky (2015), of matter waves,” Nature (London) 463, 926.
“High-precision spectroscopy of ultracold molecules in an
104

Müller, H, P. L. Stanwix, M. E. Tobar, E. Ivanov, P. Wolf, Namba, Toshio (2012), “Precise measurement of positron-
S. Herrmann, A. Senger, E. Kovalchuk, and A. Pe- ium,” Progress of Theoretical and Experimental Physics
ters (2007), “Tests of relativity by complementary rotat- 2012, 04D003.
ing Michelson-Morley experiments,” Phys. Rev. Lett. 99, Nanopoulos, D V, Asim Yildiz, and Paul H Cox (1979), “On
050401. the electric dipole moment of the neutron,” Phys. Lett. B
Müntinga, H, H. Ahlers, M. Krutzik, A. Wenzlawski, 87, 53.
S. Arnold, D. Becker, K. Bongs, H. Dittus, H. Duncker, Nemitz, N, T. Ohkubo, M. Takamoto, I. Ushijima, M. Das,
N. Gaaloul, et al. (2013), “Interferometry with Bose- N. Ohmae, and H. Katori (2016), “Frequency ratio of Yb
Einstein Condensates in Microgravity,” Phys. Rev. Lett. and Sr clocks with 5 × 10−17 uncertainty at 150 seconds
110, 093602. averaging time,” Nature Photonics 10, 258.
Murata, J, and S. Tanaka (2015), “A review of short-range Neronov, Yu I, and N. N. Seregin (2014), “Determination of
gravity experiments in the LHC era,” Class. and Quantum the spinspin coupling constant of the HD isotopologue of
Gravity 32, 033001. hydrogen for the estimate of existence of nonelectromag-
Murphy, M T, V. V. Flambaum, S. Muller, and C. Henkel netic spin-dependent interaction,” JETP Lett. 100, 609.
(2008a), “Strong limit on a variable proton-to-electron Nesvizhevsky, V V, G. Pignol, and K. V. Protasov (2008),
mass ratio from molecules in the distant universe,” Science “Neutron scattering and extra-short-range interactions,”
320, 1611. Phys. Rev. D 77, 034020.
Murphy, M T, V. V. Flambaum, J. K. Webb, V. A. Dzuba, Neville, D E (1980), “Experimental bounds on the coupling
J. X. Prochaska, and A. M. Wolfe (2004), “Constraining strength of torsion potentials,” Phys. Rev. D 21, 2075.
variations in the fine-structure constant, quark masses and Neville, D E (1982), “Experimental bounds on the coupling
the strong interaction,” in Astrophysics, Clocks and Fun- of massless spin-1 torsion,” Phys. Rev. D 25, 573.
damental Constants, Lecture Notes in Physics, Vol. 648, Newman, R D, E. C. Berg, and P. E. Boynton (2009), “Tests
edited by S. G. Karshenboim and E. Peik (Springer, Berlin) of the gravitational inverse square law at short ranges,”
p. 131. Space Science Reviews 148, 175.
Murphy, M T, C. R. Locke, P. S. Light, A. N. Luiten, and Nguyen, A T, D. Budker, D. DeMille, and M. Zolotorev
J. S. Lawrence (2012), “Laser frequency comb techniques (1997), “Search for parity nonconservation in atomic dys-
for precise astronomical spectroscopy,” Mon. Not. R. As- prosium,” Phys. Rev. A 56, 3453.
tron. Soc. 422, 761. Ni, W T, T. C. P. Chui, S.-S. Pan, and B.-Y. Cheng (1994),
Murphy, M T, J. K. Webb, and V. V. Flambaum (2007), “Search for anomalous spin-spin interactions between elec-
“Comment on “limits on the time variation of the elec- trons using a dc SQUID,” Physica B (Amsterdam) 194,
tromagnetic fine-structure constant in the low energy limit 153.
from absorption lines in the spectra of distant quasars”,” Ni, W-T, S.-S. Pan, H.-C. Yeh, L.-S. Hou, and J. Wan (1999),
Phys. Rev. Lett. 99, 239001. “Search for an axionlike spin coupling using a paramagnetic
Murphy, M T, J. K. Webb, and V. V. Flambaum (2008b), salt with a dc SQUID,” Phys. Rev. Lett. 82, 2439.
“Revision of VLT/UVES constraints on a varying fine- Nicholson, T L, S. L. Campbell, R. B. Hutson, G. E. Marti,
structure constant,” Mon. Not. R. Astron. Soc. 384, 1053. B. J. Bloom, R. L. McNally, W. Zhang, M. D. Bar-
Murphy, M T, J. K. Webb, and V. V. Flambaum (2008c), rett, M. S. Safronova, G. F. Strouse, W. L. Tew, and
“Revisiting VLT/UVES constraints on a varying fine- J. Ye (2015), “Systematic evaluation of an atomic clock at
structure constant,” in Precision Spectroscopy in Astro- 2×10−18 total uncertainty,” Nature Commun. 6, 6896.
physics, edited by N. C. Santos, L. Pasquini, A. C. M. Nicolis, A, R. Rattazzi, and E. Trincherini (2009), “Galileon
Correia, and M. Romaniello, p. 95. as a local modification of gravity,” Phys. Rev. D 79, 064036.
Murphy, M T, J. K. Webb, V. V. Flambaum, M. J. Drinkwa- Nobili, A M, M. Shao, R. Pegna, G. Zavattini, S. G. Turyshev,
ter, F. Combes, and T. Wiklind (2001a), “Improved con- D. M. Lucchesi, A. De Michele, S. Doravari, G. L. Comandi,
straints on possible variation of physical constants from H I Saravanan, et al. (2012), “Galileo Galilei (GG): space test
21-cm and molecular QSO absorption lines,” Mon. Not. R. of the weak equivalence principle to 10−17 and laboratory
Astron. Soc. 327, 1244. demonstrations,” Class. Quantum Gravity 29, 184011.
Murphy, M T, J. K. Webb, V. V. Flambaum, J. X. Prochaska, Norrgard, E B, E. R. Edwards, D. J. McCarron, M. H. Stei-
and A. M. Wolfe (2001b), “Further constraints on variation necker, D. DeMille, Shah Saad Alam, S. K. Peck, N. S.
of the fine-structure constant from alkali-doublet QSO ab- Wadia, and L. R. Hunter (2017), “Hyperfine structure of
sorption lines,” Mon. Not. R. Astron. Soc. 327, 1237. the B 3 Π1 state and predictions of optical cycling behavior
Myers, R C, and M. Pospelov (2003), “Ultraviolet modifica- in the X→B transition of TlF,” Phys. Rev. A 95, 062506.
tions of dispersion relations in effective field theory,” Phys. Nörtershäuser, W, C. Geppert, A. Krieger, K. Pachucki,
Rev. Lett. 90, 211601. M. Puchalski, K. Blaum, M. L. Bissell, N. Frömmgen,
Nagasawa, M, and M. Kawasaki (1994), “Collapse of axionic M. Hammen, M. Kowalska, et al. (2015), “Precision test
domain wall and axion emission,” Phys. Rev. D 50, 4821. of many-body QED in the Be+ 2p fine structure doublet
Nagashima, Y (2014), “Experiments on positronium negative using short-lived isotopes,” Phys. Rev. Lett. 115, 033002.
ions,” Phys. Rep. 545, 95. Nuñez Portela, M, J. E. van den Berg, H. Bekker, O. Böll,
Nagel, M, S. R. Parker, E. V. Kovalchuk, P. L. Stanwix, J. G. E. A. Dijck, G. S. Giri, S. Hoekstra, K. Jungmann, A. Mo-
Hartnett, E. N. Ivanov, A. Peters, and M. E. Tobar (2015), hanty, C. J. G. Onderwater, et al. (2013), “Towards a pre-
“Direct terrestrial test of Lorentz symmetry in electrody- cise measurement of atomic parity violation in a single Ra+
namics to 10−18 ,” Nature Commun. 6, 8174. ion,” Hyperfine Interact. 214, 157.
Nakai, Yuichiro, and Matthew Reece (2016), “Electric dipole Olive, K A, M. Peloso, and A. J. Peterson (2012), “Where are
moments in natural supersymmetry,” arXiv1612.08090. the walls? Spatial variation in the fine-structure constant,”
Phys. Rev. D 86, 043501.
105

Olive, K A, M. Peloso, and J.-P. Uzan (2011), “Wall of fun- Peebles, P J E, and B. Ratra (2003), “The cosmological con-
damental constants,” Phys. Rev. D 83, 043509. stant and dark energy,” Rev. Mod. Phys. 75, 559.
Olive, K A, and M. Pospelov (2008), “Environmental depen- Peik, E, and M. Okhapkin (2015), “Nuclear clocks based on
dence of masses and coupling constants,” Phys. Rev. D 77, resonant excitation of gamma-transitions,” C. R. Physique
043524. 16, 516.
Olive, KA, et al. (2015), “Review of particle physics,” Particle Peik, E, and C. Tamm (2003), “Nuclear laser spectroscopy of
Data Group, Chin. Phys. C, 38, 090001 (2014) and 2015 the 3.5 eV transition in 229 Th,” Europhys. Lett. 61, 181.
update. (URL: http://pdg.lbl.gov). Peil, S, S. Crane, J. L. Hanssen, T. B. Swanson, and C. R. Ek-
Ong, A, J. C. Berengut, and V. V. Flambaum (2014), “Op- strom (2013), “Tests of local position invariance using con-
tical transitions in highly charged ions for detection of tinuously running atomic clocks,” Phys. Rev. A 87, 010102.
variations in the fine-structure constant,” in Fundamen- Peres, A (1978), “Test of equivalence principle for particles
tal Physics in Particle Traps, Springer Tracts in Mod- with spin,” Phys. Rev. D 18, 2739.
ern Physics, Vol. 256, edited by W. Quint and M. Vogel Pérez, P, D. Banerjee, F. Biraben, D. Brook-Roberge,
(Springer, Berlin - Heidelberg) p. 293. M. Charlton, P. Cladé, P. Comini, P. Crivelli, O. Dalka-
Onofrio, R (2006), “Casimir forces and non-Newtonian grav- rov, P. Debu, et al. (2015), “The GBAR antimatter gravity
itation,” New J. Phys. 8, 237. experiment,” Hyperfine Interact. 233, 21.
Onofrio, Roberto (2013), “Proton radius puzzle and quantum Perlmutter, S (2012), “Nobel lecture: Measuring the accelera-
gravity at the Fermi scale,” Europhys. Lett. 104, 20002. tion of the cosmic expansion using supernovae,” Rev. Mod.
Overduin, J, F. Everitt, P. Worden, and J. Mester (2012), Phys. 84, 1127.
“Step and fundamental physics,” Class. Quantum Gravity Perlmutter, S, G. Aldering, G. Goldhaber, R. A. Knop, P. Nu-
29, 184012. gent, P. G. Castro, S. Deustua, S. Fabbro, A. Goobar, D. E.
Padmanabhan, T (2003), “Cosmological constant the weight Groom, et al. (1999), “Measurements of ω and λ from 42
of the vacuum,” Phys. Rep. 380, 235. high-redshift supernovae,” Astrophys. J. 517, 565.
Pal, Rupsi, Dansha Jiang, M S Safronova, and U I Safronova Peskin, Michael E, and Daniel V. Schroeder (1995), An intro-
(2009), “Calculation of parity-nonconserving amplitude duction to Quantum Field Theory (Perseus Books, Read-
and other properties of Ra+ ,” Phys. Rev. A 79, 62505. ing, Massachusetts).
Panda, C D, B. R. O’Leary, A. D. West, J. Baron, P. W. Peters, Achim, Keng Yeow Chung, and Steven Chu (1999),
Hess, C. Hoffman, E. Kirilov, C. B. Overstreet, E. P. West, “Measurement of gravitational acceleration by dropping
D. DeMille, J. M. Doyle, and G. Gabrielse (2016), “Stim- atoms,” Nature (London) 400, 849.
ulated Raman adiabatic passage preparation of a coherent Petukhov, A K, G. Pignol, D. Jullien, and K. H. Ander-
superposition of ThO H 3 ∆1 states for an improved elec- sen (2010), “Polarized 3 He as a probe for short-range spin-
tron electric-dipole-moment measurement,” Phys. Rev. A dependent interactions,” Phys. Rev. Lett. 105, 170401.
93 (5), 52110. Phillips, W D (1998), “Nobel lecture: Laser cooling and trap-
Parthey, C G, A. Matveev, J. Alnis, A. Beyer, R. Pohl, ping of neutral atoms,” Rev. Mod. Phys. 70, 721.
K. Predehl, T. Udem, N. Kolachevsky, M. Abgrall, Phipp, S J, N. H. Edwards, P. E. G. Baird, and S. Nakayama
D. Rovera, C. Salomon, P. Laurent, and T. W. Hänsch (1996), “A measurement of parity non-conserving optical
(2011), “Precision spectroscopy on atomic hydrogen,” rotation in atomic lead,” J. Phys. B 29, 1861.
in Society of Photo-Optical Instrumentation Engineers Piegsa, F M, and G. Pignol (2012), “Limits on the axial
(SPIE) Conference Series, Proceedings of the SPIE, Vol. coupling constant of new light bosons,” Phys. Rev. Lett.
8132, p. 813202. 108, 181801.
Patrignani, C et al (2016), “Review of particle physics,” Chin. Pihan-Le Bars, H, C. Guerlin, Q. G. Bailey, S. Bize, and
Phys. C40 (10), 100001. P. Wolf (2017), “Improved tests of Lorentz invariance in
Pauli, W (1924), “Zur Frage der theoretischen Deutung der the matter sector using atomic clocks,” arXiv:1701.06902.
Satelliten einiger Spektrallinien und ihrer Beeinflussung Pilipenko, S V (2013), “Paper-and-pencil cosmological calcu-
durch magnetische Felder,” Naturwissenschaften 12, 741. lator,” arXiv:1303.5961.
Pašteka, L F, A. Borschevsky, V. V. Flambaum, and P. Schw- Pohl, Randolf (2016), “Laser spectroscopy of muonic hydro-
erdtfeger (2015), “Search for the variation of fundamental gen and the puzzling proton,” J. Phys. Soc. Japan 85,
constants: Strong enhancements in X 2 Π cations of dihalo- 091003.
gens and hydrogen halides,” Phys. Rev. A 92, 012103. Pohl, Randolf, Aldo Antognini, François Nez, Fernando D.
Peccei, R (2008), “The strong CP problem and axions,” Lect. Amaro, François Biraben, João M. R. Cardoso, Daniel S.
Notes Phys. 741, 3. Covita, Andreas Dax, Satish Dhawan, Luis M. P. Fernan-
Peccei, R, and H. Quinn (1977a), “Constraints imposed by des, et al. (2010), “The size of the proton,” Nature (Lon-
CP conservation in the presence of pseudoparticles,” Phys. don) 466 (7303), 213–216.
Rev. D 16, 1791. Pohl, Randolf, Ronald Gilman, Gerald a Miller, and
Peccei, R D, and Helen R Quinn (1977b), “CP conservation in Krzysztof Pachucki (2013a), “Muonic Hydrogen and the
the presence of pseudoparticles,” Phys. Rev. Lett. 38 (25), Proton Radius Puzzle,” Annu. Rev. Nucl. Part. Sci. 63 (1),
1440. 175–204, arXiv:arXiv:1301.0905v2.
Peck, S K, D. K. Kim, D. Stein, D. Orbaker, A. Foss, M. T. Pohl, Randolf, Ronald Gilman, Gerald A. Miller, and
Hummon, and L. R. Hunter (2012), “Limits on local Krzysztof Pachucki (2013b), “Muonic hydrogen and the
Lorentz invariance in mercury and cesium,” Phys. Rev. A proton radius puzzle,” Ann. Rev. Nucl. Part. Sci. 63, 175.
86, 012109. Pohl, Randolf, François Nez, Luis M. P. Fernandes, Fer-
Peck, S K, N. Lane, D. G. Ang, and L. R. Hunter (2016), nando D. Amaro, François Biraben, João M. R. Cardoso,
“Using tensor light shifts to measure and cancel a cells Daniel S. Covita, Andreas Dax, Satish Dhawan, Marc
quadrupolar frequency shift,” Phys. Rev. A 93, 023426. Diepold, et al. (2016), “Laser spectroscopy of muonic deu-
106

terium,” Science 353, 669. 67.


Pokotilovski, Yu N (2006), “Constraints on new interactions Quintero-Pérez, M, T. E. Wall, S. Hoekstra, and H. L. Beth-
from neutron scattering experiments,” Physics of Atomic lem (2014), “Preparation of an ultra-cold sample of ammo-
Nuclei 69, 924. nia molecules for precision measurements,” J. Mol. Spec-
Poli, N, C. W. Oates, P. Gill, and G. M. Tino (2013), “Optical trosc. 300, 112.
atomic clocks,” Nuovo Cimento Rivista Serie 36, 555. Rabey, I, J. Devlin, B. Sauer, J. Hudson, M. Tarbutt, and
Poli, N, F.-Y. Wang, M. G. Tarallo, A. Alberti, M. Prevedelli, E. Hinds (2016), “Sensitivity improvements to the YbF
and G. M. Tino (2011), “Precision measurement of gravity electron electric dipole moment experiment,” in APS Meet-
with cold atoms in an optical lattice and comparison with ing Abstracts.
a classical gravimeter,” Phys. Rev. Lett. 106, 038501. Raffelt, G (2012), “Limits on a CP-violating scalar axion-
Pollock, S J, and M C Welliver (1999), “Effects of neutron nucleon interaction,” Phys. Rev. D 86, 015001.
spatial distributions on atomic parity nonconservation in Raffelt, G, and D. Seckel (1988), “Bounds on exotic-particle
cesium,” Phys. Lett. B 464, 177. interactions from SN1987A,” Phys. Rev. Lett. 60, 1793.
Porsev, S G, K. Beloy, and A. Derevianko (2009), “Precision Raffelt, G, and A. Weiss (1995), “Red giant bound on the
determination of electroweak coupling from atomic parity axion-electron coupling reexamined,” Phys. Rev. D 51,
violation and implications for particle physics,” Phys. Rev. 1495.
Lett. 102, 181601. Raffelt, G G (1999), “Particle physics from stars,” Annu. Rev.
Porsev, S G, K. Beloy, and A. Derevianko (2010), “Precision Nucl. Part. Sci. 49, 163.
determination of weak charge of 133 Cs from atomic parity Rahmani, H, R. Srianand, N. Gupta, P. Petitjean, P. No-
violation,” Phys. Rev. D 82, 36008. terdaeme, and D. A. Vásquez (2012), “Constraining the
Porsev, S G, K. V. Koshelev, I. I. Tupitsyn, M. G. Kozlov, variation of fundamental constants at z ∼ 1.3 using 21-cm
D. Reimers, and S. A. Levshakov (2007), “Transition fre- absorbers,” Mon. Not. R. Astron. Soc. 425, 556.
quency shifts with fine-structure-constant variation for fe Ramsey, N F (1979), “The tensor force between two protons
ii: Breit and core-valence correlation corrections,” Phys. at long range,” Physica A (Amsterdam) 96, 285.
Rev. A 76, 052507. Ramsey-Musolf, M J (1999), “Low-energy parity-violation
Pospelov, M, S. Pustelny, M. P. Ledbetter, D. F. Jack- and new physics,” Phys. Rev. C 60, 015501.
son Kimball, W. Gawlik, and D. Budker (2013), “De- Ramsey-Musolf, Michael J, and Shelley A. Page (2006),
tecting domainwalls of axionlike models using terrestrial “Hadronic Parity Violation: A New View Through the
experiments,” Phys. Rev. Lett. 110, 021803. Looking Glass,” Ann. Rev. Nucl. Part. Sci. 56, 1.
Pospelov, Maxim (2009), “Secluded U(1) below the weak Randall, L, and R. Sundrum (1999a), “An alternative to
scale,” Phys. Rev. D 80, 95002. compactification,” Phys. Rev. Lett. 83, 4690.
Pospelov, Maxim, and Adam Ritz (2005), “Electric dipole Randall, L, and R. Sundrum (1999b), “Large mass hierarchy
moments as probes of new physics,” Ann. Phys. (N. Y). from a small extra dimension,” Phys. Rev. Lett. 83, 3370.
318, 119. Redondo, J, and A. Ringwald (2011), “Light shining through
Pospelov, Maxim, and Michael Romalis (2004), “Lorentz in- walls,” Contemporary Physics 52, 211.
variance on trial,” Physics Today 57, 40. Refregier, A (2003), “Weak gravitational lensing by large-
Prescott, C Y, W. B. Atwood, R. L. A. Cottrell, H. DeStae- scale structure,” Annu. Rev. Astron. Astrophys. 41, 645.
bler, Edward L. Garwin, A. Gonidec, R. H. Miller, L. S. Regan, B, Eugene Commins, Christian Schmidt, and David
Rochester, T. Sato, D. J. Sherden, et al. (1978), “Par- DeMille (2002), “New limit on the electron electric dipole
ity non-conservation in inelastic electron scattering,” Phys. moment,” Phys. Rev. Lett. 88, 18.
Lett. B 77, 347. Riess, A G (2012), “Nobel lecture: My path to the accelerat-
Preskill, J, M. B. Wise, and F. Wilczek (1983), “Cosmology ing universe,” Rev. Mod. Phys. 84, 1165.
of the invisible axion,” Phys. Lett. B 120, 127. Riess, A G, A. V. Filippenko, P. Challis, A. Clocchiatti,
Primakoff, H (1951), “Photo-production of neutral mesons A. Diercks, P. M. Garnavich, R. L. Gilliland, C. J. Hogan,
in nuclear electric fields and the mean life of the neutral S. Jha, R. P. Kirshner, et al. (1998), “Observational evi-
meson,” Phys. Rev. 81, 899. dence from supernovae for an accelerating universe and a
Pruttivarasin, T, M. Ramm, S. G. Porsev, I. I. Tupitsyn, cosmological constant,” Astron. J. 116, 1009.
M. S. Safronova, M. A. Hohensee, and H. Häffner (2015), Ritter, R C, C. E. Goldblum, W.-T. Ni, G. T. Gillies, and
“Michelson-Morley analogue for electrons using trapped C. C. Speake (1990), “Experimental test of equivalence
ions to test Lorentz symmetry,” Nature (London) 517, 592. principle with polarized masses,” Phys. Rev. D 42, 977.
Puchalski, M, and K. Pachucki (2014), “Quantum electro- Roberts, B M, Y. V. Stadnik, V. A. Dzuba, V. V. Flam-
dynamics corrections to the 2P fine splitting in Li,” Phys. baum, N. Leefer, and D. Budker (2014a), “Limiting P-odd
Rev. Lett. 113, 073004. interactions of cosmic fields with electrons, protons, and
Purcell, E M, and N. F. Ramsey (1950), “On the possibil- neutrons,” Phys. Rev. Lett. 113, 081601.
ity of electric dipole moments for elementary particles and Roberts, B M, Y. V. Stadnik, V. A. Dzuba, V. V. Flambaum,
nuclei,” Phys. Rev. 78, 807. N. Leefer, and D. Budker (2014b), “Parity-violating inter-
Pustelny, S, D. F. Jackson Kimball, C. Pankow, M. P. Led- actions of cosmic fields with atoms, molecules, and nuclei:
better, P. Wlodarczyk, P. Wcislo, M. Pospelov, J. R. Concepts and calculations for laboratory searches and ex-
Smith, J. Read, W. Gawlik, and D. Budker (2013), “The tracting limits,” Phys. Rev. D 90, 096005.
global network of optical magnetometers for exotic physics Roberts, Benjamin M, Geoffrey Blewitt, Conner Dailey, Mac
(GNOME): A novel scheme to search for physics beyond Murphy, Maxim Pospelov, Alex Rollings, Jeff Sherman,
the standard model,” Annalen der Physik 525, 659. Wyatt Williams, and Andrei Derevianko (2017), “GPS
PVDIS-Collaboration, (2014), “Measurement of parity viola- as a dark matter detector: Orders-of-magnitude improve-
tion in electron-quark scattering,” Nature (London) 506, ment on couplings of clumpy dark matter to atomic clocks,”
107

arXiv:1704.06844. Safronova, M S (2014), “Time trials for fundamental con-


Roberts, BM, V.A. Dzuba, and V.V. Flambaum (2015), stants,” Physics Online Journal , 117.
“Parity and time-reversal violation in atomic systems,” Safronova, M S, V. A. Dzuba, V. V. Flambaum, U. I.
Ann. Rev. Nucl. Part. Sci. 65, 63. Safronova, S. G. Porsev, and M. G. Kozlov (2014a),
Robilliard, C, R. Battesti, M. Fouche, J. Mauchain, A.-M. “Atomic properties of Cd-like and Sn-like ions for the devel-
Sautivet, F. Amiranoff, and C. Rizzo (2007), “No light opment of frequency standards and search for the variation
shining through a wall: Results from a photoregeneration of the fine-structure constant,” Phys. Rev. A 90, 052509.
experiment,” Phys. Rev. Lett. 99, 190403. Safronova, M S, V. A. Dzuba, V. V. Flambaum, U. I.
Romalis, M V, and R. R. Caldwell (2013), “Laboratory search Safronova, S. G. Porsev, and M. G. Kozlov (2014b),
for a quintessence field,” arXiv:1302.1579. “Highly charged ions for atomic clocks, quantum informa-
Romalis, Michael V, Dong Sheng, Brian Saam, and Thad G. tion, and search for α variation,” Phys. Rev. Lett. 113,
Walker (2014), “Comment on “new limit on lorentz- 030801.
invariance- and CPT-violating neutron spin interactions Safronova, M S, V. A. Dzuba, V. V. Flambaum, U. I.
using a free-spin-precession 3 He−129 Xe comagnetometer”,” Safronova, S. G. Porsev, and M. G. Kozlov (2014c), “Study
Phys. Rev. Lett. 113, 188901. of highly-charged Ag-like and In-like ions for the develop-
Rosenband, T, D. B. Hume, P. O. Schmidt, C. W. Chou, ment of atomic clocks and search for α-variation,” Phys.
A. Brusch, L. Lorini, W. H. Oskay, R. E. Drullinger, T. M. Rev. A 90, 042513.
Fortier, J. E. Stalnaker, et al. (2008), “Frequency ratio of Safronova, M S, and W. R. Johnson (2000), “High-precision
Al+ and Hg+ single-ion optical clocks; metrology at the calculation of the parity-nonconserving amplitude in fran-
17th decimal place,” Science 319, 1808. cium,” Phys. Rev. A 62, 022112.
Rosenberry, M, and T. Chupp (2001), “Atomic electric dipole Sahoo, B K (2010), “Ab initio studies of electron correlation
moment measurement using spin exchange pumped masers effects in the atomic parity violating amplitudes in Cs and
of 129 Xe and 3 He,” Phys. Rev. Lett. 86, 22. Fr,” J. Phys. B 43 (8).
Rosi, G (2016), “Challenging the “Big G” measurement with Sainis, S, J. Sage, E. Tiesinga, S. Kotochigova, T. Bergeman,
atoms and light,” J. Phys. B 49, 202002. and D. DeMille (2012), “Detailed spectroscopy of the Cs2
Rosi, G, L. Cacciapuoti, F. Sorrentino, M. Menchetti, a3 Σu + state and implications for measurements sensitive
M. Prevedelli, and G. M. Tino (2015), “Measurement of to variation of the electron-proton mass ratio,” Phys. Rev.
the gravity-field curvature by atom interferometry,” Phys. A 86, 022513.
Rev. Lett. 114, 013001. Sakharov, A D (1967), “Violation of CP invariance, C asym-
Rosi, G, F. Sorrentino, L. Cacciapuoti, M. Prevedelli, and metry, and baryon asymmetry of the universe,” JETP Lett.
G. M. Tino (2014), “Precision measurement of the New- 5, 24.
tonian gravitational constant using cold atoms,” Nature Sakurai, J J, and J. J. Napolitano (2011), Modern Quantum
(London) 510, 518. Mechanics, 2nd ed. (Addison Wesley, Boston).
Rosner, J L (2000), “Atomic parity violation and precision Salucci, P, and A. Borriello (2003), “The intriguing distribu-
electroweak physics - An updated analysis,” Phys. Rev. D tion of dark matter in galaxies,” in Particle Physics in the
61, 016006. New Millennium (Springer) p. 66.
Rosner, Jonathan L (2002), “Role of present and future Salumbides, E J, G. D. Dickenson, T. I. Ivanov, and
atomic parity violation experiments in precision elec- W. Ubachs (2011), “QED effects in molecules: Test on
troweak tests,” Phys. Rev. D 65, 73026. rotational quantum states of H2 ,” Phys. Rev. Lett. 107,
Rouven, Essig, John A. Jaros, William Wester, P. Hans- 043005.
son Adrian, S. Andreas, T. Averett, O. Baker, B. Batell, Salumbides, E J, A. N. Schellekens, B. Gato-Rivera, and
M. Battaglieri, J. Beacham, et al. (2013), Working Group W. Ubachs (2015), “Constraints on extra dimensions
Report: New Light Weakly Coupled Particles, Tech. Rep., from precision molecular spectroscopy,” New J. Phys. 17,
arXiv:1311.0029. 033015.
Rubin, V C, and W. K. Ford Jr. (1970), “Rotation of the Sandars, P G H (1965), “The electric dipole moment of an
andromeda nebula from a spectroscopic survey of emission atom,” Phys. Lett. 14, 13.
regions,” Astrophys. J. 159, 379. Sandars, PGH (1966), “Enhancement factor for the electric
Rubin, V C, W. K. Ford Jr., and N. Thonnard (1980), “Ro- dipole moment of the valence electron in an alkali atom,”
tational properties of 21 SC galaxies with a large range of Phys. Lett. 22, 6.
luminosities and radii, from NGC 4605/R = 4kpc/to UGC Sandars, PGH (1967), “Measurability of the proton electric
2885/R = 122 kpc,” Astrophys. J. 238, 471. dipole moment,” Phys. Rev. Lett. 19, 1396.
Rudolph, J K, S. Bernitt, S. W. Epp, R. Steinbrügge, C. Beil- Santamaria, L, V. Di Sarno, I. Ricciardi, S. Mosca, M. De
mann, G. V. Brown, S. Eberle, A. Graf, Z. Harman, Rosa, G. Santambrogio, P. Maddaloni, and P. De Natale
N. Hell, M. Leutenegger, A. Müller, K. Schlage, H.-C. (2014), “Assessing the time constancy of the proton-to-
Wille, H. Yavaş, J. Ullrich, and J. R. Crespo López-Urrutia electron mass ratio by precision ro-vibrational spectroscopy
(2013), “X-ray resonant photoexcitation: Linewidths and of a cold molecular beam,” J. Mol. Spectrosc. 300, 116.
energies of Kα transitions in highly charged Fe ions,” Phys. Sapirstein, J, K. Pachucki, A. Veitia, and K. T. Cheng (2003),
Rev. Lett. 111, 103002. “Radiative corrections to parity-nonconserving transitions
Ruoso, G, A. Lombardi, A. Ortolan, R. Pengo, C. Braggio, in atoms,” Phys. Rev. A 67, 52110.
G. Carugno, C. S. Gallo, and C. C. Speake (2016), “The Sathyaprakash, B, M. Abernathy, F. Acernese, P. Ajith,
QUAX proposal: a search of galactic axion with magnetic B. Allen, P. Amaro-Seoane, N. Andersson, S. Aoudia,
materials,” in Journal of Physics Conference Series, Vol. K. Arun, P. Astone, and et al. (2012), “Scientific objec-
718, p. 042051. tives of Einstein telescope,” Class. Quantum Gravity 29,
124013.
108

Sato, T, Y. Ichikawa, Y. Ohtomo, Y. Sakamoto, S. Ko- Semertzidis, Y (2017), “The axion dark matter search at
jima, C. Funayama, T. Suzuki, M. Chikamori, E. Hikota, CAPP: a comprehensive approach,” in APS April Meeting
M. Tsuchiya, et al. (2015), “EDM measurement in 129 Xe Abstracts.
atom using dual active feedback nuclear spin maser,” Hy- Shabaev, V M, A. N. Artemyev, V. A. Yerokhin, O. M.
perfine Interact. 230, 147. Zherebtsov, and G. Soff (2001), “Towards a test of QED
Saulson, P R (1984), “Terrestrial gravitational noise on a in investigations of the hyperfine splitting in heavy ions,”
gravitational wave antenna,” Phys. Rev. D 30, 732. Phys. Rev. Lett. 86, 3959.
Schellekens, A N (2013), “Life at the interface of particle Shabaev, V M, D. A. Glazov, N. S. Oreshkina, A. V. Volotka,
physics and string theory,” Rev. Mod. Phys. 85, 1491. G. Plunien, H.-J. Kluge, and W. Quint (2006), “g-factor
Scherk, J (1979), “Antigravity: a crazy idea?” Phys. Lett. B of heavy ions: A new access to the fine structure constant,”
88, 265. Phys. Rev. Lett. 96, 253002.
Schiff, L I (1963), “Measurability of nuclear electric dipole Shabaev, V M, D. A. Glazov, M. B. Shabaeva, V. A. Yerokhin,
moments,” Phys. Rev. 132, 2194. G. Plunien, and G. Soff (2002), “g factor of high-Z lithi-
Schiller, S (2007), “Hydrogenlike highly charged ions for tests umlike ions,” Phys. Rev. A 65, 062104.
of the time independence of fundamental constants,” Phys. Shabaev, V M, K. Pachucki, I. I. Tupitsyn, and
Rev. Lett. 98, 180801. V. A. Yerokhin (2005), “QED corrections to the parity-
Schiller, S, D. Bakalov, and V. I. Korobov (2014), “Simplest nonconserving 6S − 7S amplitude in 133 Cs,” Phys. Rev.
molecules as candidates for precise optical clocks,” Phys. Lett. 94 (21), 213002.
Rev. Lett. 113, 023004. Shabalin, E P (1978), “Electric dipole moment of the quark
Schiller, S, G. M. Tino, P. Gill, C. Salomon, U. Sterr, E. Peik, in a gauge theory with left handed-fermions,” Yad. Fiz. 28,
A. Nevsky, A. Görlitz, D. Svehla, G. Ferrari, et al. (2009), 151, [Sov. J. Nucl. Phys. 28, 75 (1978)].
“Einstein Gravity Explorer - a medium-class fundamental Shelkovnikov, A, R. J. Butcher, C. Chardonnet, and A. Amy-
physics mission,” Experimental Astronomy 23, 573. Klein (2008), “Stability of the proton-to-electron mass ra-
Schlamminger, S (2014), “Fundamental constants: A cool way tio,” Phys. Rev. Lett. 100, 150801.
to measure big G,” Nature (London) 510, 478. Shi, H, S. Bartalucci, S. Bertolucci, C. Berucci, A. M. Bra-
Schlamminger, S, K.-Y. Choi, T. A. Wagner, J. H. Gundlach, gadireanu, M. Cargnelli, A. Clozza, C. Curceanu, L. De
and E. G. Adelberger (2008), “Test of the equivalence prin- Paolis, S. Di Matteo, et al. (2016), “Searches for the vi-
ciple using a rotating torsion balance,” Phys. Rev. Lett. olation of Pauli exclusion principle at LNGS in VIP(-2)
100, 041101. experiment,” in Journal of Physics Conference Series, Vol.
Schlamminger, S, J. H. Gundlach, and R. D. Newman (2015), 718, p. 042055.
“Recent measurements of the gravitational constant as a Shifman, M, A. Vainshtein, and V. Zakharov (1980), “Can
function of time,” Phys. Rev. D 91, 121101. confinement ensure natural CP invariance of strong inter-
Schlippert, D, J. Hartwig, H. Albers, L. L. Richardson, actions?” Nucl. Phys. B 166, 493.
C. Schubert, A. Roura, W. P. Schleich, W. Ertmer, and Sigurdson, K, A. Kurylov, and M. Kamionkowski (2003),
E. M. Rasel (2014), “Quantum test of the universality of “Spatial variation of the fine-structure parameter and the
free fall,” Phys. Rev. Lett. 112, 203002. cosmic microwave background,” Phys. Rev. D 68, 103509.
Schmidt, B P (2012), “Nobel lecture: Accelerating expan- Sikivie, P (1983), “Experimental tests of the “invisible” ax-
sion of the universe through observations of distant super- ion,” Phys. Rev. Lett. 51, 1415.
novae,” Rev. Mod. Phys. 84, 1151. Sikivie, P (1985), “Detection rates for “invisible”-axion
Schmidt, T (1937), “Über die magnetischen momente der searches,” Phys. Rev. D 32, 2988.
atomkerne,” Z. Physik 106, 358. Sikivie, P (2014), “Axion dark matter detection using atomic
Schmöger, L, O. O. Versolato, M. Schwarz, M. Kohnenand, transitions,” Phys. Rev. Lett. 113 (20), 201301.
A. Windberger, B. Piest, S. Feuchtenbeiner, J. Pedregosa- Sikivie, P, N. Sullivan, and D. B. Tanner (2014), “Proposal
Gutierrez, T. Leopold, P. Micke, A. K. Hansen, T. M. for axion dark matter detection using an LC circuit,” Phys.
Baumann, M. Drewsen, J. Ullrich, P. O. Schmidt, and Rev. Lett. 112, 131301.
J. R. Crespo López-Urrutia (2015), “Coulomb crystalliza- Sikivie, Pierre (2012), “The strong CP problem,” C. R. Phys.
tion of highly charged ions,” Science 347, 6227. 13, 176.
Schutte, C. J. H.,, Bertie, J. E., Bunker, P. R., Hougen, J. Skripnikov, L V (2016), “Combined 4-component and rel-
T., Mills, I. M., Watson, J. K. G., and Winnewisser, B. ativistic pseudopotential study of ThO for the electron
P. (1997a), “Notations and conventions in molecular spec- electric dipole moment search,” J. Chem. Phys. 145 (21),
troscopy: Part 1. general spectroscopic notation (iupac rec- 214301.
ommendations 1997),” Pure Appl. Chem. 69, 1633. Smiciklas, M, J. M. Brown, L. W. Cheuk, S. J. Smullin,
Schutte, C. J. H.,, Bertie, J. E., Bunker, P. R., Hougen, J. and M. V. Romalis (2011), “New test of local Lorentz in-
T., Mills, I. M., Watson, J. K. G., and Winnewisser, B. variance using a 21 Ne-Rb-K comagnetometer,” Phys. Rev.
P. (1997b), “Notations and conventions in molecular spec- Lett. 107, 171604.
troscopy: Part 2. symmetry notation (iupac recommenda- Smorra, C, K. Blaum, L. Bojtar, M. Borchert, K. A.
tions 1997),” Pure Appl. Chem. 69, 1641. Franke, T. Higuchi, N. Leefer, H. Nagahama, Y. Matsuda,
Scott, D (2006), “The standard cosmological model,” Can. J. A. Mooser, et al. (2015), “BASE - the baryon antibaryon
Phys. 84, 419. symmetry experiment,” Euro. Phys. J. Spec. Topics 224,
Seidel, ST, M.D. Lachmann, D. Becker, J. Grosse, M.A. Popp, 3055.
J.B. Wang, T. Wendrich, and E.M. Rasel (2015), “Atom Solà, J (2013), “Cosmological constant and vacuum energy:
interferomenty on sounding rockets,” European Rocket and old and new ideas,” Journal of Physics Conference Series
Baloon: Programmes and Related research, ESA Special 453, 012015.
Publications 730, 309.
109

Sondag, A, and H. Dittus (2016), “Electrostatic positioning Swallows, M D, T. H. Loftus, W. C. Griffith, B. R. Heckel,
system for a free fall test at drop tower bremen and an E. N. Fortson, and M. V. Romalis (2013), “Techniques
overview of tests for the weak equivalence principle in past, used to search for a permanent electric dipole moment of
present and future,” Adv. Space Research 58, 644. the 199 Hg atom and the implications for CP violation,”
Srianand, R, H. Chand, P. Petitjean, and B. Aracil (2004), Phys. Rev. A 87 (1), 012102.
“Limits on the time variation of the electromagnetic fine- Talmadge, C, J.-P. Berthias, R. W. Hellings, and E. M.
structure constant in the low energy limit from absorption Standish (1988), “Model-independent constraints on pos-
lines in the spectra of distant quasars,” Phys. Rev. Lett. sible modifications of Newtonian gravity,” Phys. Rev. Lett.
92, 121302. 61, 1159.
Srianand, R, H. Chand, P. Petitjean, and B. Aracil (2007), Tamm, C, N. Huntemann, B. Lipphardt, V. Gerginov, N. Ne-
“Reply to the comment by M. T. Murphy, J. K. Webb, and mitz, M. Kazda, S. Weyers, and E. Peik (2014), “Cs-based
V. V. Flambaum,” Phys. Rev. Lett. 99, 239002. optical frequency measurement using cross-linked optical
Stadnik, Y V, and V. V. Flambaum (2014a), “Axion-induced and microwave oscillators,” Phys. Rev. A 89, 023820.
effects in atoms, molecules, and nuclei: Parity nonconserva- Tan, W-H, S.-Q. Yang, C.-G. Shao, J. Li, A.-B. Du, B.-F.
tion, anapole moments, electric dipole moments, and spin- Zhan, Q.-L. Wang, P.-S. Luo, L.-C. Tu, and J. Luo (2016),
gravity and spin-axion momentum couplings,” Phys. Rev. “New test of the gravitational inverse-square law at the sub-
D 89, 043522. millimeter range with dual modulation and compensation,”
Stadnik, Y V, and V. V. Flambaum (2014b), “Searching for Phys. Rev. Lett. 116, 131101.
topological defect dark matter via nongravitational signa- Tarallo, M G, T. Mazzoni, N. Poli, D. V. Sutyrin, X. Zhang,
tures,” Phys. Rev. Lett. 113, 151301. and G. M. Tino (2014), “Test of Einstein equivalence prin-
Stadnik, Y V, and V. V. Flambaum (2015a), “Nuclear spin- ciple for 0-spin and half-integer-spin atoms: Search for spin-
dependent interactions: searches for WIMP, axion and gravity coupling effects,” Phys. Rev. Lett. 113, 023005.
topological defect dark matter, and tests of fundamental Tarbutt, M R, B. E. Sauer, J. J. Hudson, and E. A. Hinds
symmetries,” Eur. Phys. J. C 75, 110. (2013), “Design for a fountain of YbF molecules to measure
Stadnik, Y V, and V. V. Flambaum (2015b), “Searching for the electron’s electric dipole moment,” New J. Phys. 15,
dark matter and variation of fundamental constants with 053034.
laser and maser interferometry,” Phys. Rev. Lett. 114, Tardiff, E R, E. T. Rand, G. C. Ball, T. E. Chupp, A. B.
161301. Garnsworthy, P. Garrett, M. E. Hayden, C. A. Kierans,
Stadnik, Y V, and V. V. Flambaum (2016), “Enhanced effects W. Lorenzon, M. R. Pearson, et al. (2014), “The radon
of variation of the fundamental constants in laser interfer- EDM apparatus,” Hyperfine Interact. 225, 197.
ometers and application to dark-matter detection,” Phys. Taveras, V, and N. Yunes (2008), “Barbero-immirzi param-
Rev. A 93, 063630. eter as a scalar field: K-inflation from loop quantum grav-
Stadnik, Yevgeny V, and Victor V. Flambaum (2014c), “New ity?” Phys. Rev. D 78, 064070.
atomic probes for dark matter detection: Axions, axion-like Taylor, T R (1990), “Dilaton, gaugino condensation and su-
particles and topological defects,” Mod. Phys. Lett. A 29, persymmetry breaking,” Phys. Lett. B 252, 59.
1440007. Taylor, T R, and G. Veneziano (1988), “Dilaton couplings at
Stamper-Kurn, D M, and M. Ueda (2013), “Spinor Bose large distances,” Phys. Lett. B 213, 450.
gases: Symmetries, magnetism, and quantum dynamics,” Terrano, W A, E. G. Adelberger, J. G. Lee, and B. R. Heckel
Rev. Mod. Phys. 85, 1191. (2015), “Short-range, spin-dependent interactions of elec-
Streater, R F, and A. S. Wightman (2000), PCT, Spin and trons: A probe for exotic pseudo-goldstone bosons,” Phys.
Statistics, and All That (Princeton Univ., Princeton). Rev. Lett. 115, 201801.
Sundrum, R (1999), “Towards an effective particle-string res- Testera, G, S. Aghion, C. Amsler, A. Ariga, T. Ariga,
olution of the cosmological constant problem,” J. High En- A. Belov, G. Bonomi, P. Braunig, J. Bremer, R. Brusa,
ergy Physics 1999, 1. et al. (2015), “The AEgIS experiment,” Hyperfine Inter-
Sushkov, A O, W. J. Kim, D. A. R. Dalvit, and S. K. Lamore- act. 233, 13.
aux (2011a), “New experimental limits on non-Newtonian Thorne, K S, and J. B. Hartle (1985), “Laws of motion and
forces in the micrometer range,” Phys. Rev. Lett. 107, precession for black holes and other bodies,” Phys. Rev. D
171101. 31, 1815.
Sushkov, A O, W. J. Kim, D. A. R. Dalvit, and S. K. Lamore- Tino, G M (2001), “Spectroscopic tests of spin-statistics con-
aux (2011b), “Observation of the thermal Casimir force,” nection and symmetrization postulate of quantum mechan-
Nature Phys. 7, 230. ics,” Phys. Scr. T 95, 62.
Sushkov, O P (2001), “Breit-interaction correction to the hy- Tino, G M, F. Sorrentino, D. Aguilera, B. Battelier,
perfine constant of an external s electron in a many-electron A. Bertoldi, Q. Bodart, K. Bongs, P. Bouyer, C. Braxmaier,
atom,” Phys. Rev. A 63, 42504, 0010028. L. Cacciapuoti, et al. (2013), “Precision gravity tests with
Sushkov, O P, and V V Flambaum (1978), “Parity breaking atom interferometry in space,” Nucl. Phys. B Proc. Suppl.
effects in diatomic molecules,” Sov. J. Exp. Theor. Phys. 243, 203.
48, 608. Tobar, M E, P. L. Stanwix, J. J. McFerran, J. Guéna, M. Ab-
Sushkov, O P, V. V. Flambaum, and I. B. Khriplovich (1984), grall, S. Bize, A. Clairon, P. Laurent, P. Rosenbusch,
“Possibility of investigating P- and T-odd nuclear forces D. Rovera, and G. Santarelli (2013), “Testing local po-
in atomic and molecular experiments,” Sov. Phys. JETP sition and fundamental constant invariance due to periodic
60 (November 1984), 873. gravitational and boost using long-term comparison of the
Svrcek, P, and E. Witten (2006), “Axions in string theory,” SYRTE atomic fountains and H-masers,” Phys. Rev. D 87,
J. High Energy Phys. 06, 051. 122004.
110

Tokunaga, S K, C Stoeffler, F Auguste, A Shelkovnikov, 58, 1918.


C Daussy, A Amy-Klein, C Chardonnet, and B Darquié Uzan, J-P (2011), “Varying constants, gravitation and cos-
(2013), “Probing weak force-induced parity violation by mology,” Living Reviews in Relativity 14, 2.
high-resolution mid-infrared molecular spectroscopy,” Mol. Uzan, J-P (2013), “Variation of fundamental constants on
Phys. 111, 2363. sub- and super-hubble scales: From the equivalence prin-
Truppe, S, R. J. Hendricks, S. K. Tokunaga, H. J. ciple to the multiverse,” in American Institute of Physics
Lewandowski, M. G. Kozlov, C. Henkel, E. A. Hinds, and Conference Series, American Institute of Physics Confer-
M. R. Tarbutt (2013), “A search for varying fundamental ence Series, Vol. 1514, edited by M. P. Dabrowski, A. Bal-
constants using hertz-level frequency measurements of cold cerzak, and T. Denkiewicz, p. 14.
CH molecules,” Nature Commun. 4, 2600. Uzan, Jean-Philippe (2015), “The stability of fundamental
Trzcinska, A, J. Jastrzebski, P. Lubinski, F. J. Hartmann, constants,” C. R. Physique 16, 576.
R. Schmidt, T. von Egidy, and B. Klos (2001), “Neutron Vainshtein, A I (1972), “To the problem of nonvanishing grav-
density distributions deduced from antiprotonic atoms,” itation mass,” Phys. Lett. B 39, 393.
Phys. Rev. Lett. 87, 82501. Van Tilburg, K, N. Leefer, L. Bougas, and D. Budker (2015),
Tsigutkin, K, D. Dounas-Frazer, A. Family, J. E. Stalnaker, “Search for ultralight scalar dark matter with atomic spec-
V. V. Yashchuk, and D. Budker (2009), “Observation of troscopy,” Phys. Rev. Lett. 115, 011802.
a large atomic parity violation effect in ytterbium,” Phys. van Zoest, T, N. Gaaloul, Y. Singh, H. Ahlers, W. Herr, S. T.
Rev. Lett. 103, 071601. Seidel, W. Ertmer, E. Rasel, M. Eckart, E. Kajari, et al.
Tsujikawa, S (2013), “Quintessence: a review,” Class. Quan- (2010), “Bose-Einstein condensation in microgravity,” Sci-
tum Gravity 30, 214003. ence 328, 1540.
Tullney, K, F. Allmendinger, M. Burghoff, W. Heil, Varga, K (2014), “Comment on the “Stability of the five-
S. Karpuk, W. Kilian, S. Knappe-Grüneberg, W. Müller, body bi-positronium ion Ps2 e− ” [Phys. Lett. A 372 (2008)
U. Schmidt, A. Schnabel, F. Seifert, Yu. Sobolev, and 6721],” Phys. Lett. A 378, 529.
L. Trahms (2013), “Constraints on spin-dependent short- Varga, K, J. Usukura, and Y. Suzuki (1998), “Second bound
range interaction between nucleons,” Phys. Rev. Lett. 111, state of the positronium molecule and biexcitons,” Phys.
100801. Rev. Lett. 80, 1876.
Tureanu, Anca (2013), “CPT and Lorentz invariance: Their Vasilakis, G, J. M. Brown, T. W. Kornack, and M. V.
relation and violation,” Journal of Physics Conference Se- Romalis (2009), “Limits on new long range nuclear spin-
ries 474, 012031. dependent forces set with a K-3 He comagnetometer,” Phys.
Turner, M S (1988), “Axions from SN1987A,” Phys. Rev. Rev. Lett. 103, 261801.
Lett. 60, 1797. Vasil’ev, B V (1969), Zh. Eksp. Teor. Fiz. Pis’ma Red. 9, 299.
Turner, M S (1990), “Windows on the axion,” Phys. Rep. Velyukhov, G E (1968), Zh. Eksp. Teor. Fiz. Pis’ma Red. 8,
197, 67. 372.
Tyson, J A, G. P. Kochanski, and I. P. Dell’Antonio (1998), Venema, B J, P. K. Majumder, S. K. Lamoreaux, B. R.
“Detailed mass map of CL 0024+ 1654 from strong lens- Heckel, and E. N. Fortson (1992), “Search for a coupling of
ing,” Astrophys. J. Lett. 498, L107. Earth’s gravitational field to nuclear spins in atomic mer-
Tzanavaris, P, M. T. Murphy, J. K. Webb, V. V. Flambaum, cury,” Phys. Rev. Lett. 68, 135.
and S. J. Curran (2007), “Probing variations in fundamen- Vessot, R F C, M. W. Levine, E. M. Mattison, E. L. Blomberg,
tal constants with radio and optical quasar absorption-line T. E. Hoffman, G. U. Nystrom, B. F. Farrel, R. Decher,
observations,” Mon. Not. R. Astron. Soc. 374, 634. P. B. Eby, and C. R. Baugher (1980), “Test of relativis-
Tzanavaris, P, J. K. Webb, M. T. Murphy, V. V. Flambaum, tic gravitation with a space-borne hydrogen maser,” Phys.
and S. J. Curran (2005), “Limits on variations in fundamen- Rev. Lett. 45, 2081.
tal constants from 21-cm and ultraviolet quasar absorption Vetter, P A, D. M. Meekhof, P.K. Majumder, S. K. Lamore-
lines,” Phys. Rev. Lett. 95, 041301. aux, and E. N. Fortson (1995), “Precise test of electroweak
Ubachs, W, J. Bagdonaite, E. J. Salumbides, M. T. Murphy, theory from a new measurement of parity nonconservation
and L. Kaper (2016), “Colloquium: Search for a drifting in atomic thallium,” Phys. Rev. Lett. 74, 2658.
proton-electron mass ratio from H2 ,” Rev. Mod. Phys. 88, Vietze, L, P. Klos, J. Menendez, W. C. Haxton, and
021003. A. Schwenk (2015), “Nuclear structure aspects of spin-
Ueda, Masahito (2014), “Topological aspects in spinor Bose- independent WIMP scattering off xenon,” Phys. Rev. D
Einstein condensates,” Rep. Prog. Phys. 77, 122401. 91, 043520.
Ulmer, S, C. C. Rodegheri, K. Blaum, H. Kracke, A. Mooser, Vilenkin, A (1985), “Cosmic strings and domain walls,” Phys.
W. Quint, and J. Walz (2011), “Observation of spin Rep. 121, 263.
flips with a single trapped proton,” Phys. Rev. Lett. 106, Volotka, A V, D. A. Glazov, O. V. Andreev, V. M. Shabaev,
253001. I. I. Tupitsyn, and G. Plunien (2012), “Test of many-
Ulmer, S, C. Smorra, A. Mooser, K. Franke, H. Nagahama, electron QED effects in the hyperfine splitting of heavy
G. Schneider, T. Higuchi, S. van Gorp, K. Blaum, Y. Mat- high-z ions,” Phys. Rev. Lett. 108, 073001.
suda, W. Quint, J. Walz, and Y. Yamazaki (2015), “High- Volotka, A V, D. A. Glazov, G. Plunien, and V. M. Shabaev
precision comparison of the antiproton-to-proton charge- (2013), “Progress in quantum electrodynamics theory of
to-mass ratio,” Nature (London) 524, 196. highly charged ions,” Annalen der Physik 525, 636.
Ushijima, I, M. Takamoto, M. Das, T. Ohkubo, and H. Katori Volotka, A V, D. A. Glazov, V. M. Shabaev, I. I. Tupitsyn,
(2015), “Cryogenic optical lattice clocks,” Nature Photon- and G. Plunien (2014), “Many-electron QED corrections
ics 9, 185. to the g factor of lithiumlike ions,” Phys. Rev. Lett. 112,
Usukura, J, K. Varga, and Y. Suzuki (1998), “Signature of 253004.
the existence of the positronium molecule,” Phys. Rev. A
111

Volotka, A V, and G. Plunien (2014), “Nuclear polariza- Whitmore, J B, and M. T. Murphy (2015), “Impact of instru-
tion study: New frontiers for tests of QED in heavy highly mental systematic errors on fine-structure constant mea-
charged ions,” Phys. Rev. Lett. 113, 023002. surements with quasar spectra,” Mon. Not. R. Astron. Soc.
von der Wense, L, B. Seiferle, M. Laatiaoui, J. B. Neumayr, 447, 446.
H.-J. Maier, H.-F. Wirth, C. Mokry, J. Runke, K. Eber- Wichmann, E H (2001), “Symmetries and the connection be-
hardt, C. E. Düllmann, N. G. Trautmann, and P. G. Thi- tween spin and statistics in rigorous quantum field theory,”
rolf (2016), “Direct detection of the 229 Th nuclear clock p. 201.
transition,” Nature (London) 533, 47. Wilczek, F (1978), “Problem of strong P and T invariance in
Vretenar, D, G A Lalazissis, and P Ring (2000), “Neutron the presence of instantons,” Phys. Rev. Lett. 40, 279.
density distributions for atomic parity nonconservation ex- Wilczek, F (1982), “Axions and family symmetry breaking,”
periments,” Phys. Rev. C 62, 45502. Phys. Rev. Lett. 49, 1549.
Vutha, A C, W. C. Campbell, Y. V. Gurevich, N. R. Hut- Wilczynska, M R, J. K. Webb, J. A. King, M. T. Murphy,
zler, M. Parsons, D. Patterson, E. Petrik, B. Spaun, J. M. M. B. Bainbridge, and V. V. Flambaum (2015), “A new
Doyle, G. Gabrielse, and D. DeMille (2010), “Search for analysis of fine-structure constant measurements and mod-
the electric dipole moment of the electron with thorium elling errors from quasar absorption lines,” Mon. Not. R.
monoxide,” J. Phys. B 43, 074007. Astron. Soc. 454, 3082.
Wagner, A, G. Rybka, M. Hotz, L. J. Rosenberg, S. J. Asz- Will, C M (2014), “The confrontation between general rela-
talos, G. Carosi, C. Hagmann, D. Kinion, K. Van Bibber, tivity and experiment,” Living Reviews in Relativity 17,
J. Hoskins, et al. (2010), “Search for hidden sector photons 4.
with the ADMX detector,” Phys. Rev. Lett. 105, 171801. Williams, J, S.-W. Chiow, N. Yu, and H. Müller (2016),
Wagner, A, S. Sturm, F. Köhler, D. A. Glazov, A. V. Volotka, “Quantum test of the equivalence principle and space-time
G. Plunien, W. Quint, G. Werth, V. M. Shabaev, and aboard the international space station,” New J. Phys. 18,
K. Blaum (2013), “g factor of lithiumlike silicon 28 Si11+ ,” 025018.
Phys. Rev. Lett. 110, 033003. Williams, J G, S. G. Turyshev, and D. H. Boggs (2004),
Wagner, T A, S. Schlamminger, J. H. Gundlach, and “Progress in lunar laser ranging tests of relativistic grav-
E. G. Adelberger (2012), “Torsion-balance tests of the weak ity,” Phys. Rev. Lett. 93, 261101.
equivalence principle,” Class. and Quantum Gravity 29, Williams, J G, S. G. Turyshev, and D. H. Boggs (2012),
184002. “Lunar laser ranging tests of the equivalence principle,”
Wall, T E, A. M. Alonso, B. S. Cooper, A. Deller, S. D. Class. Quantum Gravity 29, 184004.
Hogan, and D. B. Cassidy (2015), “Selective production Wilson, Kenneth G (1970), “Operator-product expansions
of rydberg-Stark states of positronium,” Phys. Rev. Lett. and anomalous dimensions in the Thirring model,” Phys.
114, 173001. Rev. D 2, 1473.
Wan-Ping, Z, Z. Peng, and Q. Hao-Xue (2015), “Detecting Windberger, A, J. R. Crespo López-Urrutia, H. Bekker, N. S.
extra dimensions by hydrogen-like atoms,” Open Physics Oreshkina, J. C. Berengut, V. Bock, A. Borschevsky, V. A.
13. Dzuba, E. Eliav, Z. Harman, et al. (2015), “Identification
Wansbeek, L W, B. K. Sahoo, R. G. E. Timmermans, of the predicted 5s − 4f level crossing optical lines with
K. Jungmann, B. P. Das, and D. Mukherjee (2008), applications to metrology and searches for the variation of
“Atomic parity nonconservation in Ra+ ,” Phys. Rev. A fundamental constants,” Phys. Rev. Lett. 114, 150801.
78 (5), 50501. Windpassinger, Patrick, and Klaus Sengstock (2013), “En-
Warrington, R B, C D Thompson, and D N Stacey (1993), “A gineering novel optical lattices,” Rep. Prog. Phys. 76,
new measurement of parity-non-conserving optical rotation 086401.
at 648 nm in atomic bismuth,” Europhys. Lett. 24, 641. Wineland, D J (2013), “Nobel lecture: Superposition, entan-
Wcislo, P, P. Morzyński, M. Bober, A. Cygan, D. Lisak, glement, and raising Schrödinger’s cat,” Rev. Mod. Phys.
R. Ciurylo, and M. Zawada (2016), “Experimental con- 85, 1103.
straint on dark matter detection with optical atomic Wineland, D J, J. J. Bollinger, D. J. Heinzen, W. M. Itano,
clocks,” Nature Astronomy 1, 0009. and M. G. Raizen (1991), “Search for anomalous spin-
Webb, J K, V. V. Flambaum, C. W. Churchill, M. J. Drinkwa- dependent forces using stored ion spectroscopy,” Phys. Rev.
ter, and J. D. Barrow (1999), “Search for time variation Lett. 67, 1735.
of the fine structure constant,” Phys. Rev. Lett. 82, 884. Wineland, D J, and N. F. Ramsey (1972), “Atomic deuterium
Webb, J K, J. A. King, M. T. Murphy, V. V. Flambaum, maser,” Phys. Rev. A 5, 821.
R. F. Carswell, and M. B. Bainbridge (2011), “Indications Wolf, P, C. J. Bordé, A. Clairon, L. Duchayne, A. Landra-
of a spatial variation of the fine structure constant,” Phys. gin, P. Lemonde, G. Santarelli, W. Ertmer, E. Rasel, F. S.
Rev. Lett. 107, 191101. Cataliotti, et al. (2009), “Quantum physics exploring grav-
Weinberg, S (1978), “A new light boson?” Phys. Rev. Lett. ity in the outer solar system: the SAGAS project,” Exper-
40, 223. imental Astronomy 23, 651.
Weinberg, Steven (1976), “Gauge theory of CP nonconserva- Wolf, P, F. Chapelet, S. Bize, and A. Clairon (2006), “Cold
tion,” Phys. Rev. Lett. 37, 657. atom clock test of Lorentz invariance in the matter sector,”
Weinberg, Steven (1989), “The cosmological constant prob- Phys. Rev. Lett. 96, 060801.
lem,” Rev. Mod. Phys. 61, 1. Wood, C S, S. C. Bennett, D. Cho, B. P. Masterson, J. L.
Weiss, David S, Fang Fang, and Jingbiao Chen (2003), “Mea- Roberts, C. E. Tanner, and C. E. Wieman (1997), “Mea-
suring the electric dipole moment of Cs and Rb in an optical surement of parity nonconservation and an anapole mo-
lattice,” in APS April Meet. Abstr., Vol. 1, p. 1008. ment in cesium,” Science 275, 1759.
Wheeler, J A (1946), “Polyelectrons,” Annals of the New York Wu, C S, E. Ambler, R. W. Hayward, D. D. Hoppes, and R. P.
Academy of Sciences 48, 219. Hudson (1957), “Experimental test of parity conservation
112

in beta decay,” Phys. Rev. 105, 1413. Vol. 43, p. 1660193.


Yamaguchi, A, M. Kolbe, H. Kaser, T. Reichel, A. Gottwald, Young, B A (1969), “Search for a gravity shift in the proton
and E. Peik (2015), “Experimental search for the low- larmor frequency,” Phys. Rev. Lett. 22, 1445.
energy nuclear transition in 229 Th with undulator radia- Yu, N, and M. Tinto (2011), “Gravitational wave detec-
tion,” New J. Phys. 17, 053053. tion with single-laser atom interferometers,” Gen. Relativ.
Yamazaki, T, A. Miyazaki, T. Suehara, T. Namba, S. Asai, Gravit. 43, 1943.
T. Kobayashi, H. Saito, I. Ogawa, T. Idehara, and Yudin, V I, A. V. Taichenachev, and A. Derevianko (2014),
S. Sabchevski (2012), “Direct observation of the hyperfine “Magnetic-dipole transitions in highly-charged ions as a ba-
transition of ground-state positronium,” Phys. Rev. Lett. sis of ultra-precise optical clocks,” Phys. Rev. Lett. 113,
108, 253401. 233003.
Yamazaki, Yasunori, and Stefan Ulmer (2013), “CPT sym- Yunes, N, K. Yagi, and F. Pretorius (2016), “Theoreti-
metry tests with cold p and antihydrogen,” Annalen der cal physics implications of the binary black-hole mergers
Physik 525, 493. GW150914 and GW151226,” Phys. Rev. D 94, 084002.
Yan, H, and W. M. Snow (2013), “New limit on possible long- Zel’dovich, Ya B (1958), “Electromagnetic interaction with
range parity-odd interactions of the neutron from neutron- parity violation,” Sov. Phys. JETP 6, 1184.
spin rotation in liquid 4 He,” Phys. Rev. Lett. 110, 082003. Zeldovich, Ya B (1959), “Parity nonconservation in the 1st or-
Yan, H, G. A. Sun, S. M. Peng, Y. Zhang, C. Fu, H. Guo, der in the weak-interaction constant in electron scattering
and B. Q. Liu (2015), “Searching for new spin- and velocity- and other effects,” Sov. Phys. JETP 9, 682.
dependent interactions by spin relaxation of polarized 3 He Zelevinsky, T, S. Kotochigova, and J. Ye (2008), “Precision
gas,” Phys. Rev. Lett. 115, 182001. test of mass-ratio variations with lattice-confined ultracold
Yan, Z-C, W. Nörtershäuser, and G. W. F. Drake (2008), molecules,” Phys. Rev. Lett. 100, 043201.
“High precision atomic theory for Li and Be+ : QED shifts Zeppenfeld, M, T. Gantner, R. Glöckner, M. Ibrügger,
and isotope shifts,” Phys. Rev. Lett. 100, 243002. M. Koller, A. Prehn, X. Wu, S. Chervenkov, and G. Rempe
Yan, Z-C, W. Nörtershäuser, and G. W. F. Drake (2009), (2017), “An experimental toolbox for the generation of cold
“Erratum: High precision atomic theory for Li and Be+ : and ultracold polar molecules,” in Journal of Physics Con-
QED shifts and isotope shifts [Phys. Rev. Lett. 100, 243002 ference Series, Journal of Physics Conference Series, Vol.
(2008)],” Phys. Rev. Lett. 102, 249903. 793, p. 012035.
Yang, C N (1950), “Selection rules for the dematerialization Zhitnitskii, A R (1980), “Weinberg’s model of CP violation
of a particle into two photons,” Phys. Rev. 77, 242. and T-odd correlations in weak decays,” Sov. J. Nucl. Phys.
Yang, S-Q, B.-F. Zhan, Q.-L. Wang, C.-G. Shao, L.-C. Tu, 31, 260.
W.-H. Tan, and J. Luo (2012), “Test of the gravitational Zhou, L, S. Long, B. Tang, X. Chen, F. Gao, W. Peng,
inverse square law at millimeter ranges,” Phys. Rev. Lett. W. Duan, J. Zhong, Z. Xiong, J. Wang, Y. Zhang, and
108, 081101. M. Zhan (2015), “Test of equivalence principle at 10−8 level
Yerokhin, V A, P. Indelicato, and V. M. Shabaev (2003), by a dual-species double-diffraction Raman atom interfer-
“Two-loop self-energy correction in high-Z hydrogenlike ometer,” Phys. Rev. Lett. 115, 013004.
ions,” Phys. Rev. Lett. 91, 073001. Zhou, L, S. Long, B. Tang, X. Chen, F. Gao, W. Peng,
Yerokhin, V A, and V. M. Shabaev (2015), “Nuclear recoil W. Duan, J. Zhong, Z. Xiong, J. Wang, Y. Zhang, and
effect in the Lamb shift of light hydrogenlike atoms,” Phys. M. Zhan (2015), “Test of equivalence principle at 10−8 level
Rev. Lett. 115, 233002. by a dual-species double-diffraction raman atom interfer-
Yost, D C, A. Matveev, A. Grinin, E. Peters, L. Maisenbacher, ometer,” Phys. Rev. Lett. 115, 013004.
A. Beyer, R. Pohl, N. Kolachevsky, K. Khabarova, T. W. Zimmer, S (2017), “Experimental search for the electric dipole
Hänsch, and Th. Udem (2016), “Spectroscopy of the hy- moment of Xe-129,” .
drogen 1s − 3s transition with chirped laser pulses,” Phys. Zwicky, F (1933), “Die rotverschieb ung von extragalaktis-
Rev. A 93, 042509. chen nebeln,” Helv. Phys. Acta 6, 110–127.
Youdin, A N, D. Krause, K. Jagannathan, L. R. Hunter, and Zwicky, F (2009), “Republication of: The redshift of extra-
S. K. Lamoreaux (1996), “Limits on spin-mass couplings galactic nebulae,” Gen. Relativ. Gravit. 41, 207.
within the axion window,” Phys. Rev. Lett. 77, 2170.
Youn, S (2016), “Axion research at CAPP/IBS,” in Inter-
national Journal of Modern Physics: Conference Series,

You might also like