0% found this document useful (0 votes)
8 views88 pages

Climbing To The Top of The Atlas 13 Tev Data: European Organisation For Nuclear Research (Cern)

This document summarizes measurements of top quark production and properties using data from the ATLAS detector at the LHC corresponding to an integrated luminosity of 140 fb−1 at a center-of-mass energy of 13 TeV. Key measurements include the top quark pair production cross section, single top production in different channels, differential cross sections, rare production processes, top quark mass and properties, and searches for flavor changing neutral currents. Many measurements reach precision of 10% or better.

Uploaded by

DainXB
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
8 views88 pages

Climbing To The Top of The Atlas 13 Tev Data: European Organisation For Nuclear Research (Cern)

This document summarizes measurements of top quark production and properties using data from the ATLAS detector at the LHC corresponding to an integrated luminosity of 140 fb−1 at a center-of-mass energy of 13 TeV. Key measurements include the top quark pair production cross section, single top production in different channels, differential cross sections, rare production processes, top quark mass and properties, and searches for flavor changing neutral currents. Many measurements reach precision of 10% or better.

Uploaded by

DainXB
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 88

EUROPEAN ORGANISATION FOR NUCLEAR RESEARCH (CERN)

Submitted to: Physics Report CERN-EP-2024-099


17th April 2024
arXiv:2404.10674v1 [hep-ex] 16 Apr 2024

Climbing to the Top of the ATLAS 13 TeV data

The ATLAS Collaboration

The large amount of data recorded with the ATLAS detector at the Large Hadron Collider,

corresponding to 140 fb−1 of 𝑝 𝑝 collisions at a centre-of-mass energy of 𝑠 = 13 TeV, has
brought our knowledge of the top quark to the next level. The measurement of the top–antitop
quark pair-production cross-section has reached a precision of 1.8% and the cross-section
was measured differentially up to several TeV in several observables including the top-quark
transverse momentum and top-quark-pair invariant mass. Single-top-quark production was
studied in all production modes. Rare production processes where the top quark is associated
with a vector boson, and four-top-quark production, have become accessible and precision
measurements of several of these processes have reached cross-section uncertainties of around
10% or smaller. Innovative measurements of the top-quark mass and properties have also
emerged, including the observation of quantum entanglement in the top-quark sector and tests
of lepton-flavour universality using top-quark decays. Searches for flavour-changing neutral
currents in the top-quark sector have been significantly improved, reaching branching-ratio
exclusion limits ranging from 10−3 to 10−5 . Many of these analyses have been used to set
limits on Wilson coefficients within the effective field theory framework.

© 2024 CERN for the benefit of the ATLAS Collaboration.


Reproduction of this article or parts of it is allowed as specified in the CC-BY-4.0 license.
Contents

1 Introduction 3

2 Event selection, statistical analysis and systematic uncertainties 5


2.1 Data samples 5
2.2 Object reconstruction and event selection 5
2.3 Statistical methods 6
2.4 Systematic uncertainties and Monte Carlo modelling 7

3 Top-quark pair production 9


3.1 Inclusive top-quark pair cross-section measurements 10
3.2 Differential top-quark pair cross-section measurements 13
3.3 Studies of 𝑏-jet production in top-quark-pair events 14

4 Single-top-quark production 15
4.1 Measurements in the 𝑡-channel 16
4.2 Measurements in the 𝑡𝑊 channel 17
4.3 Measurement in the 𝑠-channel 19

5 Associated production of top quarks 20


5.1 Top-quark production in association with a 𝑊 or 𝑍 boson 21
5.2 Top-quark production in association with a photon 24
5.3 Four-top-quark production 25

6 Top-quark mass 27
6.1 Direct top-quark mass measurements 28
6.2 Indirect top-quark mass measurements 31

7 Top-quark properties 31
7.1 Top-quark decay angular properties 32
7.2 Asymmetry measurements 35
7.3 Tests of QCD 37
7.4 Test of lepton-flavour universality 40

8 Search for flavour-changing neutral currents in the top-quark sector 41


8.1 Searches for top-quark FCNC processes involving a Higgs boson 42
8.2 Searches for top-quark FCNC processes involving neutral gauge bosons 43
8.3 Summary of FCNC process constraints 46

9 Limits on Wilson coefficients within effective field theory 46

10 Conclusion 51

2
1 Introduction

The top quark is the heaviest known elementary particle, with a mass of about 172.5 GeV. Discovered at
the Tevatron proton–antiproton collider at Fermilab in 1995 [1, 2], it was the subject of a large number of
measurements by the ATLAS [3] and CMS [4] experiments at the Large Hadron Collider (LHC) during
both the first data-taking phase called Run 1, with proton–proton (𝑝 𝑝) collisions at centre-of-mass energies

of 𝑠 = 7 and 8 TeV, and the subsequent higher-energy phase, referred to as Run 2. During Run 2, from

2015 to 2018, the ATLAS detector collected data from 𝑠 = 13 TeV 𝑝 𝑝 collisions with a total integrated
luminosity of 140 fb−1 . This dataset corresponds to about 100 million produced top-quark pairs. With
this unprecedented integrated luminosity and 𝑝 𝑝 collision energy, the ATLAS Collaboration has had the
opportunity to intensify its efforts to understand the nature of the top quark.
At hadron colliders, the top quark can be produced either via the strong interaction, as quark–antiquark pairs
(tt), or through electroweak (EW) processes, giving rise to single-top-quark events. Strong tt production
can be initiated, at leading order (LO) in QCD, by either gluon–gluon fusion processes or quark–antiquark
annihilation processes (see Figure 1), with the former dominating in 𝑝 𝑝 collisions at LHC energies. Single
top quarks are produced mostly through three modes, which are labelled according to the virtuality of the
𝑊 boson: the so-called 𝑡-channel, 𝑠-channel and 𝑡𝑊 production modes (see Figure 2). At the LHC, single
top production is subdominant relative to tt production.

g t g t q t

g t̄ g t̄ q̄ t̄

Figure 1: Representative Feynman diagrams for tt production at LO in QCD: gluon-initiated fusion in the 𝑠-channel
(left) and 𝑡-channel (middle), and quark–antiquark annihilation (right).

q t b t b W− b W−

W
W+

q̄ ′ b̄ q̄ q̄ ′ g t g t

Figure 2: Representative Feynman diagrams for single top-quark production at LO: from left to right, 𝑠-channel,
𝑡-channel, and the two diagrams for the 𝑡𝑊 channel.

According to the Standard Model (SM), the top quark decays almost exclusively into a 𝑊 boson and a
bottom quark, 𝑏. The top-quark lifetime is shorter than the typical hadronisation time scale, so it doesn’t
form a bound state with other quarks. Final-state topologies in top-quark production events are then mainly

3
determined by the 𝑊-boson decay modes, with the 𝑏-quark manifesting itself as a hadronic jet. Top-quark
pair-production events can give rise to three different types of final states. The fully hadronic final state is
characterised by the production of two 𝑏-quarks and four light-flavour quarks, coming from the decays of
the top quarks and 𝑊 bosons respectively, typically giving rise to six hadronic jets. The dilepton final state
is distinguished by two charged leptons (electrons or muons) and two undetected neutrinos produced in
association with a pair of jets originating from the 𝑏-quarks (referred to as 𝑏-jets in the following). Finally,
the semileptonic final state, often referred to as single-lepton or lepton-plus-jet, typically has four hadronic
jets (two of which are 𝑏-jets), an electron or a muon, and a neutrino. The 𝑊-boson decay can also lead to
the production of a 𝜏-lepton that is either reconstructed through a hadronic decay or enters one of the two
leptonic channels if it decays leptonically. Analogously, 𝑠- and 𝑡-channel single-top-production events can
give rise to either fully hadronic or single-lepton final states, with 𝑡𝑊-production events having dilepton
final states as well.
Besides these dominant production mechanisms, 𝑝 𝑝 collisions at the LHC can give rise to other processes
involving the production of top quarks that were mostly not observed at the Tevatron or in Run 1. The
SM predicts the production of tt pairs in association with photons, 𝑊 or 𝑍 bosons, Higgs bosons or even
another tt pair, while single top production can proceed via 𝑡𝛾, 𝑡𝑍 or 𝑡𝐻 processes, in addition to the main
modes mentioned above.
The interest in studying top-quark physics at the LHC, with the benefits brought by the Run 2 dataset’s size
and collision energy, is manifold. The top quark is the heaviest and most recently discovered quark. Its
mass is a particularly important fundamental parameter of the SM, linked to its vacuum stability [5, 6],
and the LHC is the best place to measure it precisely. Moreover, precise measurements of its couplings,
as well as its production and decay properties, are essential in order to fully establish its nature and its
role in the SM. In addition, a number of proposed theories beyond the SM (BSM theories) predict new or
modified top-quark production and decay mechanisms, resulting in altered kinematic distributions or even
significant enhancements in the rates of very rare processes, such as those mediated by a flavour-changing
neutral current (FCNC). In the SM, FCNC decays such as 𝑡 → 𝑞𝑍, 𝑡 → 𝑞𝛾 or 𝑡 → 𝑞𝐻 (𝑞 being a first- or
second-generation up-type quark: 𝑢 or 𝑐) are highly suppressed and below the experimental sensitivity.
Direct searches for BSM phenomena are described in Ref. [7].
Experimentally, the identification and study of top-quark events with the ATLAS experiment relies not only
on the reconstruction of hadronic jets, and identifying those coming from the fragmentation of 𝑏-quarks
through dedicated 𝑏-tagging algorithms, but also on the identification of electrons and muons, and the
measurement of the missing transverse momentum associated with the presence of undetected neutrinos.
For the all-hadronic final states, which take advantage of the larger hadronic branching fraction of the 𝑊
boson but are more challenging in terms of background contamination, combinations of multijet and 𝑏-jet
triggers [8–10] are used, while for channels with at least one electron or muon, the online event selection is
based on single-lepton triggers [11, 12]. High-performance reconstruction and identification algorithms for
all these physics objects are essential ingredients for maximising the precision of top-quark measurements.
In Run 2, the performance of these algorithms was significantly improved relative to Run 1, thanks to
detector upgrades as well as new identification algorithms and calibration techniques. In particular, the
addition of a new innermost detector layer, the Pixel detector’s Insertable B-Layer (IBL) [13, 14], together
with the adoption of new machine-learning (ML) techniques, allowed the 𝑏-tagging performance to be
dramatically improved (around 10% efficiency increase for 𝑏-jets at the same light-flavour-jet rejection
rate) [15, 16]. On the other hand, new calibration techniques allowed the systematic uncertainties associated
with 𝑐-jet and light-flavour-jet rejection to be reduced [17, 18]. Similarly, improvements in the jet
reconstruction algorithms and energy calibration (with about a factor of two reduction in the jet energy

4
scale’s uncertainty) [19–21] contributed to the overall gain in measurement precision, beyond that coming
from the increase in sample size due to the larger integrated luminosity and production cross-sections. At
the same time, improved trigger algorithms, allowing the single-lepton transverse momentum thresholds to
be kept at reasonable levels (below 27 GeV) despite the increase in instantaneous luminosity, as well as the
introduction of new techniques to mitigate the stronger impact from additional 𝑝 𝑝 collisions in the same or
a nearby bunch crossing (pile-up), such as the so-called jet-vertex-tagger (JVT) [22], provided the means
to cope with the increased pile-up activity in Run 2. Improved electron and muon identification [23–25]
also provided pile-up mitigation. Finally, innovative analysis techniques, often relying on modern ML
algorithms, as well as refined Monte Carlo (MC) simulation tools were gradually introduced and adopted
for the top-quark measurements and searches performed over the past years.
This report is organised as follows. Section 2 describes general experimental aspects of the ATLAS
Run 2 top-quark physics analyses, such as typical event and physics object selection criteria, statistical
analysis techniques and systematic uncertainties. The following sections are each devoted to a particular
set of measurements. Section 3 reports the tt cross-section measurements, while Section 4 describes
the single-top-quark measurements. Section 5 describes the measurements of top quarks produced in
association with a boson, as well as four-top-quark production. Section 6 discusses the top-quark mass
results and Section 7 the determination of other top-quark properties. Section 8 presents the searches
for flavour-changing neutral currents in the top-quark sector, and Section 9 presents limits on Wilson
coefficients within effective field theory. Section 10 gives the conclusions of this report.

2 Event selection, statistical analysis and systematic uncertainties

2.1 Data samples

A set of early Run 2 top-quark measurements, based on the 2015 𝑝 𝑝 collision dataset and corresponding to
an integrated luminosity of 3.2 fb−1 , were essential for validating the updated detector and software set-ups,
as well as the new Monte Carlo simulation settings for the tt process and the applicability of the lepton
and jet calibrations. At the completion of the 2016 𝑝 𝑝 data-taking, the integrated luminosity collected by
ATLAS in Run 2 had increased by a factor of ten to 36 fb−1 , allowing an extensive set of new measurements
in all the top-quark physics sectors. With the inclusion of the 2017 𝑝 𝑝 collision dataset, the integrated
luminosity reached 80 fb−1 . Finally, with the addition of the data collected in 2018, the full Run 2 dataset
reached an integrated luminosity of 140 fb−1 . The quoted integrated luminosity for measurements released
earlier than December 2022 [26] is 139 fb−1 , and was updated to 140 fb−1 in accord with the final Run 2
13 TeV 𝑝 𝑝 luminosity measurement [27]. This full dataset was used to produce refined results for most of
the measurements.

2.2 Object reconstruction and event selection

Events containing top quarks typically produce final states including high-momentum jets, charged leptons
and missing transverse momentum.
In Run 2, electrons and muons [24, 25, 28] were typically required to have a transverse momentum (𝑝 T )
exceeding a threshold between 25 and 30 GeV, and to be reconstructed within the geometrical acceptance

5
of the inner detector (i.e. with absolute pseudorapidity1 |𝜂| < 2.5). They were required to pass the
identification and isolation requirements in Refs. [24, 28] to improve the rejection of misidentified hadrons
faking their signatures and non-prompt leptons from hadron decays or photon conversions, as well as to
ensure sufficient precision in the measurement of their energy and momentum. A multivariate discriminant
was developed to further reject non-prompt leptons. This discriminant uses as input the energy deposits
and charged-particle tracks in a cone around the lepton direction, as well as lifetime variables [29].
Hadronic jets were reconstructed either from calorimeter energy-deposit clusters, referred to as topological
cell clusters [30], or from combined information from the calorimeters and the inner detector. The latter
was assembled by a dedicated particle-flow algorithm [31] used for most of the latest measurements. The
jet constituents were then clustered with the anti-𝑘 𝑡 algorithm [32, 33], using a radius parameter 𝑅 = 0.4, to
form so-called ‘small-𝑅 jets’ (or simply ‘jets’ in the following). Kinematic requirements on such small-𝑅
jets include a minimum 𝑝 T of 20 or 25 GeV and |𝜂| < 4.5 (although most analyses restrict jets to |𝜂| < 2.5).
A cut was imposed on the JVT output for jets reconstructed within the acceptance of the inner detector
and below a 𝑝 T threshold of 60 GeV, to reduce the contamination from jets not coming from the hard
interaction (pile-up jets).
In addition, analyses targeting particularly high-momentum or ‘boosted’ top quarks that decay hadronically,
with decay products that are highly collimated and thus difficult to reconstruct as separate small-𝑅 jets,
relied on the clustering of calorimeter energy deposits, or the reclustering of small-𝑅 jets [34], into
larger-radius jets, with the 𝑅 parameter typically set to 1.0. Such large-𝑅 jets were then tagged, relying on
jet substructure variables, and grooming procedures based on trimming or soft-drop were applied [35–38]
to mitigate the effect of pile-up and the underlying event.
To identify jets originating from 𝑏-quarks, dedicated 𝑏-tagging algorithms were implemented and
calibrated. Several gradually improved implementations were used by ATLAS Run 2 analyses, all relying
on ML techniques based mainly on secondary-vertex reconstruction and charged-track impact parameter
measurements [15, 16]. For each tagger, a certain number of working points were defined, characterised by
𝑏-tagging efficiencies between 85% and 60% and increasing rejection power against 𝑐-jets (between 2 and
∼40) and light-flavour jets (between 40 and ∼1000).
Missing transverse momentum in an event could indicate the production of neutrinos, which leave no signal
in the detector. It is calculated as the negative vector sum of the 𝑝 T of the reconstructed and calibrated
objects in the event [39]. This sum also includes the momenta of the tracks that are matched to the primary
vertex but are not associated with any other reconstructed objects.

2.3 Statistical methods

Many of the Run 2 ATLAS measurements used unfolding techniques to correct the detector-level observed
distributions in order to obtain results at parton or particle level. Differential cross-sections were generally
measured at particle level in fiducial phase spaces, and often extrapolated to obtain parton-level results in
the full phase space as well. Particle-level objects were defined as stable particles (those with lifetimes
1
ATLAS uses a right-handed coordinate system with its origin at the nominal interaction point (IP) in the centre of the detector
and the 𝑧-axis along the beam line. Observables labelled as transverse are projected onto the 𝑥–𝑦 plane. The 𝑥-axis points
from the IP to the centre of the LHC ring, and the 𝑦-axis points upwards. Cylindrical coordinates (𝑟, 𝜙) are used in the
transverse plane, 𝜙 being the azimuthal angle around the beam line. The pseudorapidity is defined in terms of the polar angle
𝜃 as 𝜂 = − ln tan(𝜃/2), and the rapidity is defined as 𝑦 = (1/2) [(𝐸 + 𝑝 𝑧 )/(𝐸 − 𝑝 𝑧 )]. The angular distance Δ𝑅 is defined as
√︃
Δ𝑅 ≡ (Δ𝜂) 2 + (Δ𝜙) 2 . The transverse momentum is 𝑝 T = 𝑝/cosh(𝜂).

6
longer than 30 ps) produced by the MC generators, and fiducial phase spaces were defined with selection
requirements close to those used to select events in the data analysis. In contrast, parton-level results,
relying on observables based on top-quark four-momenta available in MC simulation samples or fixed-order
theoretical calculations, were extrapolated from the selected region to the full phase space without any
experimental cuts, again by means of MC simulation. Particle-level results are thus less affected by
modelling systematic uncertainties and avoid extrapolations to unmeasured regions of phase space, while
parton-level results allow direct comparisons with the most precise predictions. Besides differential
cross-sections, some of the inclusive cross-sections were also measured in fiducial phase spaces, in order
to avoid extrapolating the result beyond the event topology and kinematic selection used in the analysis.
In terms of statistical analysis, most of the differential cross-section and top-quark property measurements
were based on the well-established regularised unfolding technique known as iterative Bayesian unfolding
(IBU) [40, 41], while a small number of more recent measurements adopted fully Bayesian unfolding
(FBU) [42] or binned profile-likelihood-based unfolding (PLU) [43]. In the last two techniques, systematic
uncertainties are encoded in the statistical model as constrained nuisance parameters [44], while in IBU the
systematic uncertainties in the unfolded distributions are evaluated by repeating the unfolding procedure
for each systematic variation. Moreover, FBU and PLU facilitate the combination of several signal regions
to extract a single differential cross-section, as well as the inclusion of control regions to constrain the main
background contributions simultaneously with the unfolding.
For most of the inclusive cross-section measurements, as well as for the extraction of parameters controlling
the shape of the signal-process distribution (such as the top-quark mass), maximum-likelihood fits were
performed. Both binned and unbinned likelihood models were used, with an increasing number of
measurements being based on binned profile-likelihood fits. Exclusion limits for BSM processes such
as those induced by a FCNC were also based on binned profile likelihoods and computed using the CLs
method [45] with the asymptotic approximation [46].

2.4 Systematic uncertainties and Monte Carlo modelling

With the available dataset being larger than in Run 1, and the consequent reduction of the statistical
uncertainty in most of the measurements, systematic uncertainties became more and more important.
In terms of instrumental systematic effects, most of the top-quark measurements and searches are particularly
sensitive to uncertainties in jet energy corrections and in the efficiencies of 𝑏-tagging algorithms. In Run 2,
the jet energy was corrected for pile-up effects and further calibrated, based on both MC simulation and
data [20]. Uncertainties in the jet energy scale and resolution were then extracted, based on these correction
procedures and on considerations of jet flavour, kinematic and generator dependence. Efficiencies and
misidentification rates for the 𝑏-tagging algorithms were calibrated in data by analysing tt and 𝑍+jet
events [17, 18, 47]. The uncertainties in such measurements were then propagated to each analysis
and decomposed into sets of uncorrelated sources of uncertainty, including uncertainties assigned to
extrapolations to inaccessible kinematic regimes in the calibrations. Depending on the specific analysis,
other uncertainties may also be relevant; these include uncertainties in the LHC luminosity, beam energy
and pile-up conditions, in lepton selection efficiencies and energy–momentum corrections, and in the
determination of the missing transverse momentum.
Apart from these instrumental systematic effects , uncertainties in the details of the MC event-generation
process, referred to as ‘modelling’ systematic uncertainties, became more and more important and were

7
subjected to deeper and deeper studies. In particular, Run 2 ATLAS physics measurements adopted a refined
approach, learning from the past data results and benefiting from improvements in MC generators [48].
During Run 2, the modelling of tt and single-top processes relied on MC generators that implement the
hard process at next-to-leading order (NLO) in QCD and are interfaced with parton-shower generators to
implement perturbative and non-perturbative fragmentation processes, as well as with the underlying-event
modelling and hadron decays. The nominal predictions were initially obtained with Powheg Box v2 [49]
interfaced with Pythia 6 [50], with the EvtGen [51] package used to better simulate the decay of heavy-
flavour hadrons. After the first set of early Run 2 results, the Powheg+Pythia 6 set-up was replaced by the
more recent Powheg+Pythia 8 [52], moving from the Perugia 2012 set of tuned parameters (tune) [53]
to the A14 tune [54], derived by the ATLAS Collaboration from a number of its Run 1 measurements

at 𝑠 = 7 TeV. To assess modelling uncertainties in these ‘NLO+PS’ set-ups, a number of alternative
predictions were considered, obtained either by varying certain internal parameters in the Powheg or
Pythia generators or by replacing one of the two with a different generator.
To account for missing higher orders in the hard-process simulation, QCD scale variations were implemented
in Powheg by scaling the renormalisation and factorisation scales up and down by a factor of two. Initially,
such scale variations were included together with variations of the Pythia internal parameters controlling
the amount of initial-state QCD radiation in the parton shower (specifically ‘Var3c’ in the A14 tune [54], and
ℎdamp ) to constitute a so-called initial-state radiation (ISR) uncertainty source. This was then refined [48] by
splitting the ISR uncertainty source into various components in order to deal with the larger dataset and the
consequent risk of artificially overconstraining various fits through insufficient flexibility of the systematic
uncertainty model. In addition, uncertainties in the amount of final-state radiation (FSR) were taken into
account by varying the effective value of the strong coupling constant, 𝛼sFSR , in the parton shower.
A separate uncertainty was then assigned to the choice of model for the parton-shower evolution and
hadronisation processes, quantified by comparing samples from the nominal set-up with those generated
after replacing Pythia with the Herwig parton shower. These two parton-shower programs implement
different radiation-emission-ordering algorithms as well as different hadronisation models. Similarly to
the case of the nominal Powheg+Pythia simulation samples, the definition of this systematic variation
was gradually refined during Run 2, by moving from the Herwig++ parton shower to models based on
Herwig 7 [55], thereby reducing this uncertainty in most measurements.
Finally, an uncertainty related to the choice of matching scheme between the NLO hard-process generator
and the parton shower, referred to as an NLO-matching-scheme systematic uncertainty, was considered.
It was typically evaluated by comparing the nominal-model samples with samples generated using
MadGraph5_aMC@NLO [56] in NLO mode (referred to simply as aMC@NLO in the following) instead
of Powheg.2 Here also, the exact recipe for the systematic variation evolved during Run 2. The comparison
between aMC@NLO+Pythia 8 and a modified Powheg+Pythia 8 set-up (suitable for a direct comparison)
was replaced by internal systematic variations within the same generator set-up, Powheg+Pythia 8 [58].
For this purpose, the parameter 𝑝 hard
T , regulating how the Pythia 8 radiation phase space is determined,
was varied along with the shape of the top-quark mass distribution, accessed by comparing the nominal
set-up with a set-up where the top-quark decay is handled by MadSpin [59].
Some measurements took advantage of the available higher-order predictions for top-quark kinematics,
employing next-to-next-to-leading-order (NNLO) fixed-order QCD calculations, possibly combined with

2
In some particular cases, the Sherpa [57] event generator was used instead.

8
NLO electroweak corrections [60], to reweight the tt MC sample. This was used either to define an
additional uncertainty or to correct the nominal prediction (see Section 3.2).
Uncertainties in parton distribution functions (PDFs) were propagated to the analyses either by considering
the envelope of the variations in the NNPDF3.0 PDF set [61], or by implementing the PDF4LHC Run 2
procedure [62], which aims to capture the differences between individual PDF parameterisations as well
as the internal variations of individual PDFs. Compared to Run 1, the adoption of these new procedures
allowed a factor-of-two reduction of the PDF uncertainty impact in many Run 2 measurements.
Uncertainties in the modelling of multiple partonic interactions (often referred to as the underlying event)
and of colour-reconnection effects in top-quark-pair events were also considered for precision measurements.
These uncertainties were computed from dedicated variations in the Pythia 8 set of tunable parameters [54]
and by testing different models for colour reconnection [63–65].
As is noted later in this report, a comprehensive set of such systematic variations, including generator
comparisons and parameter value changes that evolved over time to more detailed models, turned out to be
just sufficient to cover differences between the prediction and observed data in a wide variety of inspected
kinematic quantities.
Within the framework of modelling tt and single-top processes with NLO+PS generators, an additional
uncertainty concerns the procedure adopted to avoid double counting of diagrams between the generated tt
and 𝑡𝑊 samples. These two processes can both lead to 𝑊 𝑏𝑊 𝑏 final states and therefore interfere with each
other. In tt production, the two 𝑊 𝑏 systems are produced on the top-quark mass shell (also called a doubly
resonant process) while the 𝑡𝑊 process is singly resonant. Two approaches are usually followed to avoid
double counting between doubly and singly resonant diagrams [66]: removing NLO diagrams from the 𝑡𝑊
amplitude if they overlap with tt contributions (DR scheme) or adding subtraction terms built to cancel out
doubly resonant contributions in the 𝑡𝑊 amplitude (DS scheme). The nominal 𝑡𝑊 model adopted the DR
scheme, while alternative 𝑡𝑊 samples generated with the DS scheme were used as systematic variations.
As described in Section 4.2, dedicated measurements were performed by ATLAS to assess the merits of
the two alternative models, with the goal of reducing this significant source of systematic uncertainty.
The modelling of processes other than tt and single top production followed these recipes as closely
as possible, although different generator set-ups were used for some processes, and not all systematic
variations were always technically implementable [67, 68].
Further improvements in MC modelling are expected in the future, including possible refined parton-shower
parameter tunings based on ATLAS tt data, the implementation of generators modelling full tt production
and decay at NLO (the so-called ‘bb4l’ set-up [69]) and the employment of NNLO matrix-element generators
matched to parton showers, as is done in the MINNLOPS implementation [70], already compared with the
full Run 2 ATLAS data in the differential tt cross-section measurement currently being finalised [71].

3 Top-quark pair production

The unprecedented centre-of-mass energy and large 𝑝 𝑝 collision dataset collected in Run 2 gave ATLAS the
opportunity to exploit the full potential of the LHC as a ‘top-quark factory’. The tt production cross-section

in 𝑠 = 13 TeV 𝑝 𝑝 collisions is ∼3.5 times larger than at the highest centre-of-mass energy, 8 TeV, in
Run 1, and the Run 2 data sample contains over a hundred million tt pairs. With such a large data sample,
increasingly precise inclusive and differential tt cross-section measurements became possible, providing

9
valuable input for PDF fits, allowing precision tests of QCD predictions, and probing phase-space regions
sensitive to new physics processes.

3.1 Inclusive top-quark pair cross-section measurements

Increasingly precise predictions for the total tt cross-section have been produced in the past years, reaching
NNLO accuracy in the strong coupling constant 𝛼s [72], and including EW corrections at NLO [73–75] as
well as resummation of next-to-next-to-leading logarithmic (NNLL) soft gluon terms [76] and even adding

soft gluon corrections up to third order, resulting in approximate-N3 LO results [77]. At 𝑠 = 13 TeV, with
the top-quark mass value fixed to 172.5 GeV and using the PDF4LHC21 PDFs [78], the cross-section
obtained from the Top++ 2.0 program [79], at NNLO in QCD and including NNLL resummation, is:
+20 +35
832 −29 (scale) −35 (PDF+𝛼s ) pb,

where uncertainties from factorisation and renormalisation scale variations and the impact of uncertainties
in the PDFs (calculated using the PDF4LHC prescription [62]) and 𝛼s are quoted separately. An additional
uncertainty of ±23 pb arises from the uncertainty in the top-quark mass.3
The most precise measurements of the inclusive tt cross-section were performed by using the cleanest final
state, with an opposite-sign electron–muon pair plus at least one 𝑏-tagged jet. The measurement using the
2015 dataset [80] adopted the analysis technique inherited from the corresponding Run 1 measurements [81],
based on the so-called 𝑏-tag counting method. In order to minimise the impact of systematic uncertainties
related to hadronic-jet selection (including 𝑏-tagging), the tt cross-section was extracted simultaneously
with an effective value for the 𝑏-jet selection efficiency by taking as input the numbers of events with
exactly one and exactly two 𝑏-tagged jets. The cross-section was determined by solving a system of two
equations with two unknowns. The event selection requirements were kept as loose as possible, in order
to minimise the impact of many systematic uncertainties. Finally, by restricting the event topology to
opposite-sign 𝑒𝜇 pairs, the contribution from background processes was minimised, with 𝑍+jet, 𝑊+jet, 𝑡-
and 𝑠-channel single-top, and diboson production contributing very little, and single-top 𝑡𝑊 events being
almost the only ones contributing significantly. This first measurement, already achieving a precision of
4.4%, was updated with the larger 36 fb−1 dataset [82], lowering the uncertainty to 2.4%, and again with
the full Run 2 dataset [83], reaching an outstanding precision of 1.8%. The last improvement comes almost
entirely from the decrease in the luminosity uncertainty [27], resulting from improvements in the ATLAS
𝑝 𝑝 luminosity calibration transfer and long-term stability analyses, as well as changes to the van der Meer
calibration procedures. The measured values, and the corresponding statistical and systematic uncertainties,
are reported in Figure 3, together with the other ATLAS Run 2 inclusive tt cross-section measurements
performed in other channels. Good agreement with the theory prediction can be seen. The statistical
uncertainty was already subdominant with the 2015 dataset, despite the small acceptance for the 𝑒𝜇 channel.
Among the systematic uncertainties, the most important contributions were those from uncertainties in
the luminosity determination, LHC beam energy, and tt process modelling. In addition to the reduction
in the luminosity uncertainty, the beam energy uncertainty was also reduced significantly since the first
Run 2 measurement, already reaching a precision of 0.1% for the measurement based on the 2015+2016
dataset [84]. Among the tt process modelling systematic uncertainties, those associated with the choice of
parton-shower and hadronisation model constitute one of the most important sources of uncertainty for all
the measurements, even with the most up-to-date tuned Pythia 8 and Herwig 7 generators.

3
The uncertainty corresponds to a ∓1 GeV shift in the 𝑚 𝑡 value set in the computation.

10
ATLAS s = 13 TeV total uncertainty
statistical uncertainty
NNLO+NNLL (m t=172.5 GeV, PDF4LHC21)
luminosity ⊕ stat. unc.
scale uncertainty
scale ⊕ PDF ⊕ αS uncertainty ± stat. ± syst. ± lumi.

all-hadronic 36.1 fb -1
864.0 ± 4.3 ± 126.0 ± 18.0
JHEP 01 (2021) 033

lepton+jets 139 fb-1 820.0 ± 0.4 ± 37.0 ± 14.0


Phys. Lett. B 810 (2020) 135797

dilepton eµ 3.2 fb-1 818.0 ± 8.0 ± 29.6 ± 19.0


Phys. Lett. B 761 (2016)

dilepton eµ 36.1 fb-1 826.4 ± 3.6 ± 11.3 ± 15.7


Eur. Phys. J. C 80 (2020) 528

dilepton eµ 140 fb-1 829.0 ± 1.0 ± 13.2 ± 8.0


JHEP 07 (2023) 141

600 650 700 750 800 850 900 950 1000 1050
σtt [pb]

Figure 3: Summary of ATLAS inclusive tt cross-section measurements at 𝑠 = 13 TeV based on Run 2 data,
compared with the exact NNLO QCD calculation complemented with NNLL resummation (Top++ 2.0). The theory
band represents uncertainties due to renormalisation and factorisation scales, parton distribution functions and
the strong coupling constant. The uncertainties in the experimental measurements are broken down into their
statistical and systematic components, quoting the uncertainty related to the integrated luminosity separately. The
measurements and the theory calculation are quoted for 𝑚 𝑡 = 172.5 GeV.

These inclusive tt cross-section measurements were used as one of the inputs for high-precision determina-
tions of ratios of tt and 𝑍 production cross-sections at the three centre-of-mass energies where ATLAS

measurements are available: 𝑠 = 13, 8, 7 TeV. This was already done with the first measurement based on
√ √
3.2 fb−1 of Run 2 data [85]: single ratios, at a given 𝑠 for the two processes and at different 𝑠 values for

each process, as well as double ratios of the two processes at different 𝑠 values, were evaluated and then
compared with NNLO calculations using recent PDF sets, demonstrating significant power to constrain
both the gluon distribution function for Bjorken-𝑥 values near 0.1 and the light-quark sea for 𝑥 < 0.02.
The second tt cross-section paper, based on 36 fb−1 of Run 2 data [82], included updated computations of
these ratios and double ratios of tt and 𝑍 cross-sections at different energies. This inclusive tt cross-section
measurement was also used to extract the top-quark mass with an uncertainty of approximately 2 GeV, as
detailed in Section 6.
The inclusive tt cross-section was also measured in other final states, despite not reaching the same
precision. In particular, the total cross-section was measured in the lepton-plus-jet channel [86], by a
simultaneous profile-likelihood fit of three different binned observables in three different event categories,
characterised by different numbers of jets and 𝑏-tagged jets. An additional uncertainty, estimated as
the difference between the results obtained with the nominal MC generator and a sample reweighted to
the NNLO (QCD) + NLO (EW) parton-level prediction [60], was applied (as discussed in Section 3.2).
Moreover, an inclusive tt cross-section was measured in all-hadronic final states [87], but with significantly
larger uncertainties. These results are reported and compared with those obtained in the dilepton channel
in Figure 3. As can be seen, all the measurements are in good agreement with each other and with the

11
Inclusive tt cross-section σ t t [pb]
103
ATLAS
eµ + b-tagged jets
l l + b-tagged jets
l + jets
combined
s = 13.6 TeV, 29 fb -1
s = 13 TeV, 140 fb -1
s = 8 TeV, 20.2 fb -1
102 s = 7 TeV, 4.6 fb -1
s = 5.02 TeV, 0.26 fb -1
NNLO+NNLL (pp)
Czakon, Fiedler, Mitov, PRL 110 (2013) 252004
mt =172.5 GeV, PDF+ αS uncertainties from PDF4LHC21

4 6 8 10 12 14
Ratio wrt PDF4LHC21

1.1 PDF4LHC21+scale PDF4LHC21 QCD scales only

1.05
1
0.95
0.9

4 6 8 10 12 14
s [TeV]

Figure 4: Summary of ATLAS measurements of the top-quark pair production cross-section as a function of the
centre-of-mass energy and comparison with the NNLO QCD calculation complemented with NNLL resummation
(Top++ 2.0). The theory band represents uncertainties due to renormalisation and factorisation scales, parton
distribution functions and the strong coupling constant. The measurements and the theory calculation are quoted for
𝑚 𝑡 = 172.5 GeV. Measurements made at the same centre-of-mass energy are offset slightly for clarity. The figure
was originally published in Ref. [90].

theoretical predictions, with the precision of the experimental determination exceeding that of the QCD
NNLO+NNLL computation.
Inclusive tt cross-section measurements were performed by ATLAS with similar techniques at the different
LHC 𝑝 𝑝 centre-of-mass energies. In Run 1, measurements were performed in both the 𝑒𝜇 dilepton

channel [81] and the lepton-plus-jet channel [88, 89] at 𝑠 = 7 and 8 TeV. The inclusive tt cross-section

was also measured by ATLAS at 𝑠 = 5.02 TeV, analysing the ∼260 pb−1 of 𝑝 𝑝 data collected during
Run 2 at this reduced centre-of-mass energy in low-pile-up conditions. This measurement was obtained by

combining the lepton-plus-jet and dilepton channels [89]. Finally, at the new energy of 𝑠 = 13.6 TeV
achieved in Run 3, a measurement was made in the 𝑒𝜇 dilepton channel using the first 29 fb−1 of
collected data [90]. In Figure 4, all these measurements, together with the most precise ones performed

at 𝑠 = 13 TeV, are shown and compared with QCD NNLO+NNLL predictions as a function of the
centre-of-mass energy. Agreement between the various measurements and the theoretical predictions is
remarkable over the whole range of centre-of-mass energies.

12
3.2 Differential top-quark pair cross-section measurements

Besides the inclusive measurements, ATLAS used the Run 2 data to measure a large variety of differential tt
cross-sections in different final states and kinematic regimes. Shortly after the first inclusive cross-section

measurement at 𝑠 = 13 TeV, differential distributions were measured, with the same 3.2 fb−1 dataset, in
both the dilepton channel [91, 92] and the lepton-plus-jet channel [93, 94]. With the larger 36 fb−1 dataset,
a new set of measurements were performed, increasing the number of measured observables and improving
the precision in the dilepton [82] and lepton-plus-jet [95] topologies, as well as including all-hadronic final
states [87, 96]. Finally, the total Run 2 dataset was used to refine and update the differential cross-section
measurements in all three channels [83, 97, 98].
Measurements in the different channels can benefit from different opportunities offered by specific
topologies. The tt dilepton channel was used to achieve high-precision measurements as well as to
study properties of tt production in an environment characterised by lower hadronic activity than in the
other channels. In Ref. [91], quantities sensitive to jet activity in the tt process were measured in events
with an 𝑒𝜇 opposite-sign lepton pair and two 𝑏-tagged jets. These include the multiplicity of additional
jets, their 𝑝 T distributions, as well as the so-called ‘gap fraction’, i.e. the fraction of signal events not
containing additional jets in a given rapidity region. All these quantities are highly sensitive to details
of the parton-shower models and it is therefore important to measure them in order to validate the MC
generator predictions. The gap fraction was measured as a function of the 𝑝 T threshold for additional jets
and for different invariant-mass regions of the 𝑒𝜇𝑏 𝑏¯ system. In Ref. [92], using the same sample of tt
events, differential cross-sections were measured as a function of the transverse momentum and absolute
rapidity of the top quark, and of the transverse momentum, absolute rapidity and invariant mass of the tt
system. These measurements, unfolded to parton level, provide valuable input for testing higher-order QCD
predictions. In order to perform the measurements, the top-quark pair four-momenta were reconstructed
using the Neutrino Weighting technique [99], which uses top-quark and 𝑊-boson mass constraints to
infer the two neutrinos’ kinematics from only a measurement of the sum of their transverse momenta,
given by the event’s missing transverse momentum 𝐸 Tmiss . The most recent measurements in the dilepton
𝑒𝜇 channel [82, 83], published together with the corresponding inclusive cross-section determinations,
followed an alternative and complementary strategy. No attempt was made to reconstruct the top quarks
from their partially invisible decay products, and instead only leptonic kinematic variables were unfolded,
at particle level in a fiducial phase space, both individually and as double-differential distributions, resulting
in very clean and precise measurements. Such an approach was pursued previously in Run 1 data [100],
allowing an indirect determination of the top-quark mass. The technique used for the inclusive measurement
was applied to extract the cross-section in each of the bins.
The lepton-plus-jet and all-hadronic channels offered the advantage of easier reconstruction of the top-quark
four-momenta, thanks to the presence of at most one neutrino among the tt decay products. In addition,
they offered the opportunity to probe extreme kinematic regimes by providing a larger number of events and
the possibility of adopting ‘boosted top-tagging’ techniques to reconstruct highly collimated, hadronically
decaying top quarks. In particular, the all-hadronic channel, despite being affected by larger backgrounds,
allows full reconstruction of the kinematics of both top quarks in the event without relying on a measurement
of the missing transverse momentum. Moreover, it ensures a higher selection efficiency for high-energy
events, where both top quarks have a large boost and their decay products are collimated.4 Differential

4
Dilepton and lepton-plus-jet channel measurements typically need to identify at least one isolated electron or muon, while
leptons in boosted topologies tend to be produced at small angular separations from the 𝑏-jet coming from the same parent top
quark, resulting in non-isolated charged-lepton signatures, which are hard to identify correctly in hadronic environments.

13
cross-sections were measured in the lepton-plus-jet channel in both the ‘resolved’ [93–95] and ‘boosted’
topologies [93, 97], as well as in the all-hadronic channel for resolved [87] and boosted regimes [96, 98]. A
large number of distributions were unfolded at particle level to fiducial phase spaces, and at parton level
to the total phase space. These distributions included top-quark transverse momentum and rapidity (𝑦);
tt system invariant mass, 𝑝 T and 𝑦; and other observables related to top-quark kinematics, initial- and
final-state radiation emission and the PDFs. Double-differential distributions were also extracted, in both
the resolved and boosted topologies. Top-quark four-momenta were obtained in boosted topologies by
reconstructing large-𝑅 jets, with possible identification using top-tagging techniques. For the resolved
topology, top-quark kinematics were retrieved in different ways in the different measurements. In the
lepton-plus-jet channel, the so-called pseudo-top algorithm [101] was used; it relies on invariant-mass
constraints from the known mass of each of the two 𝑊 bosons and on angular distances between jets and
leptons, allowing the hadronically and leptonically decaying top quarks to be reconstructed with very
similar efficiency at both detector level and particle level. Alternatively, measurements performed at parton
level used the so-called KLFitter [102] package, which provides a constrained fitting algorithm that uses
transfer functions to relate the energies of the reconstructed objects to those of the parton-level objects.
In the all-hadronic channel, a 𝜒2 minimisation, relying on top-quark and 𝑊-boson mass constraints, was
used to reconstruct top quarks in the resolved topology. In both the lepton-plus-jet and dilepton channels,
cross-sections were measured with a precision of about 10%–20% for top-quark 𝑝 T up to ∼1.5 TeV and
𝑚 tt up to ∼3 TeV.
All these measurements were compared with a large set of predictions. Particle-level results were used to
test MC predictions from various generator set-ups, while parton-level ones could also be compared with
fixed-order calculations. The early measurements turned out to be very useful for testing and validating MC
generator set-ups for tt production. In particular, it became possible to rule out certain set-ups because they
gave a poor description of the data: for instance, the prediction from Powheg Box interfaced with Herwig++
gave a poor description of most of the inspected distributions, especially the additional-jet multiplicity, and
was later replaced with a prediction making use of the more recent Herwig 7 program. Measurements
based on larger datasets started to systematically reveal limitations in the modelling of certain kinematic
distributions by most of the MC generators. The modelling of the top-quark transverse momentum as well
as that of the tt system was found to be particularly poor, with most of the NLO+PS predictions, including
those from the most recent refined MC generator set-ups, overestimating the cross-section at high momenta,
especially in boosted topologies (see Figure 5 (a)). Agreement was improved when comparing parton-level
measurements with NNLO QCD fixed-order predictions or by reweighting the MC samples to match these
predictions for the top-quark kinematics, as can be seen in Figure 5 (b).

3.3 Studies of 𝒃-jet production in top-quark-pair events

In addition to inclusive and differential tt cross-section extractions, measurements of the tt production


process with extra 𝑏-jets were performed with Run 2 data. Cross-sections for tt + bb production were
measured [103] with the 36 fb−1 dataset. Inclusive and differential cross-sections were measured in the
single-lepton and dilepton channels, specifically for tt events with additional 𝑏-quark jets defined at particle
level in dedicated fiducial phase spaces. These measurements were compared with predictions from various
MC generator set-ups, and validated the background models for measurements of tt𝐻 production in the
𝐻 → bb decay channel as well as for new-physics searches in similar final states. Figure 6 (a) shows the
results of the measurement, in terms of inclusive fiducial cross-sections for the two channels, separately for
fiducial phase spaces with ≥3 𝑏-jets or ≥4 𝑏-jets. These measured inclusive cross-sections are higher than

14
[pb/GeV]
102
σ tt [TeV ]
-1

Data
ATLAS Data ATLAS
s=13 TeV, 139 fb-1 PWG+Py8 s = 13 TeV, 139 fb-1
PWG+PY8 PWG+PY8 (NNLO rw.)
MG5_aMC@NLO+Py8 10−1
PWG+H7 PWG+H7 (NNLO rw.)
T
1dσ

Boosted all-hadronic tt Boosted


dp

PWG+H7.1.3

t,h
10 Fiducial particle level

T
PWG+Py8 (more IFSR) MCatNLO+PY8 MCatNLO+PY8 (NNLO rw.)

d σtt / d p
PWG+Py8 (less IFSR) Fiducial phase-space
10−2
Stat. unc. Stat.+Syst. unc.
Stat. ⊕ Sys. Unc.
Stat. Unc.
1
10−3

10−1
10−4
Prediction

1.2
1.1 10−5
Data

1
0.9
0.8 1.3

Prediction
0.7

Data
0.6
0.5 1
0 0.2 0.4 0.6 0.8 1
0.7
ptt [TeV] 500 1000 1500 2000
T
pt,h [GeV]
T

(a) (b)

Figure 5: Two examples of ATLAS differential tt cross-section measurements. (a) Normalised particle-level tt-system
transverse momentum distribution extracted in all-hadronic boosted final states, compared with a number of NLO+PS
MC predictions [98]. (b) Particle-level transverse momentum distribution of hadronically decaying top quarks
measured in the lepton-plus-jet boosted channel, compared with MC predictions, including predictions where a
reweighting of top-quark kinematics is performed to match higher-order fixed-order calculations (see text for more
details) [97].

the predictions from the dedicated tt + 𝑏b NLO generators matched to parton-shower programs, but still
within the uncertainties. The differential distribution comparisons do not show significant mismodelling by
most of these generators, beyond the experimental uncertainties of the unfolded results.
Finally, the same dataset was used to study 𝑏-jets in tt events [104]. A number of observables sensitive to
quark fragmentation were measured in jets identified as originating from top-quark decays, thus providing
an almost completely pure 𝑏-quark jet sample. Top-quark-pair events were selected in the dilepton 𝑒𝜇
channel, and the set of charged-particle tracks associated with jets were separated into those from the
primary 𝑝 𝑝 interaction vertex and those from the displaced 𝑏-hadron decay secondary vertex, in order to
construct observables characterising the longitudinal and transverse momentum distributions of 𝑏-hadrons
within 𝑏-quark jets. These measurements complement the measurements performed at 𝑒 + 𝑒 − colliders, in
which the 𝑏-quarks originate from a colour-singlet (𝑍/𝛾 ∗ ), allowing the universality of the fragmentation
models to be tested. Figure 6 (b) shows one of the measured distributions, compared with various MC
predictions. Despite being in overall agreement with the models tuned at 𝑒 + 𝑒 − colliders, the measured
distributions are still affected by large experimental uncertainties or limited by the amount of data in several
bins, preventing the current results from constraining the MC models.

4 Single-top-quark production

In the SM, the dominant single-top-quark production process at hadron colliders occurs through the 𝑡𝑊 𝑏
vertex according to three distinct diagrams at LO in QCD (see Figure 2): the exchange of a virtual 𝑊 boson
in the 𝑡-channel, a 𝑊-boson exchange in the 𝑠-channel, and the associated production of a top quark and a 𝑊
boson (named 𝑡𝑊). In 𝑝 𝑝 collisions at LHC energies, the 𝑡-channel process is dominant. The cross-section

15
1/σ dσ/dzLch,b
3.5 b
Data
Pow+Py8 A14 rB = 0.855 αFSR
S = 0.127
3 Pow+Her 7.1.3 b
b

2.5 Sherpa 2.2.10

2
b

ATLAS 1.5 ATLAS


−1
lepton+jets ( 3b) s = 13 TeV, 36 fb

b
s = 13 TeV, 36.1 fb 1
1
b

0.5 b
b
0
lepton+jets ( 4b)
1.4

Data - ttX(X = H, V) 1.2

Ratio to data
e ( 3b) Stat. uncert.
Total uncert. 1.0
Sherpa 2.2 ttbb (4FS)
Powheg+Pythia8 ttbb (4FS)
0.8
e ( 4b) PowHel+Pythia8 ttbb (5FS)
PowHel+Pythia8 ttbb (4FS)
0.6
1 2 3 4 0.5 1.0 1.5
10 10 10 10 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
fid [fb] Pred./(Data - ttX)
zLch,b

(a) (b)

Figure 6: (a) Measured fiducial tt + 𝑏-jet cross-sections [103], with tt𝐻 and tt-plus-vector-boson contributions
subtracted from the data, compared with predictions from different generator set-ups. (b) Particle-level differential
cross-section as a function of 𝑧 chL,𝑏 [104], compared with predictions from different generator set-ups. The quantity
𝑧 ch
L,𝑏 is defined as the fraction of the total charged-particle momentum in a jet carried by the 𝑏-hadron decay products
along the direction of the jet.

for each of the three single-top-quark production channels is sensitive to the coupling between the 𝑊 boson
and the top quark at the 𝑊𝑡𝑏 vertex. Single top-quark production therefore presents an opportunity to test
the structure of this coupling in the SM, as well as to probe classes of new-physics models that can affect
the 𝑊𝑡𝑏 vertex. The different single-top production modes are sensitive to different BSM models, so it is
important to study them separately [105, 106].

4.1 Measurements in the 𝒕-channel

In the 𝑡-channel, a light-flavour valence quark from one of the colliding protons interacts with a 𝑏-quark
which can originate from the proton sea (in the so-called five-flavour scheme, where all the quark flavours
except the top are included in the proton PDFs) or from gluon splitting (in the five-flavour scheme and also
in the four-flavour scheme, where only the four lightest quarks are included in the proton PDFs). After the
exchange of a space-like virtual 𝑊 boson, the produced top quark or antiquark recoils against a light-flavour
quark, referred to as the spectator quark. This quark is preferentially emitted in the forward direction.
Since the density of valence 𝑢-quarks in the proton is about twice as large as that of valence 𝑑-quarks,
the single-top-quark production cross-section, 𝜎(𝑡𝑞), is about twice as large as the single-top-antiquark

production cross-section, 𝜎( 𝑡¯𝑞). For 𝑝 𝑝 collisions at 𝑠 = 13 TeV, the predicted 𝑡-channel production
cross-sections (for 𝑚 𝑡 = 172.5 GeV) are 𝜎(𝑡𝑞) = 134.2 ± 2.2 pb and 𝜎( 𝑡¯𝑞) = 80.0 ± 1.6 pb, computed at
NNLO in perturbative QCD with the MCFM program [107]. This corresponds to an increase by around

a factor of 2.5 compared to the SM 𝑡-channel cross-sections at 𝑠 = 8 TeV. Measurements of 𝑡-channel
production were made first with 3.2 fb−1 of Run 2 data [108] and later with the full dataset [109]. Separate
measurements of top-quark or top-antiquark production were conducted because they provide sensitivity

16
to different PDFs (the 𝑢-quark and 𝑑-quark PDFs). Measurements of their ratio, 𝑅𝑡 = 𝜎(𝑡𝑞)/𝜎( 𝑡¯𝑞),
also profit from systematic uncertainties partially cancelling out, allowing even higher sensitivity to the
PDFs. Events were required to have a single isolated electron or muon from the leptonic decay of the
top quark or antiquark, and exactly two jets, among which exactly one is 𝑏-tagged. The 𝑡-channel signal
was separated from the background by using a neural network (NN). A binned maximum-likelihood
fit to the NN discriminant distribution in the channel with positively or negatively charged leptons was
performed to extract the top-quark or top-antiquark inclusive cross-section, respectively. The fits yield
+6 +0.036
𝜎(𝑡𝑞) = 137 ± 8 pb, 𝜎( 𝑡¯𝑞) = 84−5 pb and 𝑅𝑡 = 1.636−0.034 , in agreement with SM predictions.
Because of the vector minus axial-vector (𝑉 − 𝐴) structure of the 𝑊𝑡𝑏 vertex, the single top quarks are
highly polarised along the direction of the momentum of the spectator quark (or opposite to it in the case of
single top antiquark production). ATLAS used the full Run 2 dataset and 𝑡-channel single top production
to probe the polarisation of the top quark and top antiquark [110]. The event selection was similar to that
used to measure the 𝑡-channel production cross-section. Among the two required jets, the non-𝑏-tagged
jet was assumed to originate from the spectator quark. The charged lepton from the top-quark decay is
the most sensitive probe of the top-quark spin, so the angular distributions of the charged leptons were
used to extract the components of the polarisation vectors. The polarisation vector was expressed in three
orthogonal directions, where the 𝑧 ′ direction was chosen to be the momentum direction of the spectator
quark [111]. The 𝑊 boson from the top-quark decay was reconstructed from the lepton kinematics and
the reconstructed missing transverse momentum, imposing a 𝑊-boson mass constraint. The top-quark
candidate was then reconstructed by combining the four-momentum of the 𝑊 boson with that of the
𝑏-tagged jet. Finally, the charged lepton’s momentum was boosted into the top-quark rest frame to define
its polar angles in the polarisation coordinate system. The differential distributions of these polar angles
could then be used to extract the three polarisation components 𝑃 𝑥 ′ , 𝑃 𝑦 ′ and 𝑃 𝑧 ′ of the top-quark and
top-antiquark polarisation vectors, using a template fit to these distributions in the top-quark rest frame.
The results demonstrate a very high degree of polarisation in 𝑡-channel production, along the direction
of the spectator quark (for top-quark events), or opposite to that direction (for top-antiquark events), in
agreement with NNLO QCD predictions. The polar-angle differential distributions were also unfolded to
particle level in a fiducial region. The normalised unfolded distributions show good agreement with the
SM prediction, with a 𝑝-value close to 1 (see Figure 7 (a)), and were used to derive competitive bounds on
anomalous 𝑡𝑊 couplings (see Section 9).

4.2 Measurements in the 𝒕𝑾 channel

At the LHC, the second largest single-top-quark production cross-section is the one for single top production
in association with a 𝑊 boson, i.e. 𝑡𝑊, which accounts for approximately 24% of the total single-top-quark

production rate at 𝑠 = 13 TeV. This process was beyond the reach of the Tevatron, and was first observed

in 𝑠 = 8 TeV data at the LHC [113, 114], with cross-section measurements in good agreement with

theoretical predictions. The expected 𝑠 = 13 TeV SM 𝑡𝑊 cross-section at NLO including next-to-next-
+1.9
to-leading logarithms is 𝜎(𝑡𝑊) = 79.3 −1.8 (scale) ± 2.2 (PDF+𝛼s ) pb [115]. ATLAS measured this
process using 3.2 fb−1 of data collected in 2015 [116] in the final state with exactly two oppositely charged
leptons and at least one 𝑏-tagged jet, using the same boosted decision tree (BDT) technique as in the

𝑠 = 8 TeV measurement. Accurately estimating the kinematic distributions of the 𝑡𝑊 process is difficult
since this process is not well-defined at higher order in QCD because of interference with tt production
(see Section 2.4). The difference between results obtained with the DR and DS treatments is then taken as

an uncertainty and is sizeable in many analyses. The 𝑠 = 13 TeV dataset made it possible to perform

17
Px' 1.5

σ d mminimax [1/GeV]
ATLAS Data, stat. uncertainty
s=13 TeV, 139 fb-1 10−2 Full uncertainty
+ -
1 Powheg+Pythia8 l ν l ν bb
Powheg+Pythia8 tt +tW (DR)
Powheg+Pythia8 tt +tW (DS)


10−3

bl
0.5 MG5_aMC+Pythia8 tt +tW (DR)
MG5_aMC+Pythia8 tt +tW (DR2)
top quark

1
0
10−4
top antiquark
−0.5 ATLAS
−5 s=13 TeV, 36.1 fb-1
10 +-
pp → l l bb+X
best Fit
−1 68% CL stat. only
68% CL stat.+syst. 2

Model/Data
NNLO SM Prediction
−1.5 1
−1.5 −1 −0.5 0 0.5 1 1.5
Pz' 0
0 100 200 300 400
mminimax
bl [GeV]

(a) (b)

Figure 7: (a) Observed best-fit polarisation measurements with their statistical (green) and statistical+systematic
(yellow) uncertainty contours at 68% CL, plotted on the two-dimensional polarisation parameter space (𝑃 𝑧 ′ , 𝑃 𝑥 ′ ) [110].
(b) Unfolded normalized differential 𝑚 minimax
𝑏ℓ cross-section compared with various theoretical models [112].

measurements that test this modelling, and ATLAS exploited this in two ways: measuring the differential
𝑡𝑊 cross-section distributions [117] and using specific variables sensitive to the interference [112].
The differential cross-section was measured [117] in the dilepton final state, using 36 fb−1 of data and
requiring the presence of exactly one 𝑏-tagged jet and no additional jets. A BDT was constructed to
separate the 𝑡𝑊 signal from the large tt background. The regions with only untagged jets or more 𝑏-tagged
jets were used as validation regions. In this channel, the top quark or the 𝑊 boson cannot be reconstructed
directly because of the undetected neutrinos. However, some observables are correlated with the kinematic
properties of the 𝑡𝑊 process and its modelling. The unfolding to particle level was performed within a
fiducial phase space and the obtained distributions were normalised. The largest uncertainties come from
the limited size of the data sample as well as tt and 𝑡𝑊 MC modelling. In general, most of the MC models
show fair agreement with the measured cross-section distributions, although the predicted distributions are
softer than the observed ones. Both the statistical and systematic uncertainties have a significant impact on
the result, with total uncertainties ranging from 10% to 50% depending on the bins. The differential 𝑡𝑊
cross-section measurement is therefore expected to improve significantly as more data is used.
Interference between 𝑡𝑊 and tt production was specifically probed in the dilepton final state using 36 fb−1
of data in a phase space with exactly two 𝑏-tagged jets (and no additional jets passing a looser 𝑏-tagging
requirement) [112]. The contributions from doubly and singly resonant amplitudes depend on the invariant
mass of the 𝑏𝑊 pairs in the event. Since the charged-lepton kinematics are correlated with those of the
𝑊 boson, the invariant mass of the 𝑏-jet and the charged lepton is an interesting observable for testing
the interference. As there are some ambiguities in forming this mass, the differential cross-section was
measured as a function of 𝑚 minimax
𝑏ℓ = min{max(𝑚 𝑏1 ℓ1 , 𝑚 𝑏2 ℓ2 ), max(𝑚 𝑏1 ℓ2 , 𝑚 𝑏2 ℓ1 )}, an observable inspired
by Refs. [118, 119], with 𝑏 𝑖 and ℓ𝑖 being√︃the two 𝑏-tagged jets and leptons respectively. The doubly
resonant contribution is suppressed above 𝑚 2𝑡 − 𝑚 𝑊 2
(where 𝑚 𝑡 and 𝑚 𝑊 are the top-quark and 𝑊-boson
masses), so the differential cross-section above this kinematic endpoint has more sensitivity to interference

18
effects. The 𝑚 minimax
𝑏ℓ distribution was unfolded to particle level using an iterative Bayesian unfolding and
the normalised unfolded distribution was compared with the predictions (see Figure 7 (b)). The modelling
systematic uncertainties, which impact the results the most, range from 1% to 22% of the unfolded yields,
while the statistical uncertainty is as large as 20%. The predictions using the DR scheme give a better
description of the normalisation of the region 𝑚 minimax
𝑏ℓ ≥ 𝑚 𝑡 but the DS scheme models the 𝑚 minimax
𝑏ℓ shape
better in the same region. In general, the DR and DS predictions bracket the data in the region of large
𝑚 minimax
𝑏ℓ , justifying the application of their difference as a systematic uncertainty. Later studies showed that
this difference could be narrowed by using dynamic scales. The full ℓ + 𝜈ℓ − 𝜈𝑏 ¯ 𝑏¯ predictions implemented in
Powheg Box Res [120, 121], which includes off-shell top-quark effects at NLO and the interference term,
gives the best predictions of 𝑚 minimax
𝑏ℓ over the full distribution. It is expected to eventually replace the
model where the tt and 𝑡𝑊 production processes are generated separately.

4.3 Measurement in the 𝒔-channel

Among the three single-top-quark production channels mediated by 𝑡𝑊 vertices, the weakest one at the
LHC is the 𝑠-channel, in which a top quark is produced with a bottom antiquark via an 𝑠-channel 𝑊-boson
exchange. This process contributed a larger fraction of the single-top events at the Tevatron, where it was
observed in proton–antiproton collisions by the CDF and D0 collaborations [122]. At the LHC, ATLAS

found evidence for this process in 𝑝 𝑝 collisions at 𝑠 = 8 TeV with an observed (expected) significance of

3.2 (3.9) standard deviations relative to the background-only hypothesis [123]. Between 𝑠 = 8 TeV and
13 TeV, the ratio of the 𝑠-channel single-top cross-section to the dominant tt background’s cross-section

decreases from 2.1% to 1.2%, making the analysis more challenging at 𝑠 = 13 TeV. In the SM, the
√ +0.40
𝑠-channel single-top production cross-section in 𝑝 𝑝 collisions at 𝑠 = 13 TeV is 𝜎𝑠-channel = 10.32 −0.36 pb,
calculated at NLO in QCD with Hathor 2.1 [124].

Despite these challenges, ATLAS measured this cross-section at 𝑠 = 13 TeV in the final state with one
lepton and exactly two 𝑏-tagged jets [125]. To extract the signal from the large background composed
of tt, 𝑊+jets and 𝑡-channel single-top events, a discriminant based on the matrix-element method [126,
127] was used. This discriminant was built from the likelihood values computed for the hypothesis
that a measured event came from a given process. The likelihood values were computed by integrating
the matrix elements for the signal or background processes. The 𝑠-channel single-top production cross-
section was measured from a binned profile-likelihood fit of this discriminant, which also allowed
the normalisation of both the tt and 𝑊+jets background processes to vary freely. The analysis yields
+3.4
𝜎𝑠-channel = 8.2 ± 0.6 (stat.) −2.8 (syst.) pb in agreement with the SM prediction. The largest systematic
uncertainty comes from the tt normalisation, followed by those from the jet energy scale and signal
modelling. This corresponds to an observed (expected) significance of 3.3 (3.9) standard deviations for
𝑠-channel production relative to the background-only hypothesis. The significance is similar to that in the

𝑠 = 8 TeV analysis, despite the larger data sample, because of the lower signal-to-background ratio at

𝑠 = 13 TeV.

The 𝑠 = 13 TeV inclusive single-top cross-section measurements are summarised in Figure 8. The
𝑡-channel cross-section measurement was used to determine that the coupling at the 𝑊𝑡𝑏 vertex is
𝑓LV · |𝑉𝑡 𝑏 | = 1.015 ± 0.031 [109], where 𝑉𝑡 𝑏 is the corresponding element of the Cabibbo–Kobayashi–
Maskawa matrix [128, 129], and 𝑓LV is a possible additional left-handed form factor [130] (in the SM,
𝑓LV = 1). Bounds were also placed on possible anomalous couplings within the framework of effective
field theory (see Section 9).

19
ATLAS
statistical uncertainty
s=13 TeV
stat. ⊕ syst. uncertainty
theory prediction
theory uncertainty

± stat. ± syst.
t-channel [140 fb-1] 221 ± 1 ± 13
arXiv:2403.02126

tW 94.0 ± 10.0 ± 25.5


-1
[3.2 fb ]
JHEP 01 (2018) 63

s-channel -1
[139 fb ]
JHEP 06 (2023) 191
8.2 ± 0.6 ± 3.1

1 10 102 103
Single top cross-section (pb)

Figure 8: Summary of the 𝑠 = 13 TeV single-top-quark cross-section measurements, compared with the
corresponding theoretical expectations. The values are quoted for 𝑚 𝑡 = 172.5 GeV. The theory predictions are taken
from Refs. [107] (𝑡-channel), [131] (𝑡𝑊) and [124] (𝑠-channel).

5 Associated production of top quarks

The high centre-of-mass energy and large Run 2 data sample have brought the study of rare top-quark
production processes to a new level. Such processes include the production of a top quark or tt pair in
association with a 𝑍, 𝑊 or 𝛾 (the process where a tt pair is produced with a Higgs boson, tt𝐻, is described
in Ref. [132]). The cross-section of each of these rare associated-production processes (‘top+𝑉’ processes,

where 𝑉 is a vector boson) is significantly larger in Run 2 than at 𝑠 = 8 TeV, where they were first
explored. These cross-sections can be related to the coupling of the top quark to the associated boson,
allowing the measurement of the top quark’s neutral-current coupling or the top quark’s Yukawa coupling.
The top+𝑉 production cross-sections could be altered by physics beyond the SM, such as a vector-like
quark [133, 134], a strongly coupled Higgs boson [135], or a heavy scalar or pseudoscalar boson [136–138].
Even if such new physics is beyond the energy reach of the LHC for direct observation, it can manifest
itself by changing the top+𝑉 cross-sections through virtual contributions involving new particles. Their
effect can be parameterised in the context of an effective field theory (EFT) framework as dimension-six

operators that extend the SM Lagrangian [139, 140] (see Section 9). At 𝑠 = 13 TeV, the simultaneous
production of four top quarks has a cross-section large enough to make its observation feasible.
As in the case of tt production, the experimental challenges differ depending on the decays of the 𝑊
bosons coming from the top quarks. The ‘golden channel’ used to explore many of the top+𝑉 processes
considers two same-sign (SS) leptons, three leptons or even four leptons. These typically occur when at
least one top quark decays leptonically and another lepton comes from the associated boson’s decay. This
so-called multilepton channel benefits from low backgrounds from SM processes, which compensates for
the small branching fraction. However, it has rare backgrounds that are challenging to evaluate, coming
from non-prompt leptons produced in hadron decays or jets misidentified as leptons (collectively called

20
‘fake leptons’) and also from prompt electrons that have a misassigned charge (for the channel with SS
leptons). The muon charge-misassignment rate is negligible.
The fake-lepton background in the multilepton channel is usually evaluated using data-driven techniques.
A common procedure is to use the ‘matrix method’ [141]. It uses two types of events: events with ‘loose’
selected leptons and events with ‘tight’ selected leptons, where loose leptons are defined by loosening
or inverting some of the selection criteria so as to increase the fraction of fake leptons. This method is
based on the fact that the numbers of events selected by using either the loose or tight lepton criteria can be
expressed as linear combinations of the numbers of events with either prompt or fake leptons, using the
fraction of prompt loose leptons meeting the tight criteria (tight lepton efficiency) and the fraction of fake
loose leptons also meeting the tight criteria (lepton fake rate). Knowing this efficiency and fake rate (as well
as the numbers of selected loose and tight leptons), the numbers of prompt and fake leptons can be extracted
by inverting these relations. The efficiency and fake rate are measured in data using control regions
enriched in either prompt or fake leptons. When using this method, the key step is to select control regions
that are kinematically representative of the signal region so that the measured efficiency and fake rate can
be applied in the signal region. Another way to estimate the non-prompt-lepton background is the ‘template
method’ [142]. This method relies on the simulation to model the kinematic distributions of fake-lepton
background processes and on control regions enriched in fake leptons to determine their normalisations.
There are usually several control regions, enriched in fake leptons from different sources (such as electrons
from photon conversions or leptons from heavy-flavour hadron decays). These control regions are then
included, together with the signal region, in the fit that extracts the signal, and normalisation factors for the
fake-lepton backgrounds are determined simultaneously with the signal strength (defined as the ratio of the
measured cross-section to the SM prediction).
The other background usually evaluated from data is the one where the reconstruction assigns the wrong
charge sign to an electron in the case where events with two SS leptons are selected. This happens when
the electron undergoes a hard bremsstrahlung followed by an asymmetric photon conversion or when the
sign of the electron track’s curvature is mismeasured. To suppress this background, another multivariate
BDT was used, taking as input the track and energy-cluster properties of the electron candidate [29]. The
charge misassignment rate was measured from the fraction of reconstructed 𝑍 → 𝑒𝑒 data events with a
same-charge electron pair. It was parameterised as a function of the electron 𝑝 T and 𝜂 and then applied to
data events satisfying the signal selection where two leptons with opposite charges are required.

5.1 Top-quark production in association with a 𝑾 or 𝒁 boson

Studying the ttZ process provides a direct probe of the weak couplings of the top quark. The coupling of the
top quark to the 𝑍 boson is not yet well constrained, leaving room for potential new-physics contributions.
Top-quark pair production in association with a 𝑊 boson is an irreducible source of SS dilepton pairs. It is
charge asymmetric in 𝑝 𝑝 collisions because it is initiated by quark–antiquark initial states, and is unusually
complex to predict because of the importance of higher-order QCD and EW corrections. Although the ttZ
and ttW processes were not observed individually in Run 1, the tt + 𝑉 production process (with 𝑉 = 𝑍 or
√ √
𝑊) was observed at 𝑠 = 8 TeV [143], reaching a precision of ∼30%. At 𝑠 = 13 TeV, the expected SM
ttZ (ttW) production cross-section increases by a factor of more than 3 (2). The Run 2 data sample made
precise measurements possible, and also opened the door to differential measurements exploring ttZ and
ttW kinematic modelling for the first time. ATLAS performed measurements in the multilepton channel

coherently for both the ttZ and ttW signals, first using 3.2 fb−1 of 𝑠 = 13 TeV data [144] and then using
36 fb−1 [145]. The full Run 2 dataset was later used to study these processes differentially [146, 147].

21

Using 36 fb−1 of 𝑠 = 13 TeV data, the ttZ and ttW cross-sections were measured simultaneously [145]
using SS dilepton events, trilepton events (3L), opposite-sign (OS) dilepton events and four-lepton events
(4L). The OS dilepton region targets ttZ events where both top quarks decay hadronically and the 𝑍
boson decays into a pair of leptons. It suffers from a large background of 𝑍+jets and tt events. BDTs
were used to separate signal from background. The SS dilepton region targets the ttW process, with one
top quark and the 𝑊 boson decaying leptonically. It was split according to the charge of the selected
lepton pairs since ttW events are preferentially produced with a positively charged 𝑊 boson. The 3L
channel is sensitive to both ttZ and ttW events. The 4L channel targets ttZ events where both the tt pair
and the 𝑍 boson decay leptonically. The ttZ and ttW signal strengths were extracted simultaneously
using a binned maximum-likelihood fit to all the control and signal regions. Since they are important
backgrounds, the normalisations of the 𝑊 𝑍, 𝑍 𝑍 and 𝑍+heavy-flavour-jets processes are determined
from data control regions. The results are 𝜎ttZ = 950 ± 80 (stat.) ± 100 (syst.) = 950 ± 130 fb and
+78
𝜎ttW = 870 ± 130 (stat.) ± 140 (syst.) = 870 ± 190 fb in agreement with SM predictions of 𝜎ttZ = 863 −89 fb
from NLO QCD including EW corrections and NNLL resummation [148] and 𝜎ttW = 745.3 ± 54.9 fb [149]
from NNLO QCD with NLO EW corrections.
With the full Run 2 dataset, differential ttZ [147, 150] and ttW [146] measurements were carried out in
independent analyses. The latest ttZ measurement was performed in three final states: the dilepton, 3L
and 4L channels. The 𝑍+jets background was estimated using simulation. However, since the modelling
of 𝑍+jets with heavy-flavour jets is challenging, the normalisations of the 𝑍 + 𝑏 and 𝑍 + 𝑐 components
were obtained in data, simultaneously with the extraction of the signal strength. The tt background was
estimated using a fully data-driven method, relying on the high tt purity of a sample requiring one electron
and one muon. A deep neutral network (DNN) was trained in each of the regions to extract the ttZ signal.
Some of the input variables for this DNN were built using the output of the tt-system reconstruction. In all
channels, the background from fake/non-prompt leptons was estimated using the template method. The
inclusive signal strength was extracted by fitting the DNN output distributions simultaneously in all three
channels using a profile-likelihood technique. The cross-section was measured to be:

𝜎ttZ = 860 ± 40 (stat.) ± 40 (syst.) = 860 ± 60 fb,

where the largest systematic uncertainty comes from background normalisations. In addition, normalised
and absolute differential distributions sensitive to the ttZ vertex (and hence interesting for constraining some
EFT operators) were extracted in the 3L and 4L channels. The dilepton channel was not used because of its
large background contamination. To correct for acceptance and detector effects, the differential distributions
were unfolded using a profile-likelihood unfolding technique. The differential observables include the
transverse momentum and absolute rapidity of the 𝑍 boson (see Figure 9 (a)), the transverse momentum
or invariant mass of the top or tt system, and the azimuthal angle Δ𝜙 or rapidity difference between the
𝑍 boson and the leptonically decaying top quark. Most observables were measured at both parton and
particle level. The compatibility of the unfolded measurements with various predictions was assessed by
computing a 𝜒2 per degree of freedom and its corresponding 𝑝-value. In all cases, the 𝑝-values indicate
good agreement between the unfolded data and the predictions. The particle-level distributions were used
to constrain EFT effects in ttZ production (see Section 9). Furthermore, detector-level observables sensitive
to polarisation and spin correlation of the top quarks were combined to explore spin correlations in ttZ
production. Simulated templates of ttZ events with and without spin correlation were used to extract a
ratio of the measured spin correlation to the SM prediction in ttZ events of 𝑓SM = 1.20 ± 0.68. The total
uncertainty is dominated mainly by its statistical component. This result is in agreement with the SM, and
represents a 1.8𝜎 departure from a scenario without spin correlations.

22
The full Run 2 dataset was also used to explore the ttW process differentially for the first time, studying the
kinematics of ttW final-state particles and of any associated jets [146] in the channel with SS dilepton
events and the 3L channel. Beyond the interest in measuring this rare process more precisely, a better
understanding of ttW production is important since indirect measurements in analyses targeting tt𝐻 or tttt
production have consistently observed larger ttW yields than the SM predicts in the ttW phase space with
additional jets. The signal regions were split by lepton charge. An inclusive cross-section was extracted by
further splitting them by lepton flavour and both the jet and 𝑏-tagged-jet multiplicities. Control regions
were defined in order to adjust the normalisation of the fake-lepton backgrounds, as well as the diboson and
ttZ backgrounds. These control regions were defined to be orthogonal to the signal regions by applying
looser lepton isolation criteria or requiring a different number of jets or 𝑏-tagged jets. The normalisations
obtained from the fit are compatible with unity. The inclusive cross-section was measured to be:
𝜎ttW = 880 ± 50 (stat.) ± 70 (syst.) = 880 ± 80 fb,

with the largest systematic uncertainty coming from the modelling of the ttW signal. This result is
higher than the SM prediction in Ref. [149], but compatible with it at the level of 1.4 standard deviations.
Separate ttW + and ttW − cross-sections were also measured, together with their ratio and the relative
charge asymmetry. Differential ttW measurements were performed as a function of observables where
discrepancies were observed previously or that are sensitive to NLO corrections. These include the number
of jets (see Figure 9 (b)), the scalar sum of the transverse momenta of jets, and separately of leptons, and
the azimuthal angle and rapidity difference between the two same-sign leptons. The corrections to particle
level were obtained using a profile-likelihood unfolding. As expected from the inclusive case, the absolute
differential measurements exceed the theoretical predictions. The normalised distributions, however, show
rather good 𝜒2 compatibility with the MC generator predictions. The total uncertainty, dominated by
data statistics, does not currently allow the modelling performed in the MC simulation to be constrained
significantly.
The process with a single 𝑍 boson and a top quark in the final state (𝑡𝑍𝑞) is another way to probe the
coupling of the top quark to the 𝑍 boson. Despite an expected cross-section ten times smaller than that
for ttZ, this process probes two electroweak couplings in a single process: the 𝑡–𝑍 and 𝑊–𝑍 couplings.
Evidence for 𝑡𝑍𝑞 production was seen with the 36 fb−1 data sample [151], while the full Run 2 dataset
allowed a definitive observation [152] as described in the following. The expected 𝑡𝑍𝑞 SM cross-section at
+5
NLO in QCD (with a dilepton invariant mass 𝑚 ℓℓ > 30 GeV) is 102 −2 fb, computed using MCFM [153].
The 𝑡𝑍𝑞 process was searched for in the 3L channel, requiring one 𝑏-tagged jet and one additional
non-𝑏-tagged jet, which is expected to be emitted preferentially at high absolute pseudorapidity. A third
jet coming from radiation was also allowed. To help separate the signal from the diboson, 𝑍+jets and
tt backgrounds, both the 𝑍-boson and the top-quark invariant masses were reconstructed. Diboson, ttZ
and tt control regions were defined in order to adjust the background normalisations or to help constrain
their uncertainties. The contribution from non-prompt-lepton background was estimated by replacing
one 𝑏-tagged jet by a lepton in the tt and 𝑍+jets MC event samples. The signal was separated from the
background by using a neural network where the most discriminating input variable is the largest invariant
mass formed by the 𝑏-tagged jet and one of the untagged jets. The signal strength was extracted from a
maximum-likelihood fit together with the normalisations of the tt and 𝑍+jets backgrounds and leads to:
𝜎𝑡 𝑍𝑞 = 97 ± 13 (stat.) ± 7 (syst.) = 97 ± 15 fb,
for 𝑚 ℓℓ > 30 GeV, in agreement with the SM prediction. The statistical significance of the result is well
above five standard deviations relative to the background-only hypothesis, which establishes observation of

this rare process using 𝑠 = 13 TeV data.

23
10 0.7

dσ / dNjets [fb]

rel
jets
C
ATLAS Data, stat. u

1 / σ dσ / dNA
1.5 Sherpa
13 TeV, 140 fb-1 0.6 aMC@NLO
8 ttW± Particle Level Powheg+H
0.5
1
6
0.4
0.5
0.2
dσ [fb × GeV-1]

ATLAS 4 0.3
Data
0.18 0
s = 13 TeV, 140 fb-1 MG5_aMC@NLO+Pythia 8
0.2
0.16 Sherpa 2.2.1 (incl.)
Sherpa 2.2.11 (multi-leg) 2 −0.5
d pZ
T

0.14 Stat. uncertainty 0.1


0.12 Total uncertainty
1.5
1.5
0.1 1.5
1 11
0.08 0.5 0.5
0.5
0.06 1.5
1.5 1.5
Particle-Level Njets
0.04 1 11
0.5 0.5
0.02 0.5

Prediction

Prediction
Prediction
1.5
1.5
1.5
Data

Data
Data
1.30 100 200 300 400 500 600 700 800 900 1000 Particle-Level Njets
Prediction

1 11
Data

1 parton_pt_z1*0.001 0.5 0.5


0.5
0.7
0 100 200 300 400 500 600 700 800 900 1000 2 3 4 5 6 7 2 3 4 5 22 33 44
2lSS 3l 2lS
2lS
Parton-level pZ [GeV]
T Particle-Level Njets

(a) (b)

Figure 9: (a) Absolute differential ttZ cross-section measurement as a function of the transverse momentum of the 𝑍
boson [147]. (b) Unfolded distribution of the absolute ttW cross-section as a function of jet multiplicity [146].

5.2 Top-quark production in association with a photon

The top quark’s coupling to a photon is tested through the measurement of tt𝛾 production and its kinematic
properties. For instance, the transverse momentum of the photon would be affected by anomalous dipole
moments of the top quark [154–156] or by EFT operators. At the LHC, tt𝛾 production was already

observed in 𝑠 = 7 TeV 𝑝 𝑝 collisions [157] and the first differential measurements were performed at
√ √
𝑠 = 8 TeV [158]. At 𝑠 = 13 TeV, benefiting from larger data samples, further differential distributions
were explored, first using 36 fb−1 of data [159] and then the full Run 2 dataset [160]. One challenge in the
tt𝛾 analyses is due to the fact that the photon can originate not only from the top quark but also from its
charged decay products, which would dilute the information about the 𝑡𝛾 coupling. Although separating
these sources is difficult, some kinematic variables such as the angular separation between a lepton and
photon in the event can help. Another challenge in such analyses is the accurate simulation of the signal.
Samples could be produced either inclusively without explicitly including a photon in the final state (but
where photons are generated through initial- or final-state radiation), or with photons at the matrix-element
level, taking the whole decay chain into account at LO. The overlap between these two kinds of samples
then needs to be removed. The tt𝛾 process also interferes with the singly resonant 𝑡𝑊 𝛾 process, similarly
to the interference of the tt and 𝑡𝑊 processes. NLO calculations are available in dedicated phase spaces,
i.e. with specific photon kinematic requirements [161–163]. Hence, in order to accurately compare the
data with theory computations, the analyses need to be performed in the same phase space. The first Run 2
analysis, using 36 fb−1 , exploited both the semileptonic and dileptonic decays of the tt pair. The analysis
using the full 139 fb−1 dataset [160] focused on the clean 𝑒𝜇 final state so that no multivariate technique
had to be used to extract the signal. It also measured the combined resonant tt𝛾 and non-resonant 𝑡𝑊 𝛾

24
production cross-section in order to compare it with the NLO QCD predictions in Refs. [162, 163]. Apart
from the backgrounds with prompt photons (such as 𝑊 𝛾 and 𝑍𝛾), there are two sources of background
with misidentified photons: hadron-fakes (photons mimicked by hadronic energy deposits, or non-prompt
photons from hadron decays) and electron-fakes (electrons mimicking photon signatures). In this analysis,
they were both estimated using MC samples since the studies performed with data-driven techniques in
the 36 fb−1 analysis [159] showed that possible data-driven corrections have a negligible effect on the
distribution shapes of relevant observables. The background from electron-fakes is a minor background
contribution in the analysis. The fiducial inclusive cross-section was extracted using a profile-likelihood fit
of the distribution of the scalar sum of all transverse momenta in the event, including leptons, photons,
jets and missing transverse momentum. This observable was found to provide good separation between
signal and background without too much sensitivity to systematic uncertainties. The fiducial region’s
definition required one electron and one muon, as well as two 𝑏-jets, and one photon with 𝐸 T > 20 GeV
and |𝜂| < 2.37 at parton level. The combined tt𝛾/𝑡𝑊 𝛾 cross-section was measured to be:
+2.6 +2.7
𝜎fid (tt𝛾 → 𝑒𝜇𝑏 𝑏𝛾)
¯ = 39.6 ± 0.8 (stat.) −2.2 (syst.) fb = 39.6 −2.3 fb,
+1.2
in good agreement with the dedicated theoretical calculation: 38.5 −2.5 fb [162, 163]. The systematic
uncertainties with the largest impact come from the uncertainties in modelling the signal. The absolute and
normalised differential cross-sections were measured as a function of the kinematics of the photon, the
angular separation between the photon and the leptons, and the pseudorapidity difference and azimuthal
angle between the two leptons. The last two of these are particularly sensitive to tt spin correlations.
The unfolded distributions generally agree well with the predictions, except for the shape of the angular
separation between the leptons and the photon or the azimuthal angle between the two leptons, which are
not well modelled by the LO MC predictions. The NLO prediction provides a better description of these
distributions. The dominant uncertainty in these differential measurements still comes from the size of the
data sample, at a level slightly below 10%.
Analogously to the 𝑡–𝑍 coupling, the coupling between the top quark and a photon can also be studied in the
single-top process 𝑡𝑞𝛾, featuring a forward light-quark jet characteristic of 𝑡-channel production. Analysing

the full 𝑠 = 13 TeV dataset led to the first observation of this process [164]. The 𝑡𝑞𝛾 cross-section was
measured in a fiducial phase space either at parton level, excluding contributions where photons are radiated
from the charged decay products, or at particle level, including these contributions. Two signal regions
were defined according to the presence or absence of a forward jet. Control regions were included to
normalise the large background from tt𝛾 production and the 𝑊 𝛾 background. A neural network was trained
to separate signal from background by using the reconstructed top-quark mass as the most discriminating
input variable. The measured fiducial cross-section (requiring one electron or muon, one 𝑏-jet, and one
photon with 𝑝 T greater than 20 GeV) is:
+75 +78
𝜎fid (𝑡𝑞𝛾) = 688 ± 23 (stat.) −71 (syst.) fb = 688 −75 fb,
+36
which is 2.1 standard deviations above the SM NLO QCD prediction of 515 −42 fb from Mad-
Graph5_aMC@NLO. The largest systematic uncertainty comes from the modelling of the tt𝛾 background.
The observed (expected) significance of the 𝑡𝑞𝛾 signal relative to the background-only hypothesis is 9.3
(6.8) standard deviations, establishing observation of this process.

5.3 Four-top-quark production

The Run 2 data sample also gave access to four-top-quark production, which is one of the rarest and
heaviest-final-state processes now accessible at the LHC, with a combined particle rest mass of almost

25
700 GeV. The Run 2 dataset is expected to contain around 1700 four-top-quark events. This multiparticle
SM process presents a promising avenue to search for signals of new physics beyond the SM. For example,
the tttt cross-section could be enhanced in top-quark-compositeness models [165] or by gluino pair
production in supersymmetric theories [166, 167]. Within the EFT approach, four-top-quark production
is uniquely sensitive to four-top-quark operators. Because of the existence of electroweak tttt Feynman
diagrams where the production of a pair of top quarks is mediated by a Higgs boson, tttt production
is also sensitive to the top-quark Yukawa coupling and its 𝐶𝑃 properties [168, 169]. Measuring the
four-top-quark production cross-section is interesting in its own right since experimental results will
challenge the state-of-the-art perturbative QCD calculation techniques. Within the SM the predicted tttt

cross-section in 𝑝 𝑝 collisions at a centre-of-mass energy of 𝑠 = 13 TeV is 𝜎tttt = 12.0 ± 2.4 fb [170–172]
at NLO in QCD including NLO electroweak corrections. This value does not include the effect of threshold
resummation at next-to-leading-logarithm accuracy, which increases the total production cross-section
by approximately 12% and reduces the scale uncertainty [173]. This corresponds to about a factor of 10

enhancement relative to the tttt cross-section at 𝑠 = 8 TeV, demonstrating the new perspective offered by
13 TeV collisions.
The analysis was first performed in the channel with two SS leptons or at least three leptons. This channel
corresponds to 12% of the total tttt production cross-section, and offers the best discovery potential. First,
a search for new phenomena targeted vector-like quark and SS top-quark pair production, and placed an
upper limit on SM tttt production using 36 fb−1 of Run 2 data [174]. Later, dedicated analyses focusing on
SM production were developed, finding first evidence for tttt production using the full 139 fb−1 Run 2
dataset [175]. With the same dataset but with improvements in object reconstruction, calibration and
selection criteria, new analysis techniques and a better understanding of major background processes and
systematic uncertainties, this process was first observed [176] as described in the following. Backgrounds
in the multilepton-channel tttt analysis arise almost entirely from previously described top+𝑉 processes
(i.e. ttW, ttZ, and tt𝐻 production) when they produce additional jets. These backgrounds with prompt
leptons were estimated using MC simulation. Because the theoretical modelling of the ttW background at
high jet multiplicity suffers from large uncertainties and since, as described above, the measured inclusive
ttW cross-section is higher than the SM expectation, a data-driven estimation of this important background
was implemented. The overall normalisation of the ttW background and the parameters of the scaling as a
function of jet multiplicity were determined from dedicated control regions together with the signal. The
fake-lepton backgrounds were evaluated using the template method. The tttt signal was separated from the
background events using a multivariate discriminant built with a graph neural network (GNN). The tttt
production cross-section and the normalisation factors for the backgrounds were determined via a binned
likelihood fit to the GNN score distribution in the signal region while also using control regions, with
systematic uncertainties included as nuisance parameters. The measured tttt production cross-section is:
+4.7 +4.6 +6.6
𝜎(𝑡 𝑡¯𝑡 𝑡¯) = 22.5 −4.3 (stat.) −3.4 (syst.) fb = 22.5 −5.5 fb.

The significance of the observed (expected) signal is found to be 6.1 (4.3) standard deviations relative to
the background-only hypothesis, providing the first observation of this process [176].
The measured production cross-section is consistent with the SM prediction to within 1.8 standard deviations.
Several limits on four-heavy-flavour-fermion EFT operators were also set using this measurement (see
Section 9). In addition to tttt production being dependent on the top Yukawa coupling, the tt𝐻 background
is a function of the same coupling. The GNN distribution was therefore used to extract limits on the top
Yukawa coupling’s strength modifier 𝜅 𝑡 , leading to 𝜅 𝑡 < 1.8 (assuming a 𝐶𝑃-even coupling). This limit
is less stringent than the ones derived from specific Higgs boson studies [177, 178] but is less model

26
dependent, without any assumption about the Higgs boson’s width. This tttt analysis also derived 95%
confidence level (CL) intervals for the cross-section of the 𝑡 𝑡¯𝑡 process, which has an experimental signature
and kinematic properties very similar to those of tttt events. This process, composed of two components
𝑡 𝑡¯𝑡𝑞 and 𝑡 𝑡¯𝑡𝑊, has an expected SM cross-section that is about 10 times smaller than for tttt, and has not yet
been observed. More data and more dedicated analyses will be needed to constrain this very rare process.
Although it is significantly less sensitive, a tttt measurement was also performed in events with a single
lepton or two OS leptons, first as a search for heavy new particles [179] using 36 fb−1 , and then targeting

the SM process with the full 𝑠 = 13 TeV dataset [180]. These final states have much higher branching
fractions (about 57% of the tttt events) but considerable background from top-quark pair production with
additional jets. Since the background in the high jet-multiplicity regions was found to be mismodelled by
MC simulation, a strategy was developed to use data to sequentially reweight the tt MC generation in several
observables to obtain a reliable tt+jets estimate. The different tt+jets components after reweighting (tt+light,
tt+≥1𝑐 and tt+≥1𝑏 jets) were further adjusted and constrained in a binned profile-likelihood fit which
extracted the signal strength. In the region most sensitive to tttt production, BDTs were used to discriminate
between signal and background events. The systematic uncertainties in the tt background prediction
were evaluated with special care since these uncertainties have the largest impact on the measurement
+15 +17
sensitivity. The tttt cross-section was measured to be 𝜎(𝑡 𝑡¯𝑡 𝑡¯) = 26 ± 8 (stat.) −13 (syst.) fb = 26 −15 fb,
which corresponds to an observed significance of 1.9 standard deviations relative to the background-only
hypothesis (while 1.0 standard deviations was expected). This result is compatible with the result in the
multilepton channel.
Now that the tttt process has been observed, the next step, as already taken for ttZ and ttW production,
will be to test the SM predictions differentially with tttt events and to better constrain the small but hard to
separate 𝑡 𝑡¯𝑡 process.

The 𝑠 = 13 TeV inclusive top+𝑋 cross-section measurements are summarised in Figure 10. As described
above, a lot of the top+𝑉 processes are intertwined, so a strategy to evaluate all of their cross-sections
coherently is needed to generically constrain new physics [181]. In order to combine the analyses in the
future, this would require the different analyses to harmonise how they define the reconstructed objects
used in their selection criteria, harmonise their systematic uncertainties, and harmonise the phase spaces
defining their control and signal regions.

6 Top-quark mass

The large value of its mass, 𝑚 𝑡 , is probably the most peculiar property of the top quark. It is close to the
electroweak scale and thus plays an important role in much of the dynamics of elementary particles via
loop diagrams. In particular, it significantly affects the radiative corrections to both the Higgs boson’s mass
and the 𝑊 boson’s mass, providing a relationship that can be used for precision tests of the consistency of
the SM [182].
Measurements of the top-quark mass are typically categorised into two families, often called ‘direct’ and
‘indirect’ measurements. Measurements relying on the reconstruction of the decay products of the top
quark, building partial or total invariant-mass observables, are usually assigned to the first category, and
are therefore sometimes referred to as ‘measurements from decay’. In contrast, extractions from total
or differential cross-section measurements, relying on the dependence of theoretical predictions on the
𝑚 𝑡 parameter in the SM Lagrangian, belong to the second category, also referred to as ‘measurements

27
ATLAS
statistical uncertainty
s=13 TeV
theory prediction stat. ⊕ syst. uncertainty
theory uncertainty

± stat. ± syst.
ttZ -1
[140 fb ] 860 ± 40 ± 40
arXiv:2312.04450

ttW [140 fb-1] 880 ± 50 ± 70


arXiv:2401.05299

tZq -1
[139 fb ] 97 ± 13 ± 7
JHEP 07 (2020) 124

ttγ /tW γ (fid.) -1


[139 fb ] 39.6 ± 0.8 ± 2.4
JHEP 09 (2020) 049

tqγ (fid.) [139 fb-1] 688 ± 23 ± 73


PRL 131 (2023) 181901

tttt [140 fb-1] 22.5 ± 4.5 ± 4.0


EPJC 83 (2023) 496

1 10 102 103 104


Cross-section (fb)


Figure 10: Summary of the 𝑠 = 13 TeV top+𝑋 cross-section measurements and comparison with the corresponding
theoretical expectations. The fiducial region used to measure the tt𝛾 cross-section requires one electron, one muon,
two 𝑏-jets and one photon at parton level. The 𝑡𝑞𝛾 fiducial region is defined by requiring one electron or muon, at
least one photon and at least one 𝑏-tagged jet at particle level.

from production’. The direct measurements typically have smaller uncertainties but inevitably rely on
predictions from MC generators to relate the considered observable to a top-quark mass value. Interpreting
these measurements in a well-defined renormalisation scheme is subject to additional uncertainties that
are challenging to evaluate [183, 184]. In addition, the most precise direct top-quark mass measurements
are affected by the relatively large uncertainties in the measurement of hadronic jet energies [185–187].
Alternatively, indirect measurements can be interpreted in a theoretically cleaner way, but are often affected
by even larger uncertainties of both experimental and theoretical nature.

6.1 Direct top-quark mass measurements

The most precise experimental 𝑚 𝑡 determination by the ATLAS Collaboration was obtained through a
combination of its Run 1 direct measurements, with the dominant contributions coming from the 8 TeV
measurements in the dilepton [188] and lepton+jets [185] channels. In order to minimise the impact of the
large jet-energy scale and tt modelling systematic uncertainties, the design of the two measurements took

28
orthogonal directions to some extent. In the dilepton channel, no attempt was made to fully reconstruct
the top quark through its decay products, in order to avoid relying on the determination of the missing
transverse momentum for the reconstruction of the two escaping neutrinos in each of the selected events.
The 𝑚 𝑡 value was then extracted with an unbinned maximum-likelihood fit to the partial top-quark invariant
mass formed by a charged lepton and the corresponding 𝑏-tagged jet, 𝑚 ℓ𝑏 , after imposing a lower bound
on the lepton–𝑏-tagged-jet system’s transverse momentum, optimised to minimise the uncertainty in the
measurement. On the other hand, the measurement performed in the lepton-plus-jet channel relied on a
simultaneous fit of three distributions: the top-quark and 𝑊-boson mass distributions as reconstructed
by a kinematic fit in each event, and the distribution of the ratio of 𝑏-jet to light-jet transverse momenta.
Thanks to these three variables, 𝑚 𝑡 could be extracted at the same time as two overall jet-energy correction
factors for 𝑏-jets and light jets, with a consequent reduction of jet energy scale uncertainties via such in
situ constraints. The combination of the two measurements, which were affected in different ways and in
different directions by some of the most relevant sources of systematic uncertainty, could then benefit from
this difference in design. The result is:

𝑚 𝑡 = 172.69 ± 0.25 (stat.) ± 0.41 (syst.) GeV,

corresponding to a total uncertainty of 0.48 GeV. Combining this with similar measurements by the

CMS Collaboration based on the Run 1 data collected at 𝑠 = 7 and 8 TeV [189], yields an even smaller
uncertainty:
𝑚 𝑡 = 172.52 ± 0.14 (stat.) ± 0.30 (syst.) GeV.

With the LHC providing a larger dataset in Run 2 than in Run 1, ATLAS had the opportunity to repeat
some of the most precise measurements performed in Run 1 and also to investigate new methods for
measuring the top-quark mass, using observables less sensitive to hadronic-jet energy determination. Using
36 fb−1 , ATLAS measured the top-quark mass using a purely leptonic observable, taking advantage of
semileptonic decays of 𝑏-hadrons in top-quark decays [190]. In this analysis the idea is to select tt events
where one of the two top quarks decays leptonically (𝑡 → 𝑊 𝑏 → ℓ𝜈𝑏), and then to require the presence of
a relatively soft muon within the hadronic jet formed in the 𝑏-quark fragmentation process. In this way, the
partial top-quark invariant mass 𝑚 ℓ 𝜇 could be built as the invariant mass of the system composed of the
prompt charged lepton ℓ from the 𝑊-boson decay (considering either an electron or muon) and the ‘soft’
muon from the 𝑏-jet. Provided that the prompt lepton and soft muon originate from the same leg of the tt
decay process, the 𝑚 ℓ 𝜇 value is strongly correlated with the top-quark mass, while having only a small
dependence on the jet reconstruction and energy determination.5 Soft muons were selected by requiring
them to have an angular separation of Δ𝑅 < 0.4 from a reconstructed jet. In order to reduce the rate of
kaons misidentified as muons, as well as the contribution from 𝑏 → 𝑐 → 𝜇 decay chains (where 𝑏 and 𝑐
here represent generic 𝑏- and 𝑐-flavoured hadrons), soft muons were required to satisfy tight identification
criteria and to have 𝑝 T > 10 GeV. The contamination from events where the two leptons come from
different legs of the tt decay was mitigated by a cut on the angular distance between the prompt lepton
and the soft muon. Finally, to control the residual 𝑏 → 𝑐 → 𝜇 contribution, due to its lower sensitivity
to 𝑚 𝑡 , selected events were separated into two categories, depending on whether the two leptons have
equal or opposite electric charge. The top-quark mass was then obtained from a simultaneous binned
profile-likelihood fit of the 𝑚 ℓ 𝜇 distribution in the two event categories , yielding:

𝑚 𝑡 = 174.41 ± 0.39 (stat.) ± 0.66 (syst.) ± 0.25 (recoil) GeV,


5
A residual jet-energy uncertainty still affects the final result, due to the jet-related selection requirements that need to be applied
in the analysis, in terms of event selection and soft-muon identification.

29
where the statistical uncertainty and the contribution from systematic uncertainties are indicated separately.
This result is complementary to other more traditional direct top-quark mass measurements, as it is largely
unaffected by jet energy scale uncertainties (with an impact of ±0.13 GeV on 𝑚 𝑡 ). However, uncertainties
in both the perturbative and non-perturbative parts of the 𝑏-fragmentation process (i.e. in the so-called
parton-shower evolution and fragmentation function) have a larger impact on the final result (∼0.2 GeV),
despite having been reduced by carefully retuning the parton-shower and hadronisation model in the MC
simulation to the most precise 𝑒 + 𝑒 − data from LEP and SLD. Moreover, uncertainties in the 𝑏-hadron decay
fractions to different final states make the largest contribution to the total uncertainty in 𝑚 𝑡 (±0.4 GeV). In
addition, studies have been performed, in contact with the theory community and MC experts, on the impact
of the choice of recoil scheme in the simulation of the top-quark decay and successive QCD radiation off
the 𝑏-quark. The impact of changing the default gluon-recoil scheme, from recoiling against the 𝑏-quark to
recoiling against the 𝑊 boson or against the top quark, was found to be sizeable. This affects the modelling
of the second and subsequent gluon emissions from quarks produced by coloured-resonance decays and
therefore changes both the fraction of jet energy carried by 𝑏-hadrons and the amount of radiation that fails
to be clustered in 𝑏-jets. These studies were used to derive a corresponding uncertainty (named the recoil
uncertainty), reported as the third contribution to the total uncertainty in the result quoted above. This
uncertainty was not included in Run 1 measurements.
In parallel, ATLAS further developed the technique used for the most precise Run 1 𝑚 𝑡 measurement,
namely its extraction from the unbinned maximum-likelihood fit in the dilepton channel [191]. By taking
advantage of the full 140 fb−1 Run 2 dataset, the analysis could be refined by optimising both the final-state
reconstruction and the event selection in order to maximise the resolution of the defined observable and to
minimise the impact of jet energy scale and tt modelling systematic uncertainties. In particular, a deep
neural network (DNN) was trained to identify the best 𝑏-jet candidate to be assigned to each of the two
charged leptons for the partial top-quark reconstruction. After this DNN-based lepton–𝑏-jet (ℓ𝑏) pairing
selects the two pairs, only the one with higher 𝑝 T is used to build the final observable, following the
same rationale as was behind the optimised event selection in the Run 1 analysis. Similarly to Run 1,
top-quark-mass templates were built from simulated tt events for the 𝑚 ℓ𝑏 observable. An unbinned
maximum-likelihood fit to the observed data events was used to extract the central value of 𝑚 𝑡 , with
systematic uncertainties being estimated by repeating the fit on varied pseudo-data samples. The resulting
measurement is:
𝑚 𝑡 = 172.21 ± 0.20 (stat.) ± 0.67 (syst.) ± 0.39 (recoil) GeV,
with the last component of the uncertainty representing the impact of the choice of recoil scheme in the
top-quark decay, as in the previously discussed analysis. The other systematic uncertainties are dominated
by jet energy scale and resolution uncertainties (with an impact of ∼0.41 GeV on 𝑚 𝑡 ) and by uncertainties in
the matching scheme used between the NLO hard-scattering and parton-shower MC generators (±0.4 GeV).
Other important uncertainties are those from the initial- and final-state QCD radiation (±0.17 GeV) and
colour-reconnection modelling (±0.27 GeV). The impact of the choice of recoil modelling is large and
similar to the case of the previously described analysis, as it directly affects the fraction of energy carried
by the undetected neutrino in the leptonic top-quark decay.
Figure 11 shows a comparison of the ATLAS Run 2 𝑚 𝑡 measurements with the results of the combination of
the ATLAS direct 𝑚 𝑡 measurements in Run 1 and of the combination of the ATLAS and CMS measurements.
It can be seen how these Run 2 ATLAS measurements are affected by mostly uncorrelated sources of
uncertainty. A possible gain can be foreseen from repeating the measurements on larger datasets or in
different final states, as well as from performing an updated combination, including Run 1 measurements.

30
ATLAS Preliminary statistical uncertainty
ATLAS+CMS Run 1 combination stat. ⊕ syst. uncertainty
arXiv:2402.08713 stat. ⊕ syst. ⊕ recoil unc.
statistical uncertainty *: Preliminary
total uncertainty
± stat. ± syst. ± recoil

ATLAS Run 1 combination 172.69 ± 0.25 ± 0.41


Eur. Phys. J. C 79 (2019) 290
[20.2 fb-1 8 TeV + 4.6 fb-1 7 TeV]

Run 2 soft-muon tag 174.41 ± 0.39 ± 0.66 ± 0.25


JHEP 06 (2023) 019
[36 fb-1 13 TeV]

Run 2 dilepton* 172.21 ± 0.20 ± 0.67 ± 0.39


ATLAS-CONF-2022-058
[140 fb-1 13 TeV]

168 170 172 174 176 178 180


mt [GeV]

Figure 11: Summary of the most precise top-quark mass measurements performed by ATLAS in Run 1 and Run 2.
The reference value shown by the vertical dashed line, with the blue bands indicating the statistical and total
uncertainties, comes from the combination of the ATLAS and CMS Run 1 top-quark mass measurements. For
each of the measurements, the uncertainty is broken down into a statistical uncertainty and a systematic uncertainty
component, with the effect of the recoil modelling indicated separately for the Run 2 measurements (see text for more
details).

6.2 Indirect top-quark mass measurements

For the indirect measurements, ATLAS took advantage of the improved precision in the determination of
the inclusive tt cross-section (see Section 3) relative to Run 1. As reported in Ref. [82], the inclusive tt
cross-section measurement in the dilepton 𝑒𝜇 channel based on 36 fb−1 of Run 2 data was used to extract
pole +2.0
the top-quark pole mass via its effect on the predicted cross-section, yielding 𝑚 𝑡 = 173.1−2.1 GeV.
The result, obtained with CT14 as the reference PDF set, is dominated by uncertainties in the theoretical
cross-section evaluated through PDF+𝛼s and QCD-scale variations. Therefore, improved experimental
measurement precision (such as that achieved in the updated cross-section measurement based on the full
Run 2 dataset) would not have significant effects on this determination. Instead, extractions from differential
cross-section measurements, especially when using the same technique as in the Run 1 top-quark mass
measurement from tt+1-jet production [192] which obtained a precision of ∼1 GeV, have the potential to
improve the precision of indirect determinations.

7 Top-quark properties

In addition to its very large mass, the top quark has other unique properties. Its decay width (Γ𝑡 ≈
1.4 GeV [193]) is larger than the QCD hadronisation scale Λ ≈ 250 MeV, so the top quark decays before it

31
can form bound states, unlike other quarks. This unique feature allows its properties to be studied through
its decay products.
Precisely predicted properties of top-quark decays are measured as a stringent test of the SM and as a probe
of contributions from BSM physics. These include the polarisation of the 𝑊 bosons produced in top-quark
decays, the spin correlations and entanglement of top-quark pairs, and their forward–backward asymmetry.
Additionally, top-quark decays are used both to test the modelling of QCD effects such as colour connection
and as a source of 𝑊 bosons to test the universality of their couplings to different lepton families.
These properties were studied extensively in Run 2 data, benefiting from a data sample much larger than
the one from Run 1. As it is cleaner than the other channels, the dilepton channel was often used for these
studies despite the reconstruction of top-quark kinematics being more complex there.

7.1 Top-quark decay angular properties

Since the top quark decays almost exclusively into a 𝑊 boson and a 𝑏-quark, its decay products are
naturally useful for studying the 𝑊𝑡𝑏 interaction vertex. This vertex and the masses of the interacting
particles define the fractions of 𝑊 bosons with longitudinal ( 𝑓0 ), left-handed ( 𝑓L ), and right-handed ( 𝑓R )
polarisation from the top-quark decays. Because of the vector minus axial-vector (𝑉 − 𝐴) nature of the
𝑊𝑡𝑏 interaction, 𝑓R is expected to be very small, making it particularly sensitive to signs of new physics.
The helicity fractions are predicted at NNLO in QCD to be 𝑓0 = 0.687 ± 0.005, 𝑓L = 0.311 ± 0.005 and
𝑓R = 0.0017 ± 0.0001 [194], with 𝑓0 + 𝑓L + 𝑓R = 1. These fractions could be altered by new physics
processes, and deviations from the SM expectation can be parameterised in terms of dimension-six EFT
operators affecting the 𝑊𝑡𝑏 vertex. The full Run 2 data sample was used to measure these fractions [195] in
the dilepton tt final state. They are accessible through the normalised differential distribution of the cosine
of the polar angle 𝜃 ∗ , defined as the angle between the momentum of the charged lepton from the 𝑊-boson
decay and the reversed momentum of the 𝑏-quark from the top-quark decay, both calculated in the 𝑊
boson’s rest frame. Measuring cos 𝜃 ∗ requires the reconstruction of the tt kinematics, which is achieved by
using the Neutrino Weighting method. Using top-quark and 𝑊-boson mass constraints, this method scans
over the neutrino pseudorapidities to find two possible kinematic solutions that are the most compatible

with the measured 𝐸 Tmiss of the event. Unlike the 𝑠 = 8 TeV results, which were obtained using a template

method applied to detector-level distributions, the 𝑊 helicity fractions in the 𝑠 = 13 TeV analysis were
extracted by unfolding the normalised differential cos 𝜃 ∗ distribution to parton level and fitting the unfolded
cross-section distribution, minimising the 𝜒2 value. In the fit, the 𝑓0 parameter was set to 𝑓0 = 1 − 𝑓L − 𝑓R .
The results are 𝑓0 = 0.684 ± 0.005(stat.) ± 0.014(syst.), 𝑓L = 0.318 ± 0.003(stat.) ± 0.008(syst.), and
𝑓R = −0.002 ± 0.002(stat.) ± 0.014(syst.) in agreement with the SM predictions to within one standard
deviation. The systematic uncertainty dominates the total uncertainty for all three helicity fractions. The
largest systematic uncertainty arises from the uncertainty in tt modelling, so better understanding of the tt
MC simulation would be needed to improve this measurement.
Top-quark pair production at the LHC is mostly mediated by the parity-invariant strong interaction, so the
top quarks and antiquarks are predicted to be produced unpolarised in the SM, while the spins of the top
quark and top antiquark are expected to be correlated. Since the spin correlation is transferred to their
decay products, and almost maximally to the two leptons [196], the study can be performed with the leptons
from the top-quark and top-antiquark decays. The dilepton channel is particularly relevant for measuring
spin correlations because, in addition to their sensitivity to spin correlations, charged leptons are easy to
identify in hadron collisions. This correlation has been observed experimentally at the LHC using both

32

𝑠 = 7 TeV [197, 198] and 8 TeV [199, 200] collisions, showing slightly stronger correlation than expected,

although with rather large experimental uncertainties. Studies of spin correlation at 𝑠 = 13 TeV with

large datasets are therefore particularly relevant. Using 36 fb−1 of 𝑠 = 13 TeV data, ATLAS performed a
measurement in the channel with one electron and one muon [201]. The full spin information of the top
quarks is encoded in the spin density matrix [202]. The simplest observable sensitive to spin correlation
is the azimuthal opening angle Δ𝜙 between the electron and the muon in the transverse plane, measured
in the laboratory frame. The spin correlation measurement can be performed inclusively, but also in
different tt invariant-mass bins since the degree of correlation is expected to vary with the tt invariant

mass. The 𝑠 = 13 TeV analysis also utilised the difference between the pseudorapidities of the two
charged leptons as an additional observable, Δ𝜂. This observable is less sensitive to spin correlation
than Δ𝜙 but, in addition to Δ𝜙, is useful when searching for the presence of supersymmetric top squarks
with a mass close to the top-quark mass. The tt invariant mass was reconstructed using the Neutrino
Weighting method. The Δ𝜙 and Δ𝜂 distributions were corrected to parton level and particle level using
an iterative Bayesian unfolding, and the resulting absolute and normalised cross-sections were compared
with predictions from NLO MC generators. The dominant systematic uncertainty in this measurement
comes from the modelling of initial- and final-state radiation. The comparison revealed several shape
effects, with data tending to be higher than the expectation at low Δ𝜙 or high Δ𝜂 (see Figure 12 (a)). The
compatibility with the SM prediction was assessed using a template fit to the normalised parton-level
cross-sections, with one template from dilepton tt events with SM spin correlation and one where spin
correlation had been removed. Using the inclusive Δ𝜙 distribution, the extracted spin-correlation fraction
+0.067
is 𝑓SM = 1.249 ± 0.024 (stat.) ± 0.061 (syst.) −0.090 (theo.). When including the template’s theoretical
uncertainties, the measurement is 2.2 standard deviations higher than the SM expectation of 𝑓SM = 1. The
value of 𝑓SM is observed to increase slightly as a function of the tt invariant mass, 𝑚 tt , but no bin shows a
significant deviation from the prediction, due to the still large statistical and systematic uncertainties, and
the relatively poor 𝑚 tt resolution. Several cross-checks were performed to understand the sensitivity of
the result to the limitations of the tt modelling, such as the impact of the narrow-width approximation,
the impact of NNLO corrections, or the use of expanded NLO predictions. None of these alternative
predictions agree completely with the measurements, even though including higher-order effects brings the
predictions closer to the data. Studies with the full Run 2 dataset or during the future LHC runs, as well
as improvements in the predictions, should shed further light on this difference. The double-differential
distributions of Δ𝜙 in bins of Δ𝜂 were used to search for the pair production of supersymmetric top squarks.
In the absence of a signal, top squarks with a mass between 170 GeV and 230 GeV were excluded for most
of the allowed neutralino mass range.
Precise measurements of the tt spin density matrix in a very restricted phase space were recently proposed as a
new laboratory to study quantum information properties, especially entanglement [203–207]. Entanglement
is a feature of quantum mechanics, where two entangled particles cannot be described independently of
each other. This has been observed in many systems but only at lower energy scales. The LHC provides
the opportunity to study this effect at high energies and in systems composed of new elementary particles.
The tt pair is a two-qubit system of two spin-1/2 fermions. Entanglement in top-quark pairs can be
studied via the spin correlation between the produced top quark and top antiquark, using leptons from their
decays as spin-analysing particles. When produced close to their production threshold (i.e. 𝑚 tt ≈ 2𝑚 𝑡 ),
the tt pairs produced through gluon–gluon fusion are in a spin-singlet state. In this case, maximum
entanglement between the top quark and top antiquark is expected. When the tt system has a larger mass,
the entanglement is reduced (but restored at very high tt mass above around 1 TeV). A simple observable
can signal the presence of entanglement close to the production threshold [203]: 𝐷 = −3⟨cos 𝜑⟩, where 𝜑
is the angle between the charged-lepton directions in the respective parent top-quark and top-antiquark

33
[1/(rad/π)]

ATLAS

1.2 s = 13 TeV, 36.1 fb−1
σ d ∆φ(l +, l −)/π

0.1 ATLAS
1.0

√s = 13 TeV, 140 fb-1

0.2
·
1

0.8

Particle-level D
Data Sherpa 0.3
Powheg Pythia8 Powheg Pythia6
Powheg Herwig7 PowPy8 rad. down
0.6
MG5 aMC@NLO Pythia8 0.4
Limit (Powheg + Herwig7)
Limit (Powheg + Pythia8)
1.05 Stat. Total Theory Uncertainty
Theory

0.5 Data
Data

1.00 Powheg + Pythia8 (hvq)


Powheg + Herwig7 (hvq)
0.95
0.6
0.2 0.4 0.6 0.8 340 < mtt- < 380 380 < mtt- < 500 mtt- > 500
Parton-level ∆φ(l +, l −)/π [rad/π] Particle-level Invariant Mass Range [GeV]

(a) (b)

Figure 12: (a) The parton-level normalised Δ𝜙 differential cross-section compared with various predictions [201]. (b)
Particle-level results for the measurement of the quantum entanglement observable 𝐷 in the signal and validation
regions, compared with various MC models [208].

rest frames. The existence of an entangled state is then demonstrated if 𝐷 < −1/3. ATLAS performed
the first study of quantum entanglement at high energy using tt events in the channel with one electron

and one muon using the full Run 2 dataset at 𝑠 = 13 TeV [208]. The measurement of 𝐷 requires the
reconstruction of the top-quark and top-antiquark kinematics. The main method used in this analysis
was the Ellipse method [209], which analytically calculates the unmeasured neutrino momenta through a
geometrical approach. If the Ellipse method failed, the Neutrino Weighting method was used instead. The
optimal window for the signal region was determined to be 340 < 𝑚 tt < 380 GeV at particle level. Two
validation regions where entanglement is expected to be small were also defined to validate the method
used for the measurement: one close to the threshold region (380 < 𝑚 tt < 500 GeV) and one at higher 𝑚 tt
(𝑚 tt > 500 GeV). The observed cos 𝜑 distribution at reconstruction level was corrected, after background
subtraction, for distortions from the detector response and event selection by using a simple calibration
curve that relates the reconstructed values to the corresponding particle-level values. The resulting value of
𝐷 at particle level in a fiducial phase space is 𝐷 = −0.547 ± 0.002 (stat.) ± 0.021 (syst.). This value is
compared with the entanglement limit at particle level, which was obtained by converting the parton-level
bound of 𝐷 = −1/3 to the particle-level equivalent taking into account parton-shower effects from the
Powheg+Pythia or Powheg+Herwig generators (see Figure 12 (b)). A large discrepancy between the
predictions from these two generators is observed. However, in both cases the observed 𝐷 value is well
below the entanglement limit, beyond five standard deviations, establishing the discovery of an entangled tt
state. This constitutes the first observation of entanglement in a quark–antiquark pair. The measurement
paves the way for further studies of fundamental quantum mechanics at the LHC, such as measurements of
quantum discord or the steering ellipsoid [210] or testing the Bell inequality in tt events [204, 211].

34
7.2 Asymmetry measurements

At leading order in QCD, tt production is symmetric under charge conjugation. However, at NLO,
an asymmetry occurs between the top quark and the top antiquark. This asymmetry comes from the
interference of initial-state and final-state radiation (ISR and FSR) diagrams and from the interference
of the Born and box diagrams for quark–antiquark initial states: 𝑞 𝑞¯ → tt. As a result, the top quark
(top antiquark) is preferentially emitted in the direction of the incoming quark (antiquark). Production
via gluon–gluon fusion, 𝑔𝑔 → tt, which is dominant at the LHC, is charge symmetric at all orders. The
asymmetry in 𝑞 𝑞¯ manifests itself by the top antiquark being produced more centrally while the top quark is
produced at more forward rapidities.
The charge asymmetry at the LHC can then be defined as:

tt 𝑁 (Δ|𝑦| > 0) − 𝑁 (Δ|𝑦| < 0)


𝐴C = ,
𝑁 (Δ|𝑦| > 0) + 𝑁 (Δ|𝑦| < 0)

with Δ|𝑦| = |𝑦 𝑡 | − |𝑦 𝑡¯ |. Since tt production by gluon–gluon fusion, which is symmetric, dominates at the
LHC, the inclusive asymmetry is expected to be small, of the order of 1%. For events where the top quark
and antiquark decay leptonically, the leptons from the top quark or antiquark inherit the directions of their
parent quarks, so a similar asymmetry can be defined using the pseudorapidities of the leptons via the
tt
Δ|𝜂| = |𝜂ℓ + | − |𝜂ℓ − | observable. This leptonic charge asymmetry 𝐴Cℓℓ is slightly diluted compared to 𝐴C
but has the advantage of not requiring the kinematic reconstruction of the tt pair. It is also interesting
to measure the asymmetries differentially as a function of kinematic variables of the tt system, such as
tt
the transverse momentum 𝑝 T , the invariant mass 𝑚 tt , and the boost of the tt system in the 𝑧 direction,
𝛽 𝑧,tt . In the SM, the charge asymmetries are expected to be enhanced for high values of some of these
variables. Several BSM models predict a modification of the asymmetry, for instance by anomalous vector
or axial-vector couplings (e.g. axigluons or a heavy 𝑍 ′ boson [212–215]). These modifications could also
be studied with EFT [216]. In particular, BSM effects are expected to be enhanced in specific kinematic
regions, such as in the phase space with large 𝛽 𝑧,tt or large 𝑚 tt [217].

Asymmetry measurements in tt events were first performed at 𝑠 = 8 TeV [218], although with rather
large uncertainties, around 60% of the predicted asymmetry. The fraction of 𝑞 𝑞-initiated
¯ top-quark pair
production decreases with increasing centre-of-mass energy, and hence the tt charge asymmetry decreases
tt
as well. Despite this disadvantage, ATLAS measured 𝐴C in both the single-lepton and dilepton channels,

and 𝐴Cℓℓ in the dilepton channel, at 𝑠 = 13 TeV [219]. The lepton+jets channel was further split into
resolved and boosted topologies. In the resolved case, the assignment of the jets to the corresponding
partons from the decaying top quarks was assessed with a BDT that aims to discriminate between signal
and the combinatorial background, separately for events with one or two 𝑏-tagged jets. In the boosted
topology, the selected large-radius jet was taken to be the hadronically decaying top quark. In both
cases, the semileptonically decaying top quark’s four-vector was reconstructed from the lepton and a
small-radius jet, calculating the neutrino four-vector from the missing transverse momentum and the
𝑊-boson mass constraint. In the dilepton channel, the tt system was reconstructed using the Neutrino
Weighting method. The differential Δ|𝑦| distributions were corrected for acceptance and detector effects
using the FBU method, where systematic uncertainties that affect the measurements are treated as nuisance
tt
parameters. The combined inclusive 𝐴C asymmetry from single-lepton and dilepton events was measured
+0.0005
to be 0.0068 ± 0.0010(stat.) ± 0.0010(syst.), in agreement with the SM calculation of 0.0064−0.0006 at
NNLO accuracy in the strong coupling with NLO electroweak corrections [220]. The SM computation

35
was performed by expanding the numerator and denominator to a given order in perturbation theory.
The measurement differs from zero by 4.7 standard deviations, providing strong evidence for tt charge
asymmetry at the LHC. The precision of the result is dominated by the lepton+jets channel because
¯
of its smaller statistical uncertainty. The 𝐴Cℓ ℓ asymmetry, measured in the dilepton channel only, is
0.0054 ± 0.0012(stat.) ± 0.0023(syst.), while the SM calculation at NLO in QCD, including NLO EW
+0.0002 tt
corrections, predicts 0.0040−0.0001 [221]. The combined 𝐴C results were interpreted in terms of EFT using
tt
new operators for four-quark interactions with different coupling chiralities (see Section 9). Differential 𝐴C
tt
measurements were performed as a function of 𝑚 tt , 𝑝 T and 𝛽 𝑧,tt , with the binning at larger values being

finer than was possible for 𝑠 = 8 TeV data (see Figure 14 (a)). Differential measurements of 𝐴Cℓℓ were
presented as a function of the invariant mass, transverse momentum and longitudinal boost of the dilepton
pairs. The results were found to be compatible with the SM predictions.
Another way to study the tt charge asymmetry is to use an observable linked to the energy difference
between the top quarks and antiquarks, Δ𝐸 = 𝐸 𝑡 − 𝐸 𝑡¯ . The energy asymmetry [222] occurs mainly through
the 𝑞𝑔 → tt𝑞 process, which is a more abundant source of events than the 𝑞 𝑞¯ → tt process at the LHC. It
is then expected to be larger than asymmetries based on rapidity. The presence of an additional jet allows
QCD effects at leading order to be investigated, while the asymmetry in 𝑞 𝑞¯ → tt only appears at NLO. In
the 𝑝 𝑝 → 𝑡 𝑡¯ 𝑗 process, the energy asymmetry can be defined as a function of the jet angle 𝜃 𝑗 :

𝜎tt 𝑗 (𝜃 𝑗 |Δ𝐸 > 0) − 𝜎tt 𝑗 (𝜃 𝑗 |Δ𝐸 < 0)


𝐴𝐸 (𝜃 𝑗 ) = ,
𝜎tt 𝑗 (𝜃 𝑗 |Δ𝐸 > 0) + 𝜎tt 𝑗 (𝜃 𝑗 |Δ𝐸 < 0)

where 𝜎tt 𝑗 (𝜃 𝑗 ) is the differential tt 𝑗 cross-section as a function of 𝜃 𝑗 . Both Δ𝐸 and 𝜃 𝑗 are defined in the
tt 𝑗 rest frame, which corresponds to the partonic centre-of-mass frame in tree-level processes. ATLAS

measured this energy asymmetry differentially using 139 fb−1 of 𝑠 = 13 TeV data [223]. The analysis
was performed in the semileptonic tt decay channel. The number of events observed at detector level was
corrected for detector effects to particle level using the FBU method. The uncertainty in the measurement
is dominated by the statistical component. The measured differential distribution was found to be in
good agreement with the SM expectation, with a 𝑝-value of 0.80. In the first bin, 0 ≤ 𝜃 𝑗 ≤ 𝜋/4, the
measured asymmetry differs from zero by 2.1 standard deviations. The sensitivity of this energy asymmetry
measurement to new physics was investigated in the context of EFT (see Section 9). The energy asymmetry
is particularly sensitive to the chirality and colour charges of the involved operators. It complements the
constraints from asymmetries built using rapidities.
The large Run 2 data sample allows asymmetries to be measured in rarer processes where tt is produced
with an associated gauge boson. Some of these processes are predicted to have a larger asymmetry than in
tt production. For instance, the ttW process is initiated at LO by a 𝑞 𝑞¯ ′ initial state. These ttW events can
then serve as an interesting tool for measuring the tt charge asymmetry since it is expected to be larger
than in tt production [224, 225]. In addition, the 𝑊 boson in this process can be radiated from the 𝑞 𝑞¯ ′
initial state, thereby serving as a way to polarise the 𝑞 𝑞¯ ′ pair and thus also the tt pair. This polarisation
further enhances the asymmetry between the decay products of the top quarks and top antiquarks, leading
to an enhanced leptonic asymmetry. The drawback of using the ttW process is, however, its much smaller
cross-section in comparison with tt production. Despite the challenges, ATLAS probed the asymmetry
in ttW events using the full Run 2 dataset [226]. The measurement was performed in the 3L channel at
detector level and also at particle level after unfolding. In order to compute 𝐴Cℓℓ , the two opposite-sign
leptons from the top-quark decays need to be separated from the one from the 𝑊 decay. This was addressed
by using a BDT. The second lepton needed to compute 𝐴Cℓℓ was taken to be the lepton with charge opposite

36
to that of the lepton selected by the BDT. A profile-likelihood fit was used to extract the signal, together
with the normalisation of each of the most relevant background processes, i.e. ttZ, non-prompt electrons
and muons, and electrons from 𝛾-conversions. The leptonic charge asymmetry in ttW events was measured
to be 𝐴Cℓℓ (𝑡 𝑡¯𝑊) = −0.12 ± 0.14(stat.) ± 0.05(syst.). The SM prediction from the Sherpa simulation [57]
+0.005
in this phase space is −0.084−0.003 (scale) ± 0.006(MC stat.). With the current precision, the result is
compatible with zero and not yet sensitive to the SM asymmetry. The asymmetry was also unfolded
to particle level in a fiducial phase space. The results are limited by statistical uncertainty, so they are
expected to improve in the years to come.
The tt𝛾 final state is also relevant for asymmetry measurements [227]. When the photon is emitted from
the initial state, the process benefits from an enhanced fraction of quark–antiquark-initiated tt production
relative to symmetric production via gluon–gluon fusion. In this process, the asymmetry arises through the
interference of QED initial-state radiation and final-state radiation, which yields a negative asymmetry.

The overall asymmetry in tt𝛾 events at 𝑠 = 13 TeV is predicted in the SM to be between −1% and
−2% depending on the phase space [228, 229]. The main challenge of studying the charge asymmetry
in tt𝛾 events comes from the fact that the asymmetry is only present for events where the photon is
radiated from the initial-state parton or from the final-state top quark or antiquark. It is diluted by tt𝛾
events where the photon is emitted from any of the charged decay products of the tt final-state system.
ATLAS measured this charge asymmetry in tt𝛾 production for the first time using the full 139 fb−1 dataset

recorded at 𝑠 = 13 TeV [230] and the semileptonic tt decay channel. The top-quark and top-antiquark
kinematic properties, in particular their rapidities, were reconstructed using the KLFitter package. The
estimation of background events with prompt or misidentified photons followed the methods developed for
the measurement of the tt𝛾 cross-section described in Section 5. A neural network was used to separate the
tt𝛾 signal from the background processes. The asymmetry value was extracted from the Δ|𝑦 𝑡 | = |𝑦 𝑡 | − |𝑦 𝑡¯ |
distribution in a fiducial region at particle level after performing a maximum-likelihood unfolding. After
tt 𝛾
the fit, the asymmetry was found to be 𝐴C = −0.003 ± 0.024(stat.) ± 0.017(syst.) (assuming a SM charge
tt
asymmetry in tt events of 𝐴C = 0.0064). The measured value is compatible with the result from NLO
tt 𝛾 tt 𝛾
simulation in the same phase space: 𝐴C = −0.014 ± 0.001. The dependence of the measured 𝐴C on the
asymmetry in tt events was also studied and found to be linear. Here also the result is still limited by the
statistical uncertainty and so is expected to be improved in the future.
The different asymmetry measurements are summarised in Figure 13. They are in agreement with the
SM predictions. The uncertainties in the SM predictions are much smaller than the current experimental
uncertainties, which are dominated by statistical uncertainties, as systematic effects largely cancel out.
Therefore, improvements in the experimental set-up, as well as the analysis of the larger datasets available
after the completion of LHC Run 3 and at the HL-LHC, will lead to more valuable comparisons.

7.3 Tests of QCD

Top pair production also serves as a unique environment to test QCD and tune the modelling of QCD effects
in MC generators. One of the aspects of QCD that can be tested with tt events is the modelling of radiation
through colour connection and colour propagation, i.e. colour flow, which is modelled by colour strings
between quarks and gluons. Experimentally, quarks and gluons appear as jets. The radiation emission
pattern between particles, governing the structure of the jets, is affected by their colour connections.
In particular, QCD predicts that there is more radiation between particles that are colour-connected.
Experimental measurements are needed to validate the phenomenological description of such predictions.

37
ATLAS
statistical uncertainty
s=13 TeV
theory prediction stat. ⊕ syst. uncertainty
theory uncertainty

± stat. ± syst.
A C(tt) =0.0068 ± 0.0010 ± 0.0010
-1
JHEP 08 (2023) 077 [139 fb ]

A ll(tt) =0.0054 ± 0.0012 ± 0.0023


-1
JHEP 08 (2023) 077 [139 fb ]

A E(ttj) =-0.0430 ± 0.0140 ± 0.0143


bin 2
EPJC 82 (2022) 374 [139 fb-1]

A C(ttγ ) =-0.003 ± 0.024 ± 0.017


PLB 843 (2023) 137848 [139 fb-1]

A ll(ttW)/10 =-0.012 ± 0.014 ± 0.005


JHEP 07 (2023) 033 [139 fb-1]

−0.14 −0.12 −0.1 −0.08 −0.06 −0.04 −0.02 0 0.02


Top asymmetry


Figure 13: Summary of the 𝑠 = 13 TeV top asymmetry measurements using the full Run 2 dataset: tt asymmetry,
lepton-based asymmetry, energy asymmetry in the bin with the largest expected asymmetry, and asymmetry in tt𝛾
and ttW events. Each of the results is compared with their corresponding theoretical expectation. The value of the tt
asymmetry in ttW events is shown divided by 10 for display purposes.

In tt production where only one of the top quarks decays leptonically, there are four quarks as decay
products. The two quarks from the decay of the 𝑊 boson, which is a colour-neutral object, are expected to
be colour-connected. The two 𝑏-quarks from the top-quark decays carry the colour of their parent top
quarks but are not expected to share any colour connection. These features can be tested experimentally.

Building on the 𝑠 = 8 TeV analysis [231], ATLAS performed a measurement of colour flow using 36 fb−1

of 𝑠 = 13 TeV data [232]. The observable used to encode the colour information was the so-called
jet-pull vector P® [233]. It is defined as the sum of the transverse momenta of the jet constituents 𝑖 weighted
by their distance from the jet axis, Δ® 𝑟 𝑖 = (Δ𝑦 𝑖 , Δ𝜙𝑖 ), in the 𝑦−𝜙 space and normalised by the jet 𝑝 T :
® Í
P = (1/𝑝 T ) 𝑖 |Δ® 𝑖
𝑟 𝑖 | 𝑝 T Δ®
𝑟 𝑖 . With two jets 𝑗1 and 𝑗2 in the 𝑦−𝜙 space, one can construct the jet-pull angle
𝜃 P between the jet-pull vector P® ( 𝑗1 ) and the vector connecting 𝑗1 and 𝑗2 . If the two jets 𝑗1 and 𝑗2 are
colour-connected, a bias toward 𝜃 P ∼ 0 is expected, whereas if they are not connected, the distribution of
the jet-pull angle is expected to be uniform. Applied to tt events, three of the four observables measured are
the jet-pull angles, 𝜃 P ( 𝑗1𝑊 , 𝑗2𝑊 ) and 𝜃 P ( 𝑗2𝑊 , 𝑗1𝑊 ), of the jets from the 𝑊-boson decays, 𝑗1𝑊 and 𝑗2𝑊 , taken
to be the two leading non-𝑏-tagged jets, and the magnitude of the pull-vector from 𝑗1𝑊 . These are expected
to show colour connections. The jet-pull angle of the two leading 𝑏-tagged jets was also measured but is not
expected to show colour connection. In this analysis, the jet-pull vectors were calculated only from tracks
in the inner detector. All four reconstructed observables were unfolded to particle level using the IBU
method and normalised, so as to study only their shape and mitigate the impact of systematic uncertainties.

38
The results were compared with several generator predictions that differ in their hadronisation modelling
(the string model for Pythia and the cluster model for Herwig). Agreement with the Herwig modelling
was found to be better. The unfolded distributions are also used to test the prediction of an exotic model
that implements flipped colour flow. The data disfavours this model, with 𝑝-values of at most 0.002 for all
the tested variables. These unfolded data can be used to tune MC generators to better model the effects of
colour connections between partons.
The top-quark sector is also useful for tuning colour models that can not be derived from QCD first
principles. In MC generators, the colour information is traced using the leading-colour approximation [234,
235], where each quark is connected to only one other parton, while each gluon, which carries a colour
and an anticolour, is connected to two other partons. The multiple parton interactions overlaid on the
hard-scattering process add additional colour lines, which could potentially lead to phase-space overlaps.
The colour-reconnection models [64] reassign colour connections between partons in order to resolve these
overlapping colour lines. There are several mechanisms that could be introduced and that should be tuned
to data. The exact involvement of the top quark and its decay products in these processes is not known. The
current models involved only the top quark itself in the colour-reconnection mechanisms, and not its decay

products, so it was important to test these models on data. With 139 fb−1 of 𝑠 = 13 TeV data, ATLAS
used tt events in the channel with one electron and one muon, and two or three jets (including exactly two
𝑏-tagged jets) to measure distributions sensitive to colour reconnection [236]. The chosen observables
were the charged-particle multiplicity outside the jets (excluding leptons from the top-quark decays), the
scalar sum of the transverse momenta of these charged particles, and the double-differential cross-section
in these two quantities. These observables need to be corrected for tracks from pile-up and from secondary
vertices, and for tracking inefficiencies. The corrected observables were unfolded to particle level using
the IBU method to obtain normalised differential cross-sections. The measurements have a precision of
2%–10% in the central bins and up to 50% in the outer bins. Agreement between the measured differential
cross-sections and various models implemented in MC generators was assessed by means of a 𝜒2 test.
The Herwig generator describes the data well for the observable built with the 𝑝 T scalar sum, while the
prediction from Pythia is better for the charge-particle multiplicity. The results were also compared
with predictions from different models of multiple parton interactions, since the chosen observables are
also sensitive to these. The results could be used for future tuning of the MC parameters for both colour
reconnection and multiple parton interactions, which should be performed simultaneously.
As discussed in Sections 2.2 and 3.2, hadronically decaying top quarks with sufficiently high transverse
momentum can be reconstructed as single large-radius jets. These boosted top-quark jets are characterised
by their distinctive substructure, which is often used to separate them from energetic jets arising from
lighter quarks or gluons [237–239]. It is therefore important to test the modelling of these features in MC
simulation, through a comparison with observed data. Moreover, deviations of top-quark-jet substructure
from the SM predictions could also serve as tests of BSM effects that may not be detectable with inclusive
cross-section measurements. ATLAS measured substructure properties of jets emerging from hadronically
decaying boosted top quarks, using the full Run 2 dataset, in tt events in both the semileptonic and
all-hadronic channels [240]. Top-quark jets were reconstructed either in the all-hadronic channel as large-𝑅
jets, using calorimeter energy deposits as input, or in the semileptonic channel as large-𝑅 reclustered jets,
using small-𝑅 particle-flow jets as input. In both cases, the anti-𝑘 𝑡 clustering algorithm with 𝑅 = 1 was
used, and top-quark candidate jets were required to have 𝑝 T > 350 GeV. In the all-hadronic channel,
in order to discriminate between tt events and QCD multijet background events, two large-𝑅 jets were
selected, with one required to be tagged by a top-quark-tagging algorithm [237] using several substructure
variables as input, and a second one without this requirement, to avoid biasing the measured observables.
Differential cross-sections were measured as a function of eight substructure variables, defined using

39
1/σtt dσtt/(d τ32 dp ) [1/GeV]
top
108 ATLAS (x104), 350 < p
T
[GeV] ≤ 500
7 top
10 s = 13 TeV, 140 fb -1 3
(x10 ), 500 < p [GeV] ≤ 550
T
top
106 l+jets, p > 350 GeV (x102), 550 < p
top
[GeV] ≤ 650

top
T T

T
105 Fiducial phase-space top
(x101), 650 < p [GeV] ≤ 750
T

104 top
(x100), 750 < p [GeV] ≤ 2000
NNLO QCD + NLO EW ATLAS 3
T

0.08 √ 10 PWG+PY8
8
Ctu = 0.5 s = 13 TeV, 139 fb −1
102
0.06 8
Ctu = 1.0 10
combination 1
0.04 single-lepton
10−1
dilepton
AtCt̄

0.02 10−2
10−3
0.00 10−4
10−5
−0.02
10−6
0 0.5 1
−0.04
< 0.5 [0.5,0.75] [0.75,1.0] [1.0,1.5] > 1.5 RC large-R jet τ32
mt t̄ [TeV]
(a) (b)

tt
Figure 14: (a) Unfolded differential charge asymmetry 𝐴C as a function of the invariant mass of the reconstructed
top-quark pair [219]. (b) Particle-level differential tt production cross-section as a function of 𝜏32 for several intervals
of top-quark transverse momentum, comparing data measurements with predictions [240].

only the charged components of the jets (see Figure 14 (b)). Double-differential distributions were also
measured, with two of these substructure variables (namely 𝜏32 and 𝐷 2 ) measured in bins of top-quark-jet
𝑝 T and mass. Results unfolded to particle level were compared with a number of MC predictions. The
nominal prediction from Powheg+Pythia 8 was found to properly model measurements of the broadness
and the two-body structure, but to poorly model variables sensitive to the three-body structure of the
top-quark jets, with the predicted jet substructure being more three-body-like than observed. On the other
hand, alternative predictions, such as those from aMC@NLO+Pythia 8 and Powheg+Herwig 7, as well
as those from Powheg+Herwig 8 with less FSR, were found to better model some of these distributions.
Overall, the measurement indicates the need to improve the models used to predict the substructure of
boosted top-quark jets.

7.4 Test of lepton-flavour universality

Top-quark production and decay can also be used as a tool to study generic properties of the SM, taking

advantage of the large tt sample available at 𝑠 = 13 TeV to perform precise measurements. One interesting
example is a test of lepton-flavour universality in top-quark decays. Lepton-flavour universality is the
SM axiom stating that the couplings of the electroweak gauge bosons to the three charged leptons are
equal. A large and unbiased sample of 𝑊 bosons can be selected in tt events and used to test this
axiom by measuring the ratio of 𝑊-boson decay rates into different charged leptons, ℓ, ℓ ′ = 𝑒, 𝜇, 𝜏:
𝑅𝑊 (ℓ/ℓ ′ ) = 𝐵(𝑊 → ℓ𝜈ℓ )/𝐵(𝑊 → ℓ ′ 𝜈ℓ ′ ).
The ratio 𝑅𝑊 (𝜇/𝜏) is particularly relevant because a measurement in 𝑒 + 𝑒 − → 𝑊 +𝑊 − events at LEP found
a deviation from unity: 𝑅𝑊 (𝜏/𝜇) = 1.070 ± 0.026 [241]. ATLAS measured this ratio by analysing the full
Run 2 dataset, identifying the 𝜏-leptons through their decay into muons [242]. Hence the key step in the
analysis is to distinguish these muons from those coming directly from 𝑊-boson decay (prompt muons).

40
This was achieved by using the distinct features of the 𝜏-lepton: its significant lifetime, giving rise to a
track with a large impact parameter, and its multibody decay into a muon and two neutrinos, leading to an
average muon transverse momentum that is lower than in direct 𝑊 → 𝜇𝜈 decays. The measurement was
performed in both the 𝑒𝜇 channel and the 𝜇𝜇 channel, with one lepton used to probe the 𝑊-boson branching
ratio and the other lepton to trigger the event and ensure a high-purity selection of top-pair events. The
fraction of muons that are prompt (𝑊 → 𝜇𝜈 𝜇 ) and the fraction that come from non-prompt 𝜏-lepton decay
(𝑊 → 𝜏𝜈 𝜏 → 𝜇𝜈 𝜇 𝜈 𝜏 ) were determined from a fit using templates that exploit shape differences between
those fractions’ distributions of both the muon’s transverse impact parameter |𝑑0 | (i.e. the distance of
closest approach of the muon track to the beam line in the 𝑥–𝑦 plane) and the muon’s transverse momentum
𝑝 T . The |𝑑0 | shape for the prompt muons was calibrated using 𝑍 → 𝜇𝜇 events. The normalisations of
the main backgrounds, i.e. prompt muons coming from 𝑍 → 𝜇𝜇 decays and non-prompt muons coming
from 𝑏- or 𝑐-hadron decays, were extracted using control regions. The ratio 𝑅𝑊 (𝜏/𝜇) was measured
from a two-dimensional profile-likelihood fit of the muon |𝑑0 | and 𝑝 T distributions, where the overall
tt normalisation was allowed to vary freely, leading to 𝑅𝑊 (𝜏/𝜇) = 0.992 ± 0.007(stat.) ± 0.011(syst.).
The result agrees with the SM expectation of lepton-flavour universality and constitutes the most precise
determination of this ratio. It demonstrates the ability of the LHC experiments to perform high-precision
measurements.
Moreover, the large tt event sample collected during Run 2 was used to precisely measure the ratio of
𝑊-boson decay rates into electrons and muons [243]. The analysis selected tt events in the 𝑒𝑒, 𝜇𝜇 and 𝑒𝜇
dilepton final states, relating the relative difference between channels to the ratio 𝑅𝑊 (𝜇/𝑒), and extracting
the ratio by means of a maximum-likelihood fit. The measurements in the three channels were obtained
with a technique similar to that used for the inclusive tt cross-section measurement in the 𝑒𝜇 channel (see
Section 3.1), with the dilepton invariant mass 𝑚 ℓℓ being exploited in the 𝑒𝑒 and 𝜇𝜇 channels to separate
signal events from the dominant 𝑍+jets background. In order to reduce the sensitivity to uncertainties in
the electron and muon identification efficiencies, the result was normalised to the square root of the ratio of
𝑍-boson decay rates to electron and muon pairs, 𝑅 𝑍 (𝜇𝜇/𝑒𝑒), which was measured simultaneously in data.
The precise value of 𝑅 𝑍 (𝜇𝜇/𝑒𝑒) from the LEP and SLD experiments (with a ∼0.3% uncertainty [193, 244])
was then used to extract the final result: 𝑅𝑊 (𝜇/𝑒) = 0.9995 ± 0.0022 (stat.) ± 0.0036 (syst.) ± 0.0014 (ext.),
where the last uncertainty refers to the external measurement of 𝑅 𝑍 (𝜇𝜇/𝑒𝑒). The SM assumption of
lepton-flavour universality in 𝑊-boson decays into electron–neutrino and muon–neutrino pairs was thus
confirmed at the 0.5% level.

8 Search for flavour-changing neutral currents in the top-quark sector

In the SM, flavour-changing neutral-current (FCNC) processes are forbidden at tree level and are strongly
suppressed in loops by the GIM mechanism [245]. In the top-quark sector, the rare decay channels in
which the top quark decays into a neutral boson (a Higgs or 𝑍 boson, a photon or a gluon) and a 𝑐-
or 𝑢-quark belong to this family and are predicted to have branching ratios (BRs) ranging from 10−12
to 10−17 [246]. In addition, the same interaction vertices can give rise to single-top FCNC production
processes, where a top quark (or antiquark) is produced in association with a neutral boson. These
processes are predicted to have negligible cross-sections in the SM, but a number of BSM scenarios predict
enhancements, increasing the FCNC top-quark decay BRs to 10−4 in some cases. These SM extensions
include the quark-singlet model [247], the two-Higgs-doublet model [248], the Minimal Supersymmetric
Standard Model (MSSM) [249], the MSSM with R-parity violation [250], models with warped extra
dimensions [251], and extended mirror-fermion models [252].

41
During LHC Run 2, ATLAS investigated essentially all the accessible channels for such FCNC searches.
Signal processes with a Higgs boson in the final state, i.e. involving the 𝑡𝐻𝑞 vertex with 𝑞 = 𝑐 or 𝑢, were
the most intensely studied. This kind of process gives rise to a number of possible final-state topologies,
depending on the Higgs boson’s decay mode, each requiring a dedicated analysis strategy and optimisation.
On the other hand, these analyses could benefit from the experience gained through the studies of the Higgs
boson production and decay modes, particularly in the context of tt𝐻 production. In addition, searches for
processes involving neutral gauge bosons, i.e. with a 𝑡𝑔𝑞, 𝑡𝛾𝑞 or 𝑡𝑍𝑞 vertex, were performed. Both FCNC
top-quark decays and single-top FCNC production processes were investigated, by eventually adopting a
global approach as suggested in Ref. [253], relating FCNC decay branching ratios to FCNC production
cross-sections via EFT operators.

8.1 Searches for top-quark FCNC processes involving a Higgs boson

The first published Run 2 searches for FCNC top-quark decays into a 𝑐- or 𝑢-quark and a Higgs boson,
separately focusing on the 𝐻 → 𝛾𝛾 [254], 𝐻 → 𝑊 +𝑊 − [255] and 𝐻 → 𝑏 𝑏/𝜏 ¯ + 𝜏 − [256] decay channels,
were all based on the dataset collected in 2015 and 2016, corresponding to an integrated luminosity of
36 fb−1 . All three analyses looked for tt events where one of the two top quarks decays into 𝑏𝑊 and the
other via a FCNC into 𝑐𝐻 or 𝑢𝐻. With the larger dataset available at the end of Run 2, updated results
were released for most of these main Higgs boson decay channels. These new analyses considered FCNC
processes not only in the top quark’s decay but also in its production, by searching for events with a single
top quark plus a Higgs boson (𝑡𝐻) that were initiated by a 𝑢- or 𝑐-quark. The results were reported in
dedicated papers for the 𝐻 → 𝛾𝛾 [257], 𝐻 → 𝑊 +𝑊 − [258] and 𝐻 → 𝜏 + 𝜏 − [259] channels, while for
Higgs boson decay into a pair of bottom quarks, ATLAS took advantage of the full Run 2 dataset to perform
a more general search for FCNC top-quark decays into a generic scalar boson and an up-type quark [260],
in a new-scalar-boson mass range from 20 to 160 GeV, thus including the SM Higgs boson.
Searches using the Higgs boson’s diphoton decay channel [254, 257] looked for final states with a pair of
energetic isolated photons with an invariant mass peaking at the Higgs boson mass, a top quark (decaying
either leptonically or hadronically), and possibly a hadronic jet from a light quark or 𝑐-quark. In the
partial-dataset analysis, events were separated into categories depending on the presence of a charged
lepton and on the compatibility of the reconstructed final states with tt decays. With the full dataset, the
analysis was further improved by using a dedicated charm-tagging algorithm to split the event categories
more finely, and using a BDT-based selection for their definition. The result was then extracted by fitting
the diphoton invariant-mass spectra with a resonant signal function centred around the Higgs boson mass
and a background function, with constraints mainly from 𝑚 𝛾𝛾 bands on either side of each signal region.
The analyses reported in Ref. [255] and Ref. [258] focused on multilepton final states, targeting Higgs
boson decays giving rise to at least one electron or muon, such as 𝐻 → 𝑊𝑊 with at least one of the two
𝑊 bosons decaying leptonically, but also 𝐻 → 𝑍 𝑍 or 𝐻 → 𝜏𝜏.6 When the SM leg of the top-quark or
top-antiquark decay gives rise to a charged lepton as well, these processes can produce events with two
same-sign electrons or muons, or even three-lepton events, for which the background from SM processes is
significantly smaller. This analysis is characterised by final states similar to those for the tt𝐻 measurement
in the multilepton channel [261] and therefore shares with it a good fraction of the event selection, event

6
The analysis vetoes events with at least one reconstructed hadronic 𝜏-lepton decay, so as to avoid statistical overlap with the
dedicated search described before. Therefore, the analysis is sensitive to 𝐻 → 𝜏𝜏 only when both 𝜏-leptons decay to an electron
or muon.

42
categorisation and background estimation. However, dedicated multivariate discriminants were employed
in the signal regions to separate the SM processes from the FCNC signal.
The 𝐻 → 𝜏 + 𝜏 − analyses [256, 259] categorised events according to the numbers of reconstructed hadronic
𝜏-lepton decays (𝜏had ) and light charged leptons from leptonic 𝜏-lepton decays (𝜏lep ), to separately target
𝜏had 𝜏had and 𝜏lep 𝜏had events.7 While the partial-dataset analysis focused on hadronic decay in the SM leg of
the tt process, the full-dataset analysis also included single-top topologies and dedicated event categories
targeting tt events with leptonic top-quark decay. Events in the various categories were processed through
dedicated kinematic reconstruction algorithms, aiming to deduce the four-momenta of the invisible decay
products for each 𝜏-lepton decay. Finally, dedicated BDT discriminants were built in each category,
separately for the 𝑡𝑐𝐻 and 𝑡𝑢𝐻 couplings, using as input a number of topological and kinematic final-state
variables, including ones with values coming from the kinematic reconstruction.
The 𝐻 → 𝑏 𝑏¯ analyses [256, 260] were also heavily based on studies and analysis techniques developed in
the context of the searches for the tt𝐻 process, in this case in the 𝑏 𝑏¯ decay channel during Run 1 [262]
and Run 2 [263]. They relied on the presence of a high-momentum electron or muon in the SM leg of
the tt decay and categorised selected events according to jet and 𝑏-tagged-jet multiplicities (4–6 jets, and
2–4 𝑏-tagged jets), with the goal of retaining events with 𝑏-quark jets that fail the 𝑏-tagging requirements,
𝑐-quark jets that are 𝑏-tagged, jets that are outside the acceptance or produced from hard QCD radiation,
as well as including background-enriched data samples, useful for placing in situ constraints on some
of the associated systematic uncertainties. In the first analysis, a dedicated likelihood discriminant was
constructed in each of the categories to assess the compatibility of the observed final states with signal
processes by considering invariant masses of jet and lepton combinations. The second analysis implemented
a discriminating variable based on a NN in each signal region for each scalar mass hypothesis.
None of these searches for FCNC top-quark decays observed any significant excess over the SM expectations,
with the largest deviation from the background-only hypothesis found in the case of the full-dataset analysis
in the 𝜏𝜏 channel, yielding a 2.3𝜎 significance. Exclusion limits were then set on the branching ratios for
the 𝑡 → 𝐻𝑐 and 𝑡 → 𝐻𝑢 processes, from each of the individual channels as well as from their combinations,
and reported in Ref. [256] and Ref. [258] for the searches based on 36 fb−1 and 140 fb−1 respectively. For
the partial-dataset analyses, the exclusion upper limits at 95% CL ranged from 1.5 × 10−3 to 5 × 10−3
depending on the channel, with combined limits of around 1 × 10−3 for both up-type-quark flavours. The
corresponding 95% CL exclusion limits obtained by combining the full-dataset analyses instead yielded:

𝐵(𝑡 → 𝑢𝐻) < 2.6 × 10−4 and 𝐵(𝑡 → 𝑐𝐻) < 3.4 × 10−4 .

These results are shown in Figure 15, together with those of the individual input analyses.

8.2 Searches for top-quark FCNC processes involving neutral gauge bosons

With the full Run 2 dataset, ATLAS also performed searches for top-quark FCNC couplings to the gluon,
the photon and the 𝑍 boson. In particular, limits on the top-quark FCNC interactions involving a gluon
were set by searching for the single-top-quark production processes 𝑢𝑔 → 𝑡 and 𝑐𝑔 → 𝑡 [264]. This
analysis relied on the selection of a sample of events compatible with the production of a single top quark,
without the accompanying jet characterising the SM single-top-quark production channels (the 𝑡-, 𝑠- and
𝑡𝑊-channels, see Section 4). The event selection was defined by requiring the presence of an electron or

7
Higgs boson decays with two 𝜏lep were targeted by the multilepton search.

43
ATLAS s = 13 TeV, 140 fb 1 ATLAS s = 13 TeV, 140 fb 1

H bb H bb

H H

H H
(t Hc) = 0 (t Hu) = 0
H VV * Observed H VV * Observed
Expected ±1 Expected ±1
Expected ±2 Expected ±2
Combination Combination
0 5 10 15 20 0 5 10 15 20
95% CL limit on (t Hu) × 104 95% CL limit on (t Hc) × 104
(a) (b)

Figure 15: 95% CL upper limits on (a) 𝐵(𝑡 → 𝐻𝑢) and (b) 𝐵(𝑡 → 𝐻𝑐) for the individual ATLAS searches based on
the full Run 2 dataset, as well as their combination (assuming (a) 𝐵(𝑡 → 𝐻𝑐) = 0 and (b) 𝐵(𝑡 → 𝐻𝑢) = 0. The
observed limits (solid lines) are compared with the expected (median) limits under the background-only hypothesis
(dotted lines). The surrounding shaded bands correspond to the 68% and 95% CL intervals around the expected
limits, denoted by ±1𝜎 and ±2𝜎, respectively. Figures taken from Ref. [258].

muon, large missing transverse momentum, and exactly one central jet passing a particularly tight 𝑏-tagging
requirement corresponding to a nominal 30% efficiency for 𝑏-quark-initiated jets and a rejection rate of
around 900 (30 000) for 𝑐-quark-initiated (light-quark-initiated) jets. Two NN discriminants were built to
discriminate between signal events and background events, including the irreducible background from
𝑡-channel single-top-quark production, based on a number of kinematic properties of final-state leptons,
jets and the missing transverse momentum. One of these discriminants was trained specifically to isolate
FCNC top-quark-production events initiated by a valence quark and a gluon, namely 𝑢𝑔 → 𝑡, characterised
by kinematic properties different from those of events from sea-quark-initiated processes such as 𝑢𝑔¯ → 𝑡¯,
𝑐𝑔 → 𝑡 and 𝑐𝑔¯ → 𝑡¯. No significant excess of events compatible with FCNC-production kinematics was
observed, and exclusion limits were set on the production cross-sections of the two signal processes. These
were then turned into 95% CL limits on top-quark FCNC-decay branching ratios:

𝐵(𝑡 → 𝑢𝑔) < 0.61 × 10−4 and 𝐵(𝑡 → 𝑐𝑔) < 3.7 × 10−4 .

Limits on the 𝑐-quark-initiated process are significantly weaker than those on the 𝑢-quark-initiated process.
This is mainly due to the predicted cross-section being lower for the 𝑐𝑔 → 𝑡 process because it can only
be initiated by sea quarks, which typically have a lower momentum fraction than the valance quarks
contributing to the 𝑢𝑔 → 𝑡 process. However, 𝑐-quark-initiated processes are still phenomenologically
relevant, especially for two-Higgs-doublet models, which predict stronger FCNC couplings between top
and charm quarks.
FCNC coupling between a top quark and a photon was also investigated deeply using Run 2 data. A search
for FCNCs in the production of a top quark with a photon, 𝑔𝑢 → 𝑡𝛾 or 𝑔𝑐 → 𝑡𝛾, was performed using
80 fb−1 of data [265]. Events with an energetic isolated photon and the typical final-state signature of
a leptonically decaying top quark (an isolated electron or muon, a 𝑏-tagged jet and missing transverse
momentum), and possibly additional jets, were selected, and a NN based on kinematic variables was

44
used to discriminate between signal and SM background processes. The analysis was updated to use the
full Run 2 dataset [266], and extended to cover both the production and decay processes, 𝑔𝑞 → 𝑡𝛾 and
𝑡 → 𝛾𝑞 (in tt events) with 𝑞 being an 𝑢- or 𝑐-quark, as target signal. As done in previous analyses, the
search for a FCNC in top-quark decay was limited to tt-production events, so the two signal processes
were characterized by similar event topologies. A multi-class DNN discriminant was built to classify
events as coming from one of the two signal processes (FCNC in production or FCNC in decay) or as
background events. Dedicated control samples were selected in order to control the normalisation of the
main background processes, tt𝛾 and 𝑊 𝛾, while data-driven scale factors were obtained to correct the
simulation of the background component induced by the misidentification of a hadron or an electron as a
photon. The DNN was trained separately for 𝑡𝛾𝑢 and 𝑡𝛾𝑐 couplings, because of the different kinematics
for the 𝑔𝑢 → 𝑡𝛾 and 𝑔𝑐 → 𝑡𝛾 processes, induced by the differences between the up- and charm-quark
PDFs, and the different 𝑏-tagging probabilities for tt events with a 𝑡 → 𝛾𝑢 decay or 𝑡 → 𝛾𝑐 decay. The
final results of the two searches, extracted with a binned profile-likelihood fit to the DNN discriminant in
the signal region and to the photon 𝑝 T distributions in the background control regions, didn’t show any
significant excess over the SM expectations, allowing exclusion limits to be set on the signal production
cross-sections and decay branching ratios, separately for right-handed (RH) and left-handed (LH) 𝑡𝛾𝑞
couplings, as well as on the relative effective coupling constants [130, 253, 267, 268]. The exclusion limits
on the branching ratios range from 2.8 × 10−5 to 22 × 10−5 for the former analysis, and from 0.85 × 10−5
to 4.5 × 10−5 for the latter analysis, depending on the light quark’s flavour and on the coupling chirality,
with the full-Run 2 results improving on the previous limits by factors of 3.3 to 5.4. For a LH 𝑡𝛾𝑞 coupling,
the 95% CL exclusion limits obtained are:
𝐵(𝑡 → 𝑢𝑍) < 0.85 × 10−5 and 𝐵(𝑡 → 𝑐𝑍) < 4.2 × 10−5 .

Finally, FCNC coupling between a top quark and a 𝑍 boson was studied in Ref. [269] and Ref. [270].
The former analysis, based on 36 fb−1 , relied only on the top-quark-pair events, with one top quark
decaying through the 𝑡 → 𝑍𝑞 channel, while the latter analysis, based on the full Run 2 dataset, included
single-top-quark production via 𝑔𝑞 → 𝑡𝑍 (with 𝑞 = 𝑢, 𝑐) as a signal process. Since the final-state topologies
are similar to those investigated in SM tt 𝑍 and 𝑡𝑍 associated-production processes (see Section 5), the
analysis naturally relied on the selection of events with three charged leptons, two of them coming from the
𝑍-boson decay and the other from the top-quark decay, plus a 𝑏-tagged jet and possibly additional jets.
In the first analysis, the top quarks were kinematically reconstructed from the final-state leptons and jets,
using 𝜒2 minimisation across all the possible jet and lepton permutations in each event. The result of the
𝜒2 minimisation was used as a discriminating variable in the three-lepton signal region, and a number of
control regions were included in the analysis to specifically target the tt 𝑍, 𝑊 𝑍, 𝑍 𝑍 and non-prompt-lepton
backgrounds. For the full-Run 2 search, multivariate discriminants were introduced to improve the signal
sensitivity beyond the gain from the larger analysed dataset. In particular, after separating selected events
into two orthogonal signal regions, individually optimised for single-top-quark and tt production signal
processes, gradient boosted decision trees (GBDTs) were trained to discriminate between FCNC signals
and SM backgrounds, using the outputs of the kinematic reconstruction as input variables. In the signal
region dedicated to single-top-quark production, two separate GBDTs were built to separately target 𝑢- and
𝑐-quark-initiated processes. Finally, the presence of signal events from RH and LH FCNC couplings was
tested separately. The full-Run 2 analysis improved the sensitivity to the branching ratios 𝐵(𝑡 → 𝑍𝑞) by a
factor of two, and set the following 95% CL exclusion limits for the LH coupling scenario:
𝐵(𝑡 → 𝑢𝛾) < 6.2 × 10−5 and 𝐵(𝑡 → 𝑐𝛾) < 1.3 × 10−4 .
Similar limits were set for the RH coupling hypothesis (6.6 × 10−5 and 1.2 × 10−4 for the 𝑡𝑍𝑢 and 𝑡𝑍𝑐
couplings, respectively).

45
ATLAS CLs exclusion upper limits on FCNC processes
expected exclusion ± 1 σ
s = 13 TeV, 139-140 fb-1
observed exclusion

t → ug
Eur. Phys. J. C 82 (2022) 334
t → cg

t → uγ
Phys. Lett. B 842 (2023) 137379
t → cγ

t → uZ
Phys. Rev. D 108 (2023) 032019
t → cZ

t → uH
arXiv:2404.02123
t → cH

10− 5 10− 4 10− 3


95% CL limit on BR

Figure 16: Summary of the 95% CL observed and expected limits on the branching ratios of the top-quark decays via
flavour-changing neutral currents (FCNC) to a quark and a neutral boson 𝑡 → 𝑞𝑋 (𝑋 = 𝑔, 𝑍, 𝛾 or 𝐻; 𝑞 = 𝑢 or 𝑐)
set by the ATLAS Collaboration using its full Run 2 dataset. The hatched area represents the ±1𝜎 band around
the expected exclusion limit. It is not shown in the case of the 𝑡𝑔𝑞 coupling, as the numbers are not quoted in the
publication. The quoted upper limits refer to the left-handed chirality hypothesis, and expected limits are computed
assuming zero Standard Model contribution to the FCNC processes.

8.3 Summary of FCNC process constraints

A summary of the exclusion limits set by the ATLAS Run 2 searches for FCNCs in the top-quark sector is
shown in Figure 16. As can be seen, exclusion limits on photon-mediated FCNC processes are the strongest,
followed by those on transitions mediated by 𝑍 bosons and gluons, while weaker limits are set on processes
involving Higgs bosons. In all cases, the sensitivity is stronger for 𝑡 → 𝑢 than for 𝑡 → 𝑐 transitions. All
of these stringent limits on top-quark FCNC interactions were also interpreted as constraints on Wilson
coefficients in an EFT framework, as detailed in Section 9.

9 Limits on Wilson coefficients within effective field theory

New physics effects from BSM theories characterised by a mass scale higher than the energy accessible
at the LHC can be investigated in the framework of the Standard Model effective field theory (SMEFT),
without specifying a particular BSM model. The SMEFT extension of the SM Lagrangian is:
∑︁ 𝐶
𝑖
LSMEFT = LSM + 𝑂 𝑖 + ...,
𝑖 Λ2

where LSM represents the SM Lagrangian, 𝑂 𝑖 are effective dimension-six operators and 𝐶𝑖 the corresponding
Wilson coefficients. Here, Λ corresponds to the cut-off scale of the effective theory and is usually set to

46
Λ = 1 TeV. Dimension-five operators induce baryon- and lepton-number-violating terms and are therefore
usually ignored. Higher-dimension operators are Λ-suppressed and usually also ignored. The operators
can be expressed in different bases. The Warsaw basis [267] is chosen for the results presented here.
Strong tt production is sensitive to the operator 𝑂 𝑡𝐺 , which modifies the top–gluon vertex and primarily
(1)
changes the overall production rate, and to the 14 operators characterising four-quark interactions: 𝑂 𝑡𝑢 ,
(8) (1) (8) (1) (8) (1) (8) (1) (8) (1,1) (3,1) (1,8) (3,8) (8)
𝑂 𝑡𝑢 , 𝑂 𝑡 𝑑 , 𝑂 𝑡 𝑑 , 𝑂 𝑡𝑞 , 𝑂 𝑡𝑞 , 𝑂 𝑄𝑢 , 𝑂 𝑄𝑢 , 𝑂 𝑄𝑑 , 𝑂 𝑄𝑑 , 𝑂 𝑄𝑞 , 𝑂 𝑄𝑞 , 𝑂 𝑄𝑞 , 𝑂 𝑄𝑞 . In particular, 𝑂 𝑡𝑞 can
significantly alter differential tt cross-sections in the high-energy tails. The 𝑡-channel single-top-quark
cross-section is particularly sensitive to the operator 𝑂 3,1 𝑄𝑞 . The 𝑊𝑡𝑏 vertex could also be altered by the
operators 𝑂 3𝜙𝑄 and 𝑂 𝑡𝑊 . The ttZ cross-section is sensitive to the 14 four-quark operators as well as to the
(1) (3)
operators modifying the top–boson vertices (𝑂 𝑡𝑊 , 𝑂 𝑡 𝐵 , 𝑂 𝑡𝐺 , 𝑂 𝐻𝑄 , 𝑂 𝐻𝑄 , 𝑂 𝐻𝑡 ). The tttt cross-section is
specifically sensitive to the four heavy-flavour fermion operators (𝑂 1𝑄𝑄 , 𝑂 1𝑄𝑡 , 𝑂 1𝑡𝑡 , 𝑂 8𝑄𝑡 ). The tt charge
tt
asymmetry 𝐴C is also sensitive to the 14 four-quark operators but to different linear combinations of these,
compared to tt production, which helps to disentangle the contributions when global fits are performed. It
tt
is also sensitive to the top–gluon operator 𝑂 𝑡𝐺 . The differential 𝐴C measurements lead to tighter limits
tt
than obtained from the inclusive 𝐴C measurement, so the differential constraints are reported here. The
energy asymmetry 𝐴𝐸 (𝜃 𝑗 ) is particularly sensitive to the chirality and the colour charges of the involved
(1) (8) (1) (8) (1,1) (1,8)
quark fields, and therefore to six specific four-quark operators (𝑂 𝑡𝑢 , 𝑂 𝑡𝑢 , 𝑂 𝑡𝑞 , 𝑂 𝑡𝑞 , 𝑂 𝑄𝑞 , 𝑂 𝑄𝑞 ).
Finally, limits on top-quark FCNC couplings can be interpreted as constraints on the Wilson coefficients
of the EFT operators inducing tree-level FCNC transitions [253]. Exclusion limits on top–Higgs FCNC
couplings can be translated into constraints on the coefficients 𝐶𝑢13𝜙 and 𝐶𝑢23𝜙 , those on the top–gluon
13 23
coupling into constraints on 𝐶𝑢𝐺 and 𝐶𝑢𝐺 , while limits on both photon- and 𝑍-boson-induced FCNC
𝑖3 𝑖3
interactions translate into constraints on linear combinations of the coefficients 𝐶𝑢𝐵 and 𝐶𝑢𝑊 , with
𝑖 = 1, 2.
The 95% CL limits on Wilson coefficients extracted for Λ = 1 TeV from the measurements and searches
described in this paper, assuming one coefficient at a time to be non-zero, are summarised in Tables 1,
2 and 3. These bounds are obtained using a parameterisation that includes the dimension-six-squared
terms.

47
Table 1: Summary of the 95% CL intervals for top–boson Wilson coefficients within SMEFT set by the top-quark
Run 2 measurements and searches. These limits are set for Λ = 1 TeV and when one coefficient at a time is assumed
to be non-zero. ℜ𝐶𝑡𝑊 and ℑ𝐶𝑡𝑊 represent the real and imaginary parts of the complex 𝐶𝑡𝑊 coefficient.

Top–boson coefficient Limits Input measurement Reference


𝐶 3𝜙𝑄 [–0.87, 1.42] 𝑡-channel cross-section [109]
ℜ𝐶𝑡𝑊 [–0.9, 1.4] 𝑡-channel polarisation [110]
[–0.84, 1.0] tt 𝑍 cross-section [147]
ℑ𝐶𝑡𝑊 [–0.8, 0.2] 𝑡-channel polarisation [110]
[–1.0, 1.0] tt 𝑍 cross-section [147]
(1)
𝐶 𝐻𝑄 [–1.4, 0.84] tt 𝑍 cross-section [147]
(3)
𝐶 𝐻𝑄 [–0.95, 2.0] tt 𝑍 cross-section [147]
𝐶 𝐻𝑡 [–2.2, 1.6] tt 𝑍 cross-section [147]
ℜ𝐶𝑡 𝐵 [–1.7, 1.6] tt 𝑍 cross-section [147]
ℑ𝐶𝑡 𝐵 [–1.9, 1.9] tt 𝑍 cross-section [147]
ℜ𝐶𝑡𝐺 [–0.23, 0.34] tt 𝑍 cross-section [147]
[–0.75, 0.14] tt charge asymmetry [219]
[–0.52, 0.15] tt cross-section [97]
ℑ𝐶𝑡𝐺 [–0.32, 0.33] tt 𝑍 cross-section [147]

48
Table 2: Summary of the 95% CL intervals for four-quark Wilson coefficients within SMEFT set by the top-quark
Run 2 measurements and searches. These limits are set for Λ = 1 TeV and when one coefficient at a time is assumed
to be non-zero.

Four-quark coefficient Limits Input measurement Reference


3,1
𝐶𝑄𝑞 [–0.37, 0.06] 𝑡-channel cross-section [109]
[–0.34, 0.23] tt 𝑍 cross-section [147]
[–0.34, 0.43] tt charge asymmetry [219]
(1)
𝐶𝑄𝑑 [–0.47, 0.46] tt 𝑍 cross-section [147]
[–0.39, 0.42] tt charge asymmetry [219]
(8)
𝐶𝑄𝑑 [–1.8, 0.62] tt 𝑍 cross-section [147]
[–0.96, 1.37] tt charge asymmetry [219]
(1,1)
𝐶𝑄𝑞 [–0.29, 0.29] tt 𝑍 cross-section [147]
[–0.52, 0.28] tt charge asymmetry [219]
[–0.65, 0.67] tt energy asymmetry [223]
(1,8)
𝐶𝑄𝑞 [–1.2, 0.080] tt 𝑍 cross-section [147]
[–0.25, 0.53] tt charge asymmetry [219]
[–1.72, 2.10] tt energy asymmetry [223]
(3,8)
𝐶𝑄𝑞 [–0.68, 0.54] tt 𝑍 cross-section [147]
[–1.23, 0.31] tt charge asymmetry [219]
(1)
𝐶𝑄𝑢 [–0.47, 0.43] tt 𝑍 cross-section [147]
[–0.31, 0.23] tt charge asymmetry [219]
(8)
𝐶𝑄𝑢 [–1.6, 0.36] tt 𝑍 cross-section [147]
[–1.78, 0.27] tt charge asymmetry [219]
(1)
𝐶𝑡𝑞 [–0.42, 0.36] tt 𝑍 cross-section [147]
[–0.20, 0.22] tt charge asymmetry [219]
[–0.69, 0.75] tt energy asymmetry [223]
(8)
𝐶𝑡𝑞 [–1.5, 0.22] tt 𝑍 cross-section [147]
[–0.51, 0.58] tt charge asymmetry [219]
[–2.01, 1.43] tt energy asymmetry [223]
[–0.64, 0.12] tt cross-section [97]
𝐶𝑡(1)
𝑑 [–0.56, 0.56] tt 𝑍 cross-section [147]
[–0.60, 0.84] tt charge asymmetry [219]
𝐶𝑡(8)
𝑑 [–2.5, 1.2] tt 𝑍 cross-section [147]
[–1.62, 1.21] tt charge asymmetry [219]
(1)
𝐶𝑡𝑢 [–0.66, 0.64] tt 𝑍 cross-section [147]
[–0.70, 0.31] tt charge asymmetry [219]
[–0.78, 0.81] tt energy asymmetry [223]
(8)
𝐶𝑡𝑢 [–2.2, 0.38] tt 𝑍 cross-section [147]
[–0.45, 0.82] tt charge asymmetry [219]
[–1.71, 1.56] tt energy asymmetry [223]
1
𝐶𝑄𝑄 [–3.5, 4.1] tttt cross-section [176]
1
𝐶𝑄𝑡 [–3.5, 3.0] tttt cross-section [176]
49
𝐶𝑡𝑡1 [–1.7, 1.9] tttt cross-section [176]
8
𝐶𝑄𝑡 [–6.2, 6.9] tttt cross-section [176]
Table 3: Summary of the 95% CL intervals for top–FCNC Wilson coefficients within SMEFT set by the top-quark
Run 2 measurements and searches. These limits are set for Λ = 1 TeV and when one coefficient at a time is assumed
to be non-zero.

FCNC coefficient Limits Input search Reference


|𝐶𝑢13,31
𝜙 | < 0.68 𝑡 → 𝐻𝑞 [258]
|𝐶𝑢23,32
𝜙 | < 0.78 𝑡 → 𝐻𝑞 [258]
13
|𝐶𝑢𝐺 | < 0.057 𝑡 → 𝑔𝑞 [264]
23
|𝐶𝑢𝐺 | < 0.14 𝑡 → 𝑔𝑞 [264]
13 ∗ 13 ∗
|𝐶𝑢𝑊 + 𝐶𝑢𝐵 | < 0.103 𝑡 → 𝛾𝑞 [266]
31 31
|𝐶𝑢𝑊 + 𝐶𝑢𝐵 | < 0.123 𝑡 → 𝛾𝑞 [266]
23 ∗ 23 ∗
|𝐶𝑢𝑊 + 𝐶𝑢𝐵 | < 0.227 𝑡 → 𝛾𝑞 [266]
32 32
|𝐶𝑢𝑊 + 𝐶𝑢𝐵 | < 0.235 𝑡 → 𝛾𝑞 [266]
13 ∗ 13 ∗
|𝐶𝑢𝑊 + 𝐶𝑢𝐵 | < 0.15 𝑡 → 𝑍𝑞 [270]
31 31
|𝐶𝑢𝑊 + 𝐶𝑢𝐵 | < 0.16 𝑡 → 𝑍𝑞 [270]
23 ∗ 23 ∗
|𝐶𝑢𝑊 + 𝐶𝑢𝐵 | < 0.22 𝑡 → 𝑍𝑞 [270]
32 32
|𝐶𝑢𝑊 + 𝐶𝑢𝐵 | < 0.21 𝑡 → 𝑍𝑞 [270]

50
10 Conclusion

With the unprecedented 𝑝 𝑝 collision dataset collected during LHC Run 2 at a centre-of-mass energy of
13 TeV, the ATLAS Collaboration brought top-quark physics to the next level. Precise measurements
challenge the most accurate theoretical predictions. The inclusive tt cross-section measurement has reached
a precision of 1.8%, exceeding the precision of the NNLO+NNLL prediction. For the most abundant
top-quark production processes, as well as for a number of top-quark associated-production processes,
the focus is moving toward the measurement of differential distributions. Along with this shift of focus,
rarer and rarer production processes can be observed and studied for the first time. Using Run 2 data, the
𝑡𝑍𝑞, 𝑡𝑞𝛾 and tttt processes have now been observed. Measurements of the ttZ, ttW and tt𝛾 cross-sections
achieving a precision of around 10% are turning into precision measurements, directly probing the top
quark’s couplings to gauge bosons. New methods for the precise determination of the top-quark mass
have been adopted, alongside those already used in Run 1. Through the identification and reconstruction
of top-quark decay products, several interesting measurements have been performed for the first time or
with unprecedented accuracy, including the precise determination of the 𝑊-boson helicity fractions, the
top quark’s polarisation and tt spin correlations, as well as results showing strong evidence for charge
asymmetry in top-quark pair production and the observation of quantum entanglement in tt events. Finally,
exclusion limits for exotic or anomalous top-quark production and decay mechanisms have been improved
significantly, challenging several new physics-model predictions.
In many cases, statistical uncertainties have become subdominant, putting more emphasis on systematic
uncertainties. Systematic uncertainties have been reduced markedly with respect to Run 1. Due to precise
understanding of detector-related systematic effects, especially those which usually dominate top-quark
physics measurements (such as uncertainties in the jet energy scale and 𝑏-tagging efficiency), the dominant
uncertainty in many of these measurements is related to the modelling of top-quark-pair production in
MC simulation. This motivated several of the measurements discussed in this report that have been or
will be used to tune such models, further reducing the impact of systematic uncertainties. Nevertheless,
some of the measurements still suffer from significant statistical uncertainty, especially those involving
rare processes, such as four- and three-top-quark production, some of those trying to precisely measure
small SM effects, e.g. the tt charge asymmetry, and those sensitive to BSM effects, such as predicted in
differential cross-section measurements in the tails of kinematic distributions.
While the Run 2 analyses of top-quark production and properties are not over yet, LHC Run 3 is bringing
particle physics to a new energy with 𝑝 𝑝 collisions at 13.6 TeV. Although the new energy increases the tt
cross-section by only 11%, it can have larger effects, for instance increasing the tttt cross-section by 20%.
There is no doubt that the higher energy will bring further advances in our understanding of the top quark,
and perhaps new physics discoveries.

Acknowledgements

We thank CERN for the very successful operation of the LHC and its injectors, as well as the support staff
at CERN and at our institutions worldwide without whom ATLAS could not be operated efficiently.
The crucial computing support from all WLCG partners is acknowledged gratefully, in particular from
CERN, the ATLAS Tier-1 facilities at TRIUMF/SFU (Canada), NDGF (Denmark, Norway, Sweden),
CC-IN2P3 (France), KIT/GridKA (Germany), INFN-CNAF (Italy), NL-T1 (Netherlands), PIC (Spain),

51
RAL (UK) and BNL (USA), the Tier-2 facilities worldwide and large non-WLCG resource providers.
Major contributors of computing resources are listed in Ref. [271].
We gratefully acknowledge the support of ANPCyT, Argentina; YerPhI, Armenia; ARC, Australia; BMWFW
and FWF, Austria; ANAS, Azerbaijan; CNPq and FAPESP, Brazil; NSERC, NRC and CFI, Canada; CERN;
ANID, Chile; CAS, MOST and NSFC, China; Minciencias, Colombia; MEYS CR, Czech Republic; DNRF
and DNSRC, Denmark; IN2P3-CNRS and CEA-DRF/IRFU, France; SRNSFG, Georgia; BMBF, HGF and
MPG, Germany; GSRI, Greece; RGC and Hong Kong SAR, China; ISF and Benoziyo Center, Israel; INFN,
Italy; MEXT and JSPS, Japan; CNRST, Morocco; NWO, Netherlands; RCN, Norway; MEiN, Poland; FCT,
Portugal; MNE/IFA, Romania; MESTD, Serbia; MSSR, Slovakia; ARRS and MIZŠ, Slovenia; DSI/NRF,
South Africa; MICINN, Spain; SRC and Wallenberg Foundation, Sweden; SERI, SNSF and Cantons of
Bern and Geneva, Switzerland; MOST, Taipei; TENMAK, Türkiye; STFC, United Kingdom; DOE and
NSF, United States of America.
Individual groups and members have received support from BCKDF, CANARIE, CRC and DRAC, Canada;
PRIMUS 21/SCI/017, CERN-CZ and FORTE, Czech Republic; COST, ERC, ERDF, Horizon 2020,
ICSC-NextGenerationEU and Marie Skłodowska-Curie Actions, European Union; Investissements d’Avenir
Labex, Investissements d’Avenir Idex and ANR, France; DFG and AvH Foundation, Germany; Herakleitos,
Thales and Aristeia programmes co-financed by EU-ESF and the Greek NSRF, Greece; BSF-NSF and
MINERVA, Israel; Norwegian Financial Mechanism 2014-2021, Norway; NCN and NAWA, Poland; La
Caixa Banking Foundation, CERCA Programme Generalitat de Catalunya and PROMETEO and GenT
Programmes Generalitat Valenciana, Spain; Göran Gustafssons Stiftelse, Sweden; The Royal Society and
Leverhulme Trust, United Kingdom.
In addition, individual members wish to acknowledge support from CERN: European Organization for
Nuclear Research (CERN PJAS); Chile: Agencia Nacional de Investigación y Desarrollo (FONDECYT
1190886, FONDECYT 1210400, FONDECYT 1230987); China: National Natural Science Foundation
of China (NSFC - 12175119, NSFC 12275265); Czech Republic: Czech Science Foundation (GACR
- 24-11373S), Ministry of Education Youth and Sports (FORTE CZ.02.01.01/00/22_008/0004632);
European Union: European Research Council (ERC - 948254, ERC 101089007), Horizon 2020 Framework
Programme (MUCCA - CHIST-ERA-19-XAI-00), Italian Center for High Performance Computing, Big Data
and Quantum Computing (ICSC, NextGenerationEU); France: Agence Nationale de la Recherche (ANR-
20-CE31-0013, ANR-21-CE31-0013, ANR-21-CE31-0022), Investissements d’Avenir Labex (ANR-11-
LABX-0012); Germany: Baden-Württemberg Stiftung (BW Stiftung-Postdoc Eliteprogramme), Deutsche
Forschungsgemeinschaft (DFG - 469666862, DFG - CR 312/5-2); Italy: Istituto Nazionale di Fisica
Nucleare (ICSC, NextGenerationEU); Japan: Japan Society for the Promotion of Science (JSPS KAKENHI
22H01227, JSPS KAKENHI 22KK0227, JSPS KAKENHI JP21H05085, JSPS KAKENHI JP22H04944);
Netherlands: Netherlands Organisation for Scientific Research (NWO Veni 2020 - VI.Veni.202.179);
Norway: Research Council of Norway (RCN-314472); Poland: Polish National Agency for Academic
Exchange (PPN/PPO/2020/1/00002/U/00001), Polish National Science Centre (NCN 2021/42/E/ST2/00350,
NCN OPUS nr 2022/47/B/ST2/03059, NCN UMO-2019/34/E/ST2/00393, UMO-2020/37/B/ST2/01043,
UMO-2021/40/C/ST2/00187, UMO-2022/47/O/ST2/00148); Slovenia: Slovenian Research Agency (ARIS
grant J1-3010); Spain: Generalitat Valenciana (Artemisa, FEDER, IDIFEDER/2018/048), Ministry of
Science and Innovation (RYC2019-028510-I, RYC2020-030254-I), PROMETEO and GenT Programmes
Generalitat Valenciana (CIDEGENT/2019/023, CIDEGENT/2019/027); Sweden: Swedish Research
Council (VR 2022-03845, VR 2022-04683, VR 2023-03403), Knut and Alice Wallenberg Foundation
(KAW 2018.0157, KAW 2019.0447, KAW 2022.0358); Switzerland: Swiss National Science Foundation

52
(SNSF - PCEFP2_194658); United Kingdom: Leverhulme Trust (Leverhulme Trust RPG-2020-004), Royal
Society (NIF-R1-231091); United States of America: Neubauer Family Foundation.

References
[1] D0 Collaboration, Observation of the top quark, Phys. Rev. Lett. 74 (1995) 2632,
arXiv: hep-ex/9503003.
[2] CDF Collaboration,
Observation of Top Quark Production in 𝑝¯ 𝑝 Collisions with the Collider Detector at Fermilab,
Phys. Rev. Lett. 74 (1995) 2626, arXiv: hep-ex/9503002.
[3] ATLAS Collaboration, The ATLAS Experiment at the CERN Large Hadron Collider,
JINST 3 (2008) S08003.
[4] CMS Collaboration, The CMS Experiment at the CERN LHC, JINST 3 (2008) S08004.
[5] S. Alekhin, A. Djouadi and S. Moch,
The top quark and Higgs boson masses and the stability of the electroweak vacuum,
Phys. Lett. B 716 (2012) 214, arXiv: 1207.0980 [hep-ph].
[6] J. R. Espinosa, Implications of the top (and Higgs) mass for vacuum stability,
PoS TOP2015 (2016) 043, ed. by L. Lista, F. Margaroli and F. Tramontano,
arXiv: 1512.01222 [hep-ph].
[7] ATLAS Collaboration, Exploration at the high-energy frontier: ATLAS Run 2 searches
investigating the exotic jungle beyond the Standard Model, (2024),
arXiv: 2403.09292 [hep-ex].
[8] ATLAS Collaboration, Performance of the ATLAS trigger system in 2015,
Eur. Phys. J. C 77 (2017) 317, arXiv: 1611.09661 [hep-ex].
[9] ATLAS Collaboration,
The performance of the jet trigger for the ATLAS detector during 2011 data taking,
Eur. Phys. J. C 76 (2016) 526, arXiv: 1606.07759 [hep-ex].
[10] ATLAS Collaboration, Configuration and performance of the ATLAS 𝑏-jet triggers in Run 2,
Eur. Phys. J. C 81 (2021) 1087, arXiv: 2106.03584 [hep-ex].
[11] ATLAS Collaboration, Performance of the ATLAS muon triggers in Run 2,
JINST 15 (2020) P09015, arXiv: 2004.13447 [physics.ins-det].
[12] ATLAS Collaboration, Performance of electron and photon triggers in ATLAS during LHC Run 2,
Eur. Phys. J. C 80 (2020) 47, arXiv: 1909.00761 [hep-ex].
[13] ATLAS Collaboration, ATLAS Insertable B-Layer: Technical Design Report,
ATLAS-TDR-19; CERN-LHCC-2010-013, 2010,
url: https://cds.cern.ch/record/1291633, Addendum: ATLAS-TDR-19-ADD-1;
CERN-LHCC-2012-009, 2012, url: https://cds.cern.ch/record/1451888.
[14] B. Abbott et al., Production and integration of the ATLAS Insertable B-Layer,
JINST 13 (2018) T05008, arXiv: 1803.00844 [physics.ins-det].
[15] ATLAS Collaboration, ATLAS flavour-tagging algorithms for the LHC Run 2 𝑝 𝑝 collision dataset,
Eur. Phys. J. C 83 (2023) 681, arXiv: 2211.16345 [physics.data-an].

53
[16] ATLAS Collaboration, ATLAS 𝑏-jet identification performance and efficiency measurement with 𝑡 𝑡¯

events in 𝑝 𝑝 collisions at 𝑠 = 13 TeV, Eur. Phys. J. C 79 (2019) 970,
arXiv: 1907.05120 [hep-ex].
[17] ATLAS Collaboration,
Calibration of the light-flavour jet mistagging efficiency of the 𝑏-tagging algorithms with 𝑍+jets

events using 139 fb−1 of ATLAS proton–proton collision data at 𝑠 = 13 TeV,
Eur. Phys. J. C 83 (2023) 728, arXiv: 2301.06319 [hep-ex].
[18] ATLAS Collaboration, Measurement of the 𝑐-jet mistagging efficiency in 𝑡 𝑡¯ events using 𝑝 𝑝

collision data at 𝑠 = 13 TeV collected with the ATLAS detector, Eur. Phys. J. C 82 (2022) 95,
arXiv: 2109.10627 [hep-ex].
[19] ATLAS Collaboration, Dependence of the Jet Energy Scale on the Particle Content of Hadronic
Jets in the ATLAS Detector Simulation, ATL-PHYS-PUB-2022-021, 2022,
url: https://cds.cern.ch/record/2808016.
[20] ATLAS Collaboration, Jet energy scale and resolution measured in proton–proton collisions at

𝑠 = 13 TeV with the ATLAS detector, Eur. Phys. J. C 81 (2021) 689,
arXiv: 2007.02645 [hep-ex].
[21] ATLAS Collaboration, In situ calibration of large-radius jet energy and mass in 13 TeV
proton–proton collisions with the ATLAS detector, Eur. Phys. J. C 79 (2019) 135,
arXiv: 1807.09477 [hep-ex].
[22] ATLAS Collaboration, Performance of pile-up mitigation techniques for jets in 𝑝 𝑝 collisions at

𝑠 = 8 TeV using the ATLAS detector, Eur. Phys. J. C 76 (2016) 581,
arXiv: 1510.03823 [hep-ex].
[23] ATLAS Collaboration, Electron reconstruction and identification in the ATLAS experiment using

the 2015 and 2016 LHC proton–proton collision data at 𝑠 = 13 TeV,
Eur. Phys. J. C 79 (2019) 639, arXiv: 1902.04655 [physics.ins-det].
[24] ATLAS Collaboration, Electron and photon efficiencies in LHC Run 2 with the ATLAS experiment,
(2023), arXiv: 2308.13362 [hep-ex].
[25] ATLAS Collaboration, Studies of the muon momentum calibration and performance of the ATLAS

detector with 𝑝 𝑝 collisions at 𝑠 = 13 TeV, Eur. Phys. J. C 83 (2023) 686,
arXiv: 2212.07338 [hep-ex].
[26] ATLAS Collaboration,

Luminosity determination in 𝑝 𝑝 collisions at 𝑠 = 13 TeV using the ATLAS detector at the LHC,
ATLAS-CONF-2019-021, 2019, url: https://cds.cern.ch/record/2677054.
[27] ATLAS Collaboration,

Luminosity determination in 𝑝 𝑝 collisions at 𝑠 = 13 TeV using the ATLAS detector at the LHC,
Eur. Phys. J. C 83 (2023) 982, arXiv: 2212.09379 [hep-ex].
[28] ATLAS Collaboration, Muon reconstruction and identification efficiency in ATLAS using the full

Run 2 𝑝 𝑝 collision data set at 𝑠 = 13 TeV, Eur. Phys. J. C 81 (2021) 578,
arXiv: 2012.00578 [hep-ex].
[29] ATLAS Collaboration, Evidence for the associated production of the Higgs boson and a top quark
pair with the ATLAS detector, Phys. Rev. D 97 (2018) 072003, arXiv: 1712.08891 [hep-ex].

54
[30] ATLAS Collaboration,
Topological cell clustering in the ATLAS calorimeters and its performance in LHC Run 1,
Eur. Phys. J. C 77 (2017) 490, arXiv: 1603.02934 [hep-ex].
[31] ATLAS Collaboration,
Jet reconstruction and performance using particle flow with the ATLAS Detector,
Eur. Phys. J. C 77 (2017) 466, arXiv: 1703.10485 [hep-ex].
[32] M. Cacciari, G. P. Salam and G. Soyez, The anti-𝑘 𝑡 jet clustering algorithm, JHEP 04 (2008) 063,
arXiv: 0802.1189 [hep-ph].
[33] M. Cacciari, G. P. Salam and G. Soyez, FastJet User Manual, Eur. Phys. J. C 72 (2012) 1896,
arXiv: 1111.6097 [hep-ph].
[34] B. Nachman, P. Nef, A. Schwartzman, M. Swiatlowski and C. Wanotayaroj,
Jets from Jets: Re-clustering as a tool for large radius jet reconstruction and grooming at the LHC,
JHEP 02 (2015) 075, arXiv: 1407.2922 [hep-ph].
[35] D. Krohn, J. Thaler and L.-T. Wang, Jet Trimming, JHEP 02 (2010) 084,
arXiv: 0912.1342 [hep-ph].
[36] M. Dasgupta, A. Fregoso, S. Marzani and G. P. Salam,
Towards an understanding of jet substructure, JHEP 09 (2013) 029, arXiv: 1307.0007 [hep-ph].
[37] A. J. Larkoski, S. Marzani, G. Soyez and J. Thaler, Soft Drop, JHEP 05 (2014) 146,
arXiv: 1402.2657 [hep-ph].
[38] ATLAS Collaboration, Optimisation of large-radius jet reconstruction for the ATLAS detector in
13 TeV proton–proton collisions, Eur. Phys. J. C 81 (2021) 334, arXiv: 2009.04986 [hep-ex].
[39] ATLAS Collaboration, The performance of missing transverse momentum reconstruction and its

significance with the ATLAS detector using 140 fb−1 of 𝑠 = 13 TeV 𝑝 𝑝 collisions, (2024),
arXiv: 2402.05858 [hep-ex].
[40] G. D’Agostini, A Multidimensional unfolding method based on Bayes’ theorem,
Nucl. Instrum. Meth. A 362 (1995) 487.
[41] G. D’Agostini, ‘Improved iterative Bayesian unfolding’,
Alliance Workshop on Unfolding and Data Correction, 2010,
arXiv: 1010.0632 [physics.data-an].
[42] G. Choudalakis, Fully Bayesian Unfolding, (2012), arXiv: 1201.4612 [physics.data-an].
[43] V. Blobel, ‘An Unfolding method for high-energy physics experiments’,
Conference on Advanced Statistical Techniques in Particle Physics, 2002 258,
arXiv: hep-ex/0208022.
[44] G. Cowan, Statistical data analysis, Oxford University Press, 1998.
[45] A. L. Read, Presentation of search results: The 𝐶 𝐿 𝑠 technique,
J. Phys. G 28 (2002) 2693, ed. by M. R. Whalley and L. Lyons.
[46] G. Cowan, K. Cranmer, E. Gross and O. Vitells,
Asymptotic formulae for likelihood-based tests of new physics, Eur. Phys. J. C 71 (2011) 1554,
arXiv: 1007.1727 [physics.data-an], Erratum: Eur. Phys. J. C 73 (2013) 2501.
[47] ATLAS Collaboration,

Measurements of 𝑏-jet tagging efficiency with the ATLAS detector using 𝑡 𝑡¯ events at 𝑠 = 13 TeV,
JHEP 08 (2018) 089, arXiv: 1805.01845 [hep-ex].

55
[48] ATLAS Collaboration,
Improvements in 𝑡 𝑡¯ modelling using NLO+PS Monte Carlo generators for Run 2,
ATL-PHYS-PUB-2018-009, 2018, url: https://cds.cern.ch/record/2630327.
[49] S. Alioli, P. Nason, C. Oleari and E. Re, A general framework for implementing NLO calculations
in shower Monte Carlo programs: the POWHEG BOX, JHEP 06 (2010) 043,
arXiv: 1002.2581 [hep-ph].
[50] T. Sjöstrand, S. Mrenna and P. Z. Skands, PYTHIA 6.4 physics and manual, JHEP 05 (2006) 026,
arXiv: hep-ph/0603175.
[51] D. J. Lange, The EvtGen particle decay simulation package,
Nucl. Instrum. Meth. A 462 (2001) 152, ed. by S. Erhan, P. Schlein and Y. Rozen.
[52] T. Sjöstrand et al., An introduction to PYTHIA 8.2, Comput. Phys. Commun. 191 (2015) 159,
arXiv: 1410.3012 [hep-ph].
[53] P. Z. Skands, Tuning Monte Carlo Generators: The Perugia Tunes, Phys. Rev. D 82 (2010) 074018,
arXiv: 1005.3457 [hep-ph].
[54] ATLAS Collaboration, ATLAS Pythia 8 tunes to 7 TeV data, ATL-PHYS-PUB-2014-021, 2014,
url: https://cds.cern.ch/record/1966419.
[55] J. Bellm et al., Herwig 7.0/Herwig++ 3.0 release note, Eur. Phys. J. C 76 (2016) 196,
arXiv: 1512.01178 [hep-ph].
[56] J. Alwall et al., The automated computation of tree-level and next-to-leading order differential
cross sections, and their matching to parton shower simulations, JHEP 07 (2014) 079,
arXiv: 1405.0301 [hep-ph].
[57] E. Bothmann et al., Event Generation with Sherpa 2.2, SciPost Phys. 7 (2019) 034,
arXiv: 1905.09127 [hep-ph].
[58] ATLAS Collaboration, Studies on the improvement of the matching uncertainty definition in
top-quark processes simulated with Powheg+Pythia8, ATL-PHYS-PUB-2023-029, 2013,
url: https://cds.cern.ch/record/2872787.
[59] P. Artoisenet, R. Frederix, O. Mattelaer and R. Rietkerk,
Automatic spin-entangled decays of heavy resonances in Monte Carlo simulations,
JHEP 03 (2013) 015, arXiv: 1212.3460 [hep-ph].
[60] M. Czakon et al., Top-pair production at the LHC through NNLO QCD and NLO EW,
JHEP 10 (2017) 186, arXiv: 1705.04105 [hep-ph].
[61] NNPDF Collaboration, R. D. Ball et al., Parton distributions for the LHC run II,
JHEP 04 (2015) 040, arXiv: 1410.8849 [hep-ph].
[62] J. Butterworth et al., PDF4LHC recommendations for LHC Run II, J. Phys. G 43 (2016) 023001,
arXiv: 1510.03865 [hep-ph].
[63] ATLAS Collaboration, A study of different colour reconnection settings for Pythia8 generator
using underlying event observables, ATL-PHYS-PUB-2017-008, 2017,
url: https://cds.cern.ch/record/2262253.
[64] T. Sjöstrand, Colour reconnection and its effects on precise measurements at the LHC, (2013),
arXiv: 1310.8073 [hep-ph].
[65] S. Argyropoulos and T. Sjöstrand, Effects of color reconnection on 𝑡 𝑡¯ final states at the LHC,
JHEP 11 (2014) 043, arXiv: 1407.6653 [hep-ph].

56
[66] S. Frixione, E. Laenen, P. Motylinski, C. D. White and B. R. Webber,
Single-top hadroproduction in association with a W boson, JHEP 07 (2008) 029,
arXiv: 0805.3067 [hep-ph].

[67] ATLAS Collaboration, Modelling of rare top quark processes at 𝑠 = 13 TeV in ATLAS,
ATL-PHYS-PUB-2020-024, 2020, url: https://cds.cern.ch/record/2730584.
[68] ATLAS Collaboration, Study of ttbb and ttW background modelling for ttH analyses,
ATL-PHYS-PUB-2022-026, 2022, url: https://cds.cern.ch/record/2810864.

¯ + ℓ − 𝜈 𝜈¯ ′ final states with
[69] ATLAS Collaboration, Studies of 𝑡 𝑡¯/𝑡𝑊 interference effects in 𝑏 𝑏ℓ
Powheg and MG5_aMC@NLO setups, ATL-PHYS-PUB-2021-042, 2021,
url: https://cds.cern.ch/record/2792254.
[70] J. Mazzitelli et al., Top-pair production at the LHC with MINNLO 𝑃𝑆 , JHEP 04 (2022) 079,
arXiv: 2112.12135 [hep-ph].
[71] ATLAS Collaboration, Measurement of differential cross sections in 𝑡 𝑡¯ and 𝑡 𝑡¯+jets production in

the lepton+jets decay mode in 𝑝 𝑝 collisions at 𝑠 = 13 TeV using 140 fb−1 of ATLAS data,
ATLAS-CONF-2023-068, 2023, url: https://cds.cern.ch/record/2873524.
[72] M. Czakon, P. Fiedler and A. Mitov,
Total Top-Quark Pair-Production Cross Section at Hadron Colliders Through 𝑂 (𝛼𝑆4 ),
Phys. Rev. Lett. 110 (2013) 252004, arXiv: 1303.6254 [hep-ph].
[73] W. Bernreuther, M. Fücker and Z.-G. Si,
Weak interaction corrections to hadronic top quark pair production,
Phys. Rev. D 74 (2006) 113005, arXiv: hep-ph/0610334.
[74] J. H. Kühn, A. Scharf and P. Uwer,
Weak interaction effects in top-quark pair production at hadron colliders,
Eur. Phys. J. C 51 (2007) 37, arXiv: hep-ph/0610335.
[75] W. Hollik and M. Kollár, NLO QED contributions to top-pair production at hadron colliders,
Phys. Rev. D 77 (2008) 014008, arXiv: 0708.1697 [hep-ph].
[76] M. Cacciari, M. Czakon, M. Mangano, A. Mitov and P. Nason, Top-pair production at hadron
colliders with next-to-next-to-leading logarithmic soft-gluon resummation,
Phys. Lett. B 710 (2012) 612, arXiv: 1111.5869 [hep-ph].
[77] N. Kidonakis, M. Guzzi and A. Tonero,
Top-quark cross sections and distributions at approximate N3 LO, Phys. Rev. D 108 (2023) 054012,
arXiv: 2306.06166 [hep-ph].
[78] R. D. Ball et al., The PDF4LHC21 combination of global PDF fits for the LHC Run III,
J. Phys. G 49 (2022) 080501, arXiv: 2203.05506 [hep-ph].
[79] M. Czakon and A. Mitov,
Top++: A Program for the Calculation of the Top-Pair Cross-Section at Hadron Colliders,
Comput. Phys. Commun. 185 (2014) 2930, arXiv: 1112.5675 [hep-ph].
[80] ATLAS Collaboration, Measurement of the 𝑡 𝑡¯ production cross-section using 𝑒𝜇 events with

𝑏-tagged jets in 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector,
Phys. Lett. B 761 (2016) 136, arXiv: 1606.02699 [hep-ex],
Erratum: Phys. Lett. B 772 (2017) 879.

57
[81] ATLAS Collaboration, Measurement of the 𝑡 𝑡¯ production cross-section using 𝑒𝜇 events with

𝑏-tagged jets in 𝑝 𝑝 collisions at 𝑠 = 7 and 8 TeV with the ATLAS detector,
Eur. Phys. J. C 74 (2014) 3109, arXiv: 1406.5375 [hep-ex],
Addendum: Eur. Phys. J. C 76 (2016) 642.
[82] ATLAS Collaboration, Measurement of the 𝑡 𝑡¯ production cross-section and lepton differential

distributions in 𝑒𝜇 dilepton events from 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector,
Eur. Phys. J. C 80 (2020) 528, arXiv: 1910.08819 [hep-ex].
[83] ATLAS Collaboration, Inclusive and differential cross-sections for dilepton 𝑡 𝑡¯ production

measured in 𝑠 = 13 TeV 𝑝 𝑝 collisions with the ATLAS detector, JHEP 07 (2023) 141,
arXiv: 2303.15340 [hep-ex].
[84] E. Todesco and J. Wenninger, Large Hadron Collider momentum calibration and accuracy,
Phys. Rev. Accel. Beams 20 (2017) 081003.
[85] ATLAS Collaboration, Measurements of top-quark pair to 𝑍-boson cross-section ratios at

𝑠 = 13, 8, 7 TeV with the ATLAS detector, JHEP 02 (2017) 117, arXiv: 1612.03636 [hep-ex].
[86] ATLAS Collaboration, Measurement of the 𝑡 𝑡¯ production cross-section in the lepton+jets channel

at 𝑠 = 13 TeV with the ATLAS experiment, Phys. Lett. B 810 (2020) 135797,
arXiv: 2006.13076 [hep-ex].
[87] ATLAS Collaboration,
Measurements of top-quark pair single- and double-differential cross-sections in the all-hadronic

channel in 𝑝 𝑝 collisions at 𝑠 = 13 TeV using the ATLAS detector, JHEP 01 (2021) 033,
arXiv: 2006.09274 [hep-ex].
[88] ATLAS Collaboration,
Measurement of the inclusive 𝑡 𝑡¯ production cross section in the lepton + jets channel in 𝑝 𝑝

collisions at 𝑠 = 7 TeV with the ATLAS detector using support vector machines,
Phys. Rev. D 108 (2023) 032014, arXiv: 2212.00571 [hep-ex].
[89] ATLAS Collaboration, Measurement of the 𝑡 𝑡¯ production cross-section in 𝑝 𝑝 collisions at

𝑠 = 5.02 TeV with the ATLAS detector, JHEP 06 (2023) 138, arXiv: 2207.01354 [hep-ex].
[90] ATLAS Collaboration, Measurement of the 𝑡 𝑡¯ cross section and its ratio to the 𝑍 production cross

section using 𝑝 𝑝 collisions at 𝑠 = 13.6 TeV with the ATLAS detector,
Phys. Lett. B 848 (2024) 138376, arXiv: 2308.09529 [hep-ex].
[91] ATLAS Collaboration,
Measurement of jet activity produced in top-quark events with an electron, a muon and two

𝑏-tagged jets in the final state in 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector,
Eur. Phys. J. C 77 (2017) 220, arXiv: 1610.09978 [hep-ex].
[92] ATLAS Collaboration, Measurements of top-quark pair differential cross-sections in the 𝑒𝜇

channel in 𝑝 𝑝 collisions at 𝑠 = 13 TeV using the ATLAS detector, Eur. Phys. J. C 77 (2017) 292,
arXiv: 1612.05220 [hep-ex].
[93] ATLAS Collaboration, Measurements of top-quark pair differential cross-sections in the

lepton+jets channel in 𝑝 𝑝 collisions at 𝑠 = 13 TeV using the ATLAS detector,
JHEP 11 (2017) 191, arXiv: 1708.00727 [hep-ex].
[94] ATLAS Collaboration, Measurements of differential cross sections of top quark pair production in

association with jets in 𝑝 𝑝 collisions at 𝑠 = 13 TeV using the ATLAS detector,
JHEP 10 (2018) 159, arXiv: 1802.06572 [hep-ex].

58
[95] ATLAS Collaboration, Measurements of top-quark pair differential and double-differential

cross-sections in the ℓ+jets channel with 𝑝 𝑝 collisions at 𝑠 = 13 TeV using the ATLAS detector,
Eur. Phys. J. C 79 (2019) 1028, arXiv: 1908.07305 [hep-ex],
Erratum: Eur. Phys. J. C 80 (2020) 1092.
[96] ATLAS Collaboration, Measurements of 𝑡 𝑡¯ differential cross-sections of highly boosted top quarks

decaying to all-hadronic final states in 𝑝 𝑝 collisions at 𝑠 = 13 TeV using the ATLAS detector,
Phys. Rev. D 98 (2018) 012003, arXiv: 1801.02052 [hep-ex].
[97] ATLAS Collaboration, Measurements of differential cross-sections in top-quark pair events with a
high transverse momentum top quark and limits on beyond the Standard Model contributions to

top-quark pair production with the ATLAS detector at 𝑠 = 13 TeV, JHEP 06 (2022) 063,
arXiv: 2202.12134 [hep-ex].
[98] ATLAS Collaboration, Differential 𝑡 𝑡¯ cross-section measurements using boosted top quarks in the
all-hadronic final state with 139 fb−1 of ATLAS data, JHEP 04 (2023) 080,
arXiv: 2205.02817 [hep-ex].
[99] D0 Collaboration, Measurement of the top quark mass using dilepton events,
Phys. Rev. Lett. 80 (1998) 2063, arXiv: hep-ex/9706014.
[100] ATLAS Collaboration, Measurement of lepton differential distributions and the top quark mass in

𝑡 𝑡¯ production in 𝑝 𝑝 collisions at 𝑠 = 8 TeV with the ATLAS detector,
Eur. Phys. J. C 77 (2017) 804, arXiv: 1709.09407 [hep-ex].
[101] ATLAS Collaboration,
Differential top-antitop cross-section measurements as a function of observables constructed from

final-state particles using 𝑝 𝑝 collisions at 𝑠 = 7 TeV in the ATLAS detector, JHEP 06 (2015) 100,
arXiv: 1502.05923 [hep-ex].
[102] J. Erdmann et al.,
A likelihood-based reconstruction algorithm for top-quark pairs and the KLFitter framework,
Nucl. Instrum. Meth. A 748 (2014) 18, arXiv: 1312.5595 [hep-ex].
[103] ATLAS Collaboration,
Measurements of inclusive and differential fiducial cross-sections of 𝑡 𝑡¯ production with additional

heavy-flavour jets in proton–proton collisions at 𝑠 = 13 TeV with the ATLAS detector,
JHEP 04 (2019) 046, arXiv: 1811.12113 [hep-ex].
[104] ATLAS Collaboration, Measurements of jet observables sensitive to 𝑏-quark fragmentation in 𝑡 𝑡¯
events at the LHC with the ATLAS detector, Phys. Rev. D 106 (2022) 032008,
arXiv: 2202.13901 [hep-ex].
[105] T. M. P. Tait and C.-P. Yuan,
Single top quark production as a window to physics beyond the standard model,
Phys. Rev. D 63 (2000) 014018, arXiv: hep-ph/0007298.
[106] Q.-H. Cao, J. Wudka and C.-P. Yuan, Search for new physics via single-top production at the LHC,
Phys. Lett. B 658 (2007) 50, arXiv: 0704.2809 [hep-ph].
[107] J. Campbell, T. Neumann and Z. Sullivan, Single-top-quark production in the 𝑡-channel at NNLO,
JHEP 02 (2021) 040, arXiv: 2012.01574 [hep-ph].
[108] ATLAS Collaboration, Measurement of the inclusive cross-sections of single top-quark and

top-antiquark 𝑡-channel production in 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector,
JHEP 04 (2017) 086, arXiv: 1609.03920 [hep-ex].

59
[109] ATLAS Collaboration, Measurement of 𝑡-channel production of single top quarks and antiquarks
in 𝑝 𝑝 collisions at 13 TeV using the full ATLAS Run 2 data sample, (2024),
arXiv: 2403.02126 [hep-ex].
[110] ATLAS Collaboration,
Measurement of the polarisation of single top quarks and antiquarks produced in the t-channel at

𝑠 = 13 TeV and bounds on the tWb dipole operator from the ATLAS experiment,
JHEP 11 (2022) 040, arXiv: 2202.11382 [hep-ex].
[111] J. A. Aguilar-Saavedra and S. Amor Dos Santos,
New directions for top quark polarization in the 𝑡-channel process, Phys. Rev. D 89 (2014) 114009,
arXiv: 1404.1585 [hep-ph].
[112] ATLAS Collaboration, Probing the quantum interference between singly and doubly resonant

top-quark production in 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector,
Phys. Rev. Lett. 121 (2018) 152002, arXiv: 1806.04667 [hep-ex].
[113] CMS Collaboration, Observation of the associated production of a single top quark and a 𝑊 boson

in 𝑝 𝑝 collisions at 𝑠 =8 TeV, Phys. Rev. Lett. 112 (2014) 231802, arXiv: 1401.2942 [hep-ex].
[114] ATLAS Collaboration, Measurement of the production cross-section of a single top quark in
association with a 𝑊 boson at 8 TeV with the ATLAS experiment, JHEP 01 (2016) 064,
arXiv: 1510.03752 [hep-ex].
[115] N. Kidonakis and N. Yamanaka,
Higher-order corrections for 𝑡𝑊 production at high-energy hadron colliders, JHEP 05 (2021) 278,
arXiv: 2102.11300 [hep-ph].
[116] ATLAS Collaboration, Measurement of the cross-section for producing a W boson in association

with a single top quark in pp collisions at 𝑠 = 13 TeV with ATLAS, JHEP 01 (2018) 063,
arXiv: 1612.07231 [hep-ex].
[117] ATLAS Collaboration, Measurement of differential cross-sections of a single top quark produced

in association with a 𝑊 boson at 𝑠 = 13 TeV with ATLAS, Eur. Phys. J. C 78 (2018) 186,
arXiv: 1712.01602 [hep-ex].
[118] C. G. Lester and D. J. Summers,
Measuring masses of semi-invisibly decaying particles pair produced at hadron colliders,
Phys. Lett. B 463 (1999) 99, arXiv: hep-ph/9906349.
[119] A. Barr, C. Lester and P. Stephens, A variable for measuring masses at hadron colliders when
missing energy is expected; 𝑚 𝑇2 : the truth behind the glamour, J. Phys. G 29 (2003) 2343,
arXiv: hep-ph/0304226.
[120] T. Ježo, J. M. Lindert, P. Nason, C. Oleari and S. Pozzorini, An NLO+PS generator for 𝑡 𝑡¯ and 𝑊𝑡
production and decay including non-resonant and interference effects,
Eur. Phys. J. C 76 (2016) 691, arXiv: 1607.04538 [hep-ph].
[121] T. Ježo and P. Nason, On the Treatment of Resonances in Next-to-Leading Order Calculations
Matched to a Parton Shower, JHEP 12 (2015) 1, arXiv: 1509.09071 [hep-ph].
[122] CDF and D0 Collaborations,
Observation of s-channel production of single top quarks at the Tevatron,
Phys. Rev. Lett. 112 (2014) 231803, arXiv: 1402.5126 [hep-ex].

60
[123] ATLAS Collaboration, Evidence for single top-quark production in the 𝑠-channel in proton-proton

collisions at 𝑠 =8 TeV with the ATLAS detector using the Matrix Element Method,
Phys. Lett. B 756 (2016) 228, arXiv: 1511.05980 [hep-ex].
[124] P. Kant et al., HatHor for single top-quark production: Updated predictions and uncertainty
estimates for single top-quark production in hadronic collisions,
Comput. Phys. Commun. 191 (2015) 74, arXiv: 1406.4403 [hep-ph].
[125] ATLAS Collaboration, Measurement of single top-quark production in the s-channel in

proton–proton collisions at 𝑠 = 13 TeV with the ATLAS detector, JHEP 06 (2023) 191,
arXiv: 2209.08990 [hep-ex].
[126] K. Kondo, Dynamical Likelihood Method for Reconstruction of Events With Missing Momentum.
1: Method and Toy Models, J. Phys. Soc. Jap. 57 (1988) 4126.
[127] K. Kondo, Dynamical likelihood method for reconstruction of events with missing momentum. 2:
Mass spectra for 2 —> 2 processes, J. Phys. Soc. Jap. 60 (1991) 836.
[128] N. Cabibbo, Unitary Symmetry and Leptonic Decays, Phys. Rev. Lett. 10 (1963) 531.
[129] M. Kobayashi and T. Maskawa, 𝐶𝑃-Violation in the Renormalizable Theory of Weak Interaction,
Prog. Theor. Phys. 49 (1973) 652.
[130] J. A. Aguilar-Saavedra, A Minimal set of top anomalous couplings, Nucl. Phys. B 812 (2009) 181,
arXiv: 0811.3842 [hep-ph].
[131] N. Kidonakis, Theoretical results for electroweak boson and single-top production,
PoS DIS2015 (2015) 170, arXiv: 1506.04072 [hep-ph].
[132] ATLAS Collaboration, Characterising the Higgs boson with ATLAS data from Run 2 of the LHC,
(2024), arXiv: 2404.05498 [hep-ex].
[133] J. A. Aguilar-Saavedra, Identifying top partners at LHC, JHEP 11 (2009) 030,
arXiv: 0907.3155 [hep-ph].
[134] J. A. Aguilar-Saavedra, R. Benbrik, S. Heinemeyer and M. Pérez-Victoria,
Handbook of vectorlike quarks: Mixing and single production, Phys. Rev. D 88 (2013) 094010,
arXiv: 1306.0572 [hep-ph].
[135] M. Perelstein, Little Higgs models and their phenomenology,
Prog. Part. Nucl. Phys. 58 (2007) 247, arXiv: hep-ph/0512128.
[136] D. Dicus, A. Stange and S. Willenbrock, Higgs decay to top quarks at hadron colliders,
Phys. Lett. B 333 (1994) 126, arXiv: hep-ph/9404359.
[137] N. Craig, F. D’Eramo, P. Draper, S. Thomas and H. Zhang,
The Hunt for the Rest of the Higgs Bosons, JHEP 06 (2015) 137, arXiv: 1504.04630 [hep-ph].
[138] N. Craig, J. Hajer, Y.-Y. Li, T. Liu and H. Zhang,
Heavy Higgs bosons at low tan 𝛽: from the LHC to 100 TeV, JHEP 01 (2017) 018,
arXiv: 1605.08744 [hep-ph].
[139] C. Degrande, J.-M. Gérard, C. Grojean, F. Maltoni and G. Servant,
Non-resonant New Physics in Top Pair Production at Hadron Colliders, JHEP 03 (2011) 125,
arXiv: 1010.6304 [hep-ph].
[140] C. Zhang,
Constraining 𝑞𝑞𝑡𝑡 operators from four-top production: a case for enhanced EFT sensitivity,
Chin. Phys. C 42 (2018) 023104, arXiv: 1708.05928 [hep-ph].

61
[141] ATLAS Collaboration,
Tools for estimating fake/non-prompt lepton backgrounds with the ATLAS detector at the LHC,
JINST 18 (2023) T11004, arXiv: 2211.16178 [hep-ex].
[142] ATLAS Collaboration, Electron and photon performance measurements with the ATLAS detector
using the 2015–2017 LHC proton–proton collision data, JINST 14 (2019) P12006,
arXiv: 1908.00005 [hep-ex].
[143] ATLAS Collaboration, Measurement of the 𝑡𝑡𝑊 and 𝑡𝑡𝑍 production cross sections in pp collisions

at 𝑠 = 8 TeV with the ATLAS detector, JHEP 11 (2015) 172, arXiv: 1509.05276 [hep-ex].
[144] ATLAS Collaboration, Measurement of the 𝑡 𝑡¯𝑍 and 𝑡 𝑡¯𝑊 production cross sections in multilepton

final states using 3.2 fb−1 of 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector,
Eur. Phys. J. C 77 (2017) 40, arXiv: 1609.01599 [hep-ex].
[145] ATLAS Collaboration, Measurement of the 𝑡 𝑡¯𝑍 and 𝑡 𝑡¯𝑊 cross sections in proton-proton collisions

at 𝑠 = 13 TeV with the ATLAS detector, Phys. Rev. D 99 (2019) 072009,
arXiv: 1901.03584 [hep-ex].
[146] ATLAS Collaboration, Measurement of the total and differential cross-sections of 𝑡 𝑡¯𝑊 production

in 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector, (2024), arXiv: 2401.05299 [hep-ex].
[147] ATLAS Collaboration,
Inclusive and differential cross-section measurements of 𝑡 𝑡¯𝑍 production in 𝑝 𝑝 collisions at

𝑠 = 13 TeV with the ATLAS detector, including EFT and spin-correlation interpretations, (2023),
arXiv: 2312.04450 [hep-ex].
[148] A. Kulesza, L. Motyka, D. Schwartländer, T. Stebel and V. Theeuwes, Associated production of a
top quark pair with a heavy electroweak gauge boson at NLO+NNLL accuracy,
Eur. Phys. J. C 79 (2019) 249, arXiv: 1812.08622 [hep-ph].
[149] L. Buonocore et al., Precise Predictions for the Associated Production of a W Boson with a
Top-Antitop Quark Pair at the LHC, Phys. Rev. Lett. 131 (2023) 231901,
arXiv: 2306.16311 [hep-ph].
[150] ATLAS Collaboration, Measurements of the inclusive and differential production cross sections of

a top-quark–antiquark pair in association with a Z boson at 𝑠 = 13 TeV with the ATLAS detector,
Eur. Phys. J. C 81 (2021) 737, arXiv: 2103.12603 [hep-ex].
[151] ATLAS Collaboration, Measurement of the production cross-section of a single top quark in
association with a Z boson in proton–proton collisions at 13 TeV with the ATLAS detector,
Phys. Lett. B 780 (2018) 557, arXiv: 1710.03659 [hep-ex].
[152] ATLAS Collaboration, Observation of the associated production of a top quark and a 𝑍 boson in

𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector, JHEP 07 (2020) 124,
arXiv: 2002.07546 [hep-ex].
[153] J. Campbell, R. K. Ellis and R. Röntsch,
Single top production in association with a 𝑍 boson at the LHC, Phys. Rev. D 87 (2013) 114006,
arXiv: 1302.3856 [hep-ph].
[154] U. Baur, A. Juste, L. H. Orr and D. Rainwater,
Probing electroweak top quark couplings at hadron colliders, Phys. Rev. D 71 (2005) 054013,
arXiv: hep-ph/0412021.
[155] A. O. Bouzas and F. Larios, Electromagnetic dipole moments of the Top quark,
Phys. Rev. D 87 (2013) 074015, arXiv: 1212.6575 [hep-ph].

62
[156] M. Schulze and Y. Soreq, Pinning down electroweak dipole operators of the top quark,
Eur. Phys. J. C 76 (2016) 466, arXiv: 1603.08911 [hep-ph].
[157] ATLAS Collaboration,
Observation of top-quark pair production in association with a photon and measurement of the 𝑡 𝑡¯𝛾

production cross section in pp collisions at 𝑠 = 7 TeV using the ATLAS detector,
Phys. Rev. D 91 (2015) 072007, arXiv: 1502.00586 [hep-ex].
[158] ATLAS Collaboration, Measurement of the 𝑡𝑡𝛾 production cross section in proton-proton

collisions at 𝑠 = 8 TeV with the ATLAS detector, JHEP 11 (2017) 086,
arXiv: 1706.03046 [hep-ex].
[159] ATLAS Collaboration, Measurements of inclusive and differential fiducial cross-sections of 𝑡 𝑡¯𝛾

production in leptonic final states at 𝑠 = 13 TeV in ATLAS, Eur. Phys. J. C 79 (2019) 382,
arXiv: 1812.01697 [hep-ex].
[160] ATLAS Collaboration, Measurements of inclusive and differential cross-sections of combined 𝑡𝑡𝛾
and 𝑡𝑊 𝛾 production in the e𝜇 channel at 13 TeV with the ATLAS detector, JHEP 09 (2020) 049,
arXiv: 2007.06946 [hep-ex].
[161] K. Melnikov, M. Schulze and A. Scharf,
QCD corrections to top quark pair production in association with a photon at hadron colliders,
Phys. Rev. D 83 (2011) 074013, arXiv: 1102.1967 [hep-ph].
[162] G. Bevilacqua, H. B. Hartanto, M. Kraus, T. Weber and M. Worek,
Hard Photons in Hadroproduction of Top Quarks with Realistic Final States, JHEP 10 (2018) 158,
arXiv: 1803.09916 [hep-ph].
[163] G. Bevilacqua, H. B. Hartanto, M. Kraus, T. Weber and M. Worek,
Precise predictions for 𝑡 𝑡¯𝛾/𝑡 𝑡¯ cross section ratios at the LHC, JHEP 01 (2019) 188,
arXiv: 1809.08562 [hep-ph].
[164] ATLAS Collaboration, Observation of Single-Top-Quark Production in Association with a Photon
Using the ATLAS Detector, Phys. Rev. Lett. 131 (2023) 181901, arXiv: 2302.01283 [hep-ex].
[165] A. Pomarol and J. Serra, Top Quark Compositeness: Feasibility and Implications,
Phys. Rev. D 78 (2008) 074026, arXiv: 0806.3247 [hep-ph].
[166] H. P. Nilles, Supersymmetry, Supergravity and Particle Physics, Phys. Rept. 110 (1984) 1.
[167] G. R. Farrar and P. Fayet, Phenomenology of the Production, Decay, and Detection of New
Hadronic States Associated with Supersymmetry, Phys. Lett. B 76 (1978) 575.
[168] Q.-H. Cao, S.-L. Chen and Y. Liu,
Probing Higgs Width and Top Quark Yukawa Coupling from 𝑡 𝑡¯𝐻 and 𝑡 𝑡¯𝑡 𝑡¯ Productions,
Phys. Rev. D 95 (2017) 053004, arXiv: 1602.01934 [hep-ph].
[169] Q.-H. Cao, S.-L. Chen, Y. Liu, R. Zhang and Y. Zhang,
Limiting top quark-Higgs boson interaction and Higgs-boson width from multitop productions,
Phys. Rev. D 99 (2019) 113003, arXiv: 1901.04567 [hep-ph].
[170] R. Frederix, D. Pagani and M. Zaro, Large NLO corrections in 𝑡 𝑡¯𝑊 ± and 𝑡 𝑡¯𝑡 𝑡¯ hadroproduction
from supposedly subleading EW contributions, JHEP 02 (2018) 031,
arXiv: 1711.02116 [hep-ph].
[171] G. Bevilacqua and M. Worek, Constraining BSM Physics at the LHC: Four top final states with
NLO accuracy in perturbative QCD, JHEP 07 (2012) 111, arXiv: 1206.3064 [hep-ph].

63
[172] T. Ježo and M. Kraus, Hadroproduction of four top quarks in the powheg box,
Phys. Rev. D 105 (2022) 114024, arXiv: 2110.15159 [hep-ph].
[173] M. van Beekveld, A. Kulesza and L. M. Valero,
Threshold resummation for the production of four top quarks at the LHC, (2022),
arXiv: 2212.03259 [hep-ph].
[174] ATLAS Collaboration, Search for new phenomena in events with same-charge leptons and 𝑏-jets

in 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector, JHEP 12 (2018) 039,
arXiv: 1807.11883 [hep-ex].
[175] ATLAS Collaboration, Evidence for 𝑡 𝑡¯𝑡 𝑡¯ production in the multilepton final state in proton–proton

collisions at 𝑠 = 13TeV with the ATLAS detector, Eur. Phys. J. C 80 (2020) 1085,
arXiv: 2007.14858 [hep-ex].
[176] ATLAS Collaboration,
Observation of four-top-quark production in the multilepton final state with the ATLAS detector,
Eur. Phys. J. C 83 (2023) 496, arXiv: 2303.15061 [hep-ex].

[177] ATLAS Collaboration, Measurement of the properties of Higgs boson production at 𝑠 = 13 TeV
in the 𝐻 → 𝛾𝛾 channel using 139 fb−1 of 𝑝 𝑝 collision data with the ATLAS experiment,
JHEP 07 (2023) 088, arXiv: 2207.00348 [hep-ex].
[178] ATLAS Collaboration, Probing the 𝐶𝑃 nature of the top-Higgs Yukawa coupling in 𝑡 𝑡¯𝐻 and 𝑡𝐻
events with 𝐻 → 𝑏 𝑏¯ decays using the ATLAS detector at the LHC, (2023),
arXiv: 2303.05974 [hep-ex].
[179] ATLAS Collaboration, Search for four-top-quark production in the single-lepton and opposite-sign

dilepton final states in 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector,
Phys. Rev. D 99 (2019) 052009, arXiv: 1811.02305 [hep-ex].
[180] ATLAS Collaboration, Measurement of the 𝑡 𝑡¯𝑡 𝑡¯ production cross section in 𝑝 𝑝 collisions at

𝑠 = 13 TeV with the ATLAS detector, JHEP 11 (2021) 118, arXiv: 2106.11683 [hep-ex].
[181] ATLAS Collaboration, Roadmap towards future combinations and Effective Field Theory
interpretations of top+X processes, ATL-PHYS-PUB-2023-030, 2023,
url: https://cds.cern.ch/record/2872789.
[182] Particle Data Group, M. Tanabashi et al., Review of Particle Physics,
Phys. Rev. D 98 (3 2018) 030001.
[183] A. H. Hoang, What is the Top Quark Mass?, Ann. Rev. Nucl. Part. Sci. 70 (2020) 225,
arXiv: 2004.12915 [hep-ph].
[184] ATLAS Collaboration,
A precise interpretation for the top quark mass parameter in ATLAS Monte Carlo simulation,
ATL-PHYS-PUB-2021-034, 2021, url: https://cds.cern.ch/record/2777332.
[185] ATLAS Collaboration, Measurement of the top quark mass in the 𝑡 𝑡¯ → lepton+jets channel from

𝑠 = 8 TeV ATLAS data and combination with previous results, Eur. Phys. J. C 79 (2019) 290,
arXiv: 1810.01772 [hep-ex].
[186] CMS Collaboration,

Measurement of the top quark mass using proton–proton data at 𝑠 = 7 and 8 TeV,
Phys. Rev. D 93 (2016) 072004, arXiv: 1509.04044 [hep-ex].

64
[187] Tevatron Electroweak Working Group,
Combination of CDF and D0 results on the mass of the top quark using up 9.7 fb−1 at the Tevatron,
(2016), arXiv: 1608.01881 [hep-ex].
[188] ATLAS Collaboration,

Measurement of the top quark mass in the 𝑡 𝑡¯ → dilepton channel from 𝑠 = 8 TeV ATLAS data,
Phys. Lett. B 761 (2016) 350, arXiv: 1606.02179 [hep-ex].
[189] ATLAS and CMS Collaborations, Combination of measurements of the top quark mass from data

collected by the ATLAS and CMS experiments at 𝑠 = 7 and 8 TeV, (2024),
arXiv: 2402.08713 [hep-ex].
[190] ATLAS Collaboration, Measurement of the top-quark mass using a leptonic invariant mass in 𝑝 𝑝

collisions at 𝑠 = 13 TeV with the ATLAS detector, JHEP 06 (2023) 019,
arXiv: 2209.00583 [hep-ex].
[191] ATLAS Collaboration, Measurement of the top-quark mass in 𝑡 𝑡¯ → dilepton events with the
ATLAS experiment using the template method in 13 TeV 𝑝 𝑝 collision data,
ATLAS-CONF-2022-058, 2022, url: https://cds.cern.ch/record/2826701.
[192] ATLAS Collaboration, Measurement of the top-quark mass in 𝑡 𝑡¯ + 1-jet events collected with the

ATLAS detector in 𝑝 𝑝 collisions at 𝑠 = 8 TeV, JHEP 11 (2019) 150,
arXiv: 1905.02302 [hep-ex].
[193] Particle Data Group, R. L. Workman et al., Review of Particle Physics, PTEP 2022 (2022) 083C01.
[194] A. Czarnecki, J. G. Körner and J. H. Piclum,
Helicity fractions of 𝑊 bosons from top quark decays at next-to-next-to-leading order in QCD,
Phys. Rev. D 81 (2010) 111503, arXiv: 1005.2625 [hep-ph].
[195] ATLAS Collaboration, Measurement of the polarisation of W bosons produced in top-quark

decays using dilepton events at 𝑠 = 13 TeV with the ATLAS experiment,
Phys. Lett. B 843 (2023) 137829, arXiv: 2209.14903 [hep-ex].
[196] A. Brandenburg, Z. G. Si and P. Uwer,
QCD-corrected spin analyzing power of jets in decays of polarized top quarks,
Phys. Lett. B 539 (2002) 235, arXiv: hep-ph/0205023.
[197] ATLAS Collaboration, Observation of spin correlation in 𝑡 𝑡¯ events from pp collisions at sqrt(s) =
7 TeV using the ATLAS detector, Phys. Rev. Lett. 108 (2012) 212001,
arXiv: 1203.4081 [hep-ex].
[198] ATLAS Collaboration, Measurements of spin correlation in top-antitop quark events from

proton-proton collisions at 𝑠 = 7 TeV using the ATLAS detector, Phys. Rev. D 90 (2014) 112016,
arXiv: 1407.4314 [hep-ex].
[199] ATLAS Collaboration, Measurement of Spin Correlation in Top-Antitop Quark Events and Search

for Top Squark Pair Production in pp Collisions at 𝑠 = 8 TeV Using the ATLAS Detector,
Phys. Rev. Lett. 114 (2015) 142001, arXiv: 1412.4742 [hep-ex].
[200] ATLAS Collaboration, Measurements of top quark spin observables in 𝑡𝑡 events using dilepton

final states in 𝑠 = 8 TeV pp collisions with the ATLAS detector, JHEP 03 (2017) 113,
arXiv: 1612.07004 [hep-ex].
[201] ATLAS Collaboration, Measurements of top-quark pair spin correlations in the 𝑒𝜇 channel at

𝑠 = 13 TeV using 𝑝 𝑝 collisions in the ATLAS detector, Eur. Phys. J. C 80 (2020) 754,
arXiv: 1903.07570 [hep-ex].

65
[202] W. Bernreuther, D. Heisler and Z.-G. Si, A set of top quark spin correlation and polarization
observables for the LHC: Standard Model predictions and new physics contributions,
JHEP 12 (2015) 1, arXiv: 1508.05271 [hep-ph].
[203] Y. Afik and J. R. M. de Nova, Entanglement and quantum tomography with top quarks at the LHC,
Eur. Phys. J. Plus 136 (2021) 907, arXiv: 2003.02280 [quant-ph].
[204] M. Fabbrichesi, R. Floreanini and G. Panizzo,
Testing Bell Inequalities at the LHC with Top-Quark Pairs, Phys. Rev. Lett. 127 (2021) 161801,
arXiv: 2102.11883 [hep-ph].
[205] C. Severi, C. D. E. Boschi, F. Maltoni and M. Sioli,
Quantum tops at the LHC: from entanglement to Bell inequalities, Eur. Phys. J. C 82 (2022) 285,
arXiv: 2110.10112 [hep-ph].
[206] J. A. Aguilar-Saavedra and J. A. Casas,
Improved tests of entanglement and Bell inequalities with LHC tops, Eur. Phys. J. C 82 (2022) 666,
arXiv: 2205.00542 [hep-ph].
[207] R. Ashby-Pickering, A. J. Barr and A. Wierzchucka, Quantum state tomography, entanglement
detection and Bell violation prospects in weak decays of massive particles, JHEP 05 (2023) 020,
arXiv: 2209.13990 [quant-ph].
[208] ATLAS Collaboration,
Observation of quantum entanglement in top-quark pairs using the ATLAS detector, (2023),
arXiv: 2311.07288 [hep-ex].
[209] B. A. Betchart, R. Demina and A. Harel,
Analytic solutions for neutrino momenta in decay of top quarks,
Nucl. Instrum. Meth. A 736 (2014) 169, arXiv: 1305.1878 [hep-ph].
[210] Y. Afik and J. R. M. de Nova, Quantum Discord and Steering in Top Quarks at the LHC,
Phys. Rev. Lett. 130 (2023) 221801, arXiv: 2209.03969 [quant-ph].
[211] Y. Afik and J. R. M. de Nova, Quantum information with top quarks in QCD,
Quantum 6 (2022) 820, arXiv: 2203.05582 [quant-ph].
[212] O. Antuñano, J. H. Kühn and G. Rodrigo,
Top quarks, axigluons, and charge asymmetries at hadron colliders,
Phys. Rev. D 77 (2008) 014003, arXiv: 0709.1652 [hep-ph].
[213] P. H. Frampton, J. Shu and K. Wang,
Axigluon as possible explanation for 𝑝 𝑝¯ → 𝑡 𝑡¯ forward-backward asymmetry,
Phys. Lett. B 683 (2010) 294, arXiv: 0911.2955 [hep-ph].
[214] J. L. Rosner, Prominent decay modes of a leptophobic 𝑍 ′ , Phys. Lett. B 387 (1996) 113,
arXiv: hep-ph/9607207.
[215] J. A. Aguilar-Saavedra and M. Pérez-Victoria, Asymmetries in t 𝑡¯ production: LHC versus Tevatron,
Phys. Rev. D 84 (2011) 115013, arXiv: 1105.4606 [hep-ph].
[216] M. P. Roselló and M. Vos,
Constraints on four-fermion interactions from the 𝑡 𝑡¯ charge asymmetry at hadron colliders,
Eur. Phys. J. C 76 (2016) 200, arXiv: 1512.07542 [hep-ex].
[217] J. A. Aguilar-Saavedra, A. Juste and F. Rubbo, Boosting the 𝑡 𝑡¯ charge asymmetry,
Phys. Lett. B 707 (2012) 92, arXiv: 1109.3710 [hep-ph].

66
[218] ATLAS and CMS Collaborations, Combination of inclusive and differential tt charge asymmetry

measurements using ATLAS and CMS data at 𝑠 = 7 and 8 TeV, JHEP 04 (2018) 033,
arXiv: 1709.05327 [hep-ex].
[219] ATLAS Collaboration,

Evidence for the charge asymmetry in pp → 𝑡𝑡 production at 𝑠 = 13 TeV with the ATLAS detector,
JHEP 08 (2023) 077, arXiv: 2208.12095 [hep-ex].
[220] M. Czakon et al.,
Top-quark charge asymmetry at the LHC and Tevatron through NNLO QCD and NLO EW,
Phys. Rev. D 98 (2018) 014003, arXiv: 1711.03945 [hep-ph].
[221] W. Bernreuther and Z.-G. Si,
Top quark and leptonic charge asymmetries for the Tevatron and LHC,
Phys. Rev. D 86 (2012) 034026, arXiv: 1205.6580 [hep-ph].
[222] S. Berge and S. Westhoff,
Top-Quark Charge Asymmetry Goes Forward: Two New Observables for Hadron Colliders,
JHEP 07 (2013) 179, arXiv: 1305.3272 [hep-ph].
[223] ATLAS Collaboration, Measurement of the energy asymmetry in 𝑡 𝑡¯ 𝑗 production at 13 TeV with the
ATLAS experiment and interpretation in the SMEFT framework, Eur. Phys. J. C 82 (2022) 374,
arXiv: 2110.05453 [hep-ex].
[224] F. Maltoni, M. L. Mangano, I. Tsinikos and M. Zaro,
Top-quark charge asymmetry and polarization in 𝑡 𝑡¯𝑊 ± production at the LHC,
Phys. Lett. B 736 (2014) 252, arXiv: 1406.3262 [hep-ph].
[225] G. Bevilacqua et al.,
NLO QCD corrections to off-shell 𝑡 𝑡¯𝑊 ± production at the LHC: correlations and asymmetries,
Eur. Phys. J. C 81 (2021) 675, arXiv: 2012.01363 [hep-ph].
[226] ATLAS Collaboration, Search for leptonic charge asymmetry in 𝑡𝑡𝑊 production in final states with

three leptons at 𝑠 = 13 TeV, JHEP 07 (2023) 033, arXiv: 2301.04245 [hep-ex].
[227] J. A. Aguilar-Saavedra, E. Álvarez, A. Juste and F. Rubbo,
Shedding light on the 𝑡 𝑡¯ asymmetry: the photon handle, JHEP 04 (2014) 188,
arXiv: 1402.3598 [hep-ph].
[228] D. Pagani, H.-S. Shao, I. Tsinikos and M. Zaro,
Automated EW corrections with isolated photons: t𝑡𝛾, t𝑡𝛾𝛾 and t𝛾j as case studies,
JHEP 09 (2021) 155, arXiv: 2106.02059 [hep-ph].
[229] J. Bergner and M. Schulze, The top quark charge asymmetry in 𝑡 𝑡¯𝛾 production at the LHC,
Eur. Phys. J. C 79 (2019) 189, arXiv: 1812.10535 [hep-ph].
[230] ATLAS Collaboration, Measurement of the charge asymmetry in top-quark pair production in
association with a photon with the ATLAS experiment, Phys. Lett. B 843 (2023) 137848,
arXiv: 2212.10552 [hep-ex].
[231] ATLAS Collaboration, Measurement of colour flow with the jet pull angle in 𝑡 𝑡¯ events using the

ATLAS detector at 𝑠 = 8 TeV, Phys. Lett. B 750 (2015) 475, arXiv: 1506.05629 [hep-ex].
[232] ATLAS Collaboration, Measurement of colour flow using jet-pull observables in 𝑡 𝑡¯ events with the

ATLAS experiment at 𝑠 = 13 TeV, Eur. Phys. J. C 78 (2018) 847, arXiv: 1805.02935 [hep-ex].

67
[233] J. Gallicchio and M. D. Schwartz, Seeing in Color: Jet Superstructure,
Phys. Rev. Lett. 105 (2010) 022001, arXiv: 1001.5027 [hep-ph].
[234] Z. Nagy and D. E. Soper, Parton showers with quantum interference: Leading color, spin averaged,
JHEP 03 (2008) 030, arXiv: 0801.1917 [hep-ph].
[235] A. Buckley et al., General-purpose event generators for LHC physics, Phys. Rept. 504 (2011) 145,
arXiv: 1101.2599 [hep-ph].
[236] ATLAS Collaboration, Measurements of observables sensitive to colour reconnection in 𝑡 𝑡¯ events

with the ATLAS detector at 𝑠 = 13 TeV, Eur. Phys. J. C 83 (2023) 518,
arXiv: 2209.07874 [hep-ex].
[237] ATLAS Collaboration,
Performance of top-quark and 𝑊-boson tagging with ATLAS in Run 2 of the LHC,
Eur. Phys. J. C 79 (2019) 375, arXiv: 1808.07858 [hep-ex].
[238] CMS Collaboration, Identification of heavy, energetic, hadronically decaying particles using
machine-learning techniques, JINST 15 (2020) P06005, arXiv: 2004.08262 [hep-ex].
[239] A. Butter et al., The Machine Learning landscape of top taggers,
SciPost Phys. 7 (2019) 014, ed. by G. Kasieczka and T. Plehn, arXiv: 1902.09914 [hep-ph].
[240] ATLAS Collaboration, Measurement of jet substructure in boosted 𝑡 𝑡¯ events with the ATLAS
detector using 140 fb−1 of 13 TeV 𝑝 𝑝 collisions, (2023), arXiv: 2312.03797 [hep-ex].
[241] S. Schael et al.,
Electroweak Measurements in Electron-Positron Collisions at W-Boson-Pair Energies at LEP,
Phys. Rept. 532 (2013) 119, arXiv: 1302.3415 [hep-ex].
[242] ATLAS Collaboration,
Test of the universality of 𝜏 and 𝜇 lepton couplings in 𝑊-boson decays with the ATLAS detector,
Nature Phys. 17 (2021) 813, arXiv: 2007.14040 [hep-ex].
[243] ATLAS Collaboration, Precise test of lepton flavour universality in 𝑊-boson decays into muons

and electrons in 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector, (2024),
arXiv: 2403.02133 [hep-ex].
[244] ALEPH, DELPHI, L3, OPAL and SLD Collaborations, LEP Electroweak Working Group, SLD
Electroweak Group, SLD Heavy Flavour Group,
Precision electroweak measurements on the 𝑍 resonance, Phys. Rept. 427 (2006) 257,
arXiv: hep-ex/0509008.
[245] S. L. Glashow, J. Iliopoulos and L. Maiani, Weak Interactions with Lepton-Hadron Symmetry,
Phys. Rev. D 2 (1970) 1285.
[246] J. A. Aguilar-Saavedra,
Top flavor-changing neutral interactions: Theoretical expectations and experimental detection,
Acta Phys. Polon. B 35 (2004) 2695, ed. by F. del Aguila, R. Pittau, A. Djouadi and
C. G. Papadopoulos, arXiv: hep-ph/0409342.
[247] J. A. Aguilar-Saavedra, Effects of mixing with quark singlets, Phys. Rev. D 67 (2003) 035003,
arXiv: hep-ph/0210112, Erratum: Phys. Rev. D 69 (2004) 099901.
[248] D. Atwood, L. Reina and A. Soni,
Phenomenology of two Higgs doublet models with flavor-changing neutral currents,
Phys. Rev. D 55 (1997) 3156, arXiv: hep-ph/9609279.

68
[249] J. J. Cao et al., Supersymmetry-induced flavor-changing neutral-current top-quark processes at the
CERN Large Hadron Collider, Phys. Rev. D 75 (2007) 075021, arXiv: hep-ph/0702264.
[250] J. M. Yang, B.-L. Young and X. Zhang,
Flavor-changing top quark decays in 𝑅-parity-violating supersymmetric models,
Phys. Rev. D 58 (1998) 055001, arXiv: hep-ph/9705341.
[251] K. Agashe, G. Perez and A. Soni,
Collider Signals of Top Quark Flavor Violation from a Warped Extra Dimension,
Phys. Rev. D 75 (2007) 015002, arXiv: hep-ph/0606293.
[252] P. Q. Hung, Y.-X. Lin, C. S. Nugroho and T.-C. Yuan,
Top Quark Rare Decays via Loop-Induced FCNC Interactions in Extended Mirror Fermion Model,
Nucl. Phys. B 927 (2018) 166, arXiv: 1709.01690 [hep-ph].
[253] G. Durieux, F. Maltoni and C. Zhang, Global approach to top-quark flavor-changing interactions,
Phys. Rev. D 91 (2015) 074017, arXiv: 1412.7166 [hep-ph].

[254] ATLAS Collaboration, Search for top quark decays 𝑡 → 𝑞𝐻, with 𝐻 → 𝛾𝛾, in 𝑠 = 13 TeV 𝑝 𝑝
collisions using the ATLAS detector, JHEP 10 (2017) 129, arXiv: 1707.01404 [hep-ex].
[255] ATLAS Collaboration,
Search for flavor-changing neutral currents in top quark decays 𝑡 → 𝐻𝑐 and 𝑡 → 𝐻𝑢 in

multilepton final states in proton–proton collisions at 𝑠 = 13 TeV with the ATLAS detector,
Phys. Rev. D 98 (2018) 032002, arXiv: 1805.03483 [hep-ex].
[256] ATLAS Collaboration, Search for top-quark decays 𝑡 → 𝐻𝑞 with 36 fb−1 of 𝑝 𝑝 collision data at

𝑠 = 13 TeV with the ATLAS detector, JHEP 05 (2019) 123, arXiv: 1812.11568 [hep-ex].
[257] ATLAS Collaboration, Search for flavor-changing neutral 𝑡𝑞𝐻 interactions with 𝐻 → 𝛾𝛾 in 𝑝 𝑝

collisions at 𝑠 = 13 TeV using the ATLAS detector, JHEP 12 (2023) 195,
arXiv: 2309.12817 [hep-ex].
[258] ATLAS Collaboration,
Search for flavour-changing neutral-current couplings between the top quark and the Higgs boson
in multi-lepton final states in 13 TeV 𝑝 𝑝 collisions with the ATLAS detector, (2024),
arXiv: 2404.02123 [hep-ex].
[259] ATLAS Collaboration,
Search for flavour-changing neutral current interactions of the top quark and the Higgs boson in

events with a pair of 𝜏-leptons in 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector,
JHEP 06 (2023) 155, arXiv: 2208.11415 [hep-ex].
[260] ATLAS Collaboration, Search for a new scalar resonance in flavour-changing neutral-current
¯ in proton–proton collisions at √𝑠 = 13 TeV
top-quark decays 𝑡 → 𝑞𝑋 (𝑞 = 𝑢, 𝑐), with 𝑋 → 𝑏 𝑏,
with the ATLAS detector, JHEP 07 (2023) 199, arXiv: 2301.03902 [hep-ex].
[261] ATLAS Collaboration, Evidence for the associated production of the Higgs boson and a top quark
pair with the ATLAS detector, Phys. Rev. D 97 (2018) 072003, arXiv: 1712.08891 [hep-ex].
[262] ATLAS Collaboration, Search for the Standard Model Higgs boson produced in association with

top quarks and decaying into 𝑏 𝑏¯ in 𝑝 𝑝 collisions at 𝑠 = 8 TeV with the ATLAS detector,
Eur. Phys. J. C 75 (2015) 349, arXiv: 1503.05066 [hep-ex].
[263] ATLAS Collaboration, Search for the standard model Higgs boson produced in association with

top quarks and decaying into a 𝑏 𝑏¯ pair in 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector,
Phys. Rev. D 97 (2018) 072016, arXiv: 1712.08895 [hep-ex].

69
[264] ATLAS Collaboration, Search for flavour-changing neutral-current interactions of a top quark and

a gluon in 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS detector, Eur. Phys. J. C 82 (2022) 334,
arXiv: 2112.01302 [hep-ex].
[265] ATLAS Collaboration, Search for flavour-changing neutral currents in processes with one top

quark and a photon using 81 fb−1 of 𝑝 𝑝 collisions at 𝑠 = 13 TeV with the ATLAS experiment,
Phys. Lett. B 800 (2020) 135082, arXiv: 1908.08461 [hep-ex].
[266] ATLAS Collaboration, Search for flavour-changing neutral-current couplings between the top

quark and the photon with the ATLAS detector at 𝑠 = 13 TeV, Phys. Lett. B 842 (2023) 137379,
arXiv: 2205.02537 [hep-ex].
[267] B. Grzadkowski, M. Iskrzyński, M. Misiak and J. Rosiek,
Dimension-Six Terms in the Standard Model Lagrangian, JHEP 10 (2010) 085,
arXiv: 1008.4884 [hep-ph].
[268] D. Barducci et al.,
Interpreting top-quark LHC measurements in the standard-model effective field theory,
(2018), ed. by J. A. Aguilar-Saavedra et al., arXiv: 1802.07237 [hep-ph].
[269] ATLAS Collaboration, Search for flavour-changing neutral current top-quark decays 𝑡 → 𝑞𝑍 in

proton–proton collisions at 𝑠 = 13 TeV with the ATLAS detector, JHEP 07 (2018) 176,
arXiv: 1803.09923 [hep-ex].
[270] ATLAS Collaboration, Search for flavor-changing neutral-current couplings between the top quark

and the 𝑍 boson with proton–proton collisions at 𝑠 = 13 TeV with the ATLAS detector,
Phys. Rev. D 108 (2023) 032019, arXiv: 2301.11605 [hep-ex].
[271] ATLAS Collaboration, ATLAS Computing Acknowledgements, ATL-SOFT-PUB-2023-001, 2023,
url: https://cds.cern.ch/record/2869272.

70
The ATLAS Collaboration

G. Aad 103 , E. Aakvaag 16 , B. Abbott 121 , S. Abdelhameed 117a , K. Abeling 55 , N.J. Abicht 49 ,
S.H. Abidi 29 , M. Aboelela 44 , A. Aboulhorma 35e , H. Abramowicz 152 , H. Abreu 151 ,
Y. Abulaiti 118 , B.S. Acharya 69a,69b,k , A. Ackermann 63a , C. Adam Bourdarios 4 ,
L. Adamczyk 86a , S.V. Addepalli 26 , M.J. Addison 102 , J. Adelman 116 , A. Adiguzel 21c ,
T. Adye 135 , A.A. Affolder 137 , Y. Afik 39 , M.N. Agaras 13 , J. Agarwala 73a,73b ,
A. Aggarwal 101 , C. Agheorghiesei 27c , A. Ahmad 36 , F. Ahmadov 38,x , W.S. Ahmed 105 ,
S. Ahuja 96 , X. Ai 62e , G. Aielli 76a,76b , A. Aikot 164 , M. Ait Tamlihat 35e , B. Aitbenchikh 35a ,
M. Akbiyik 101 , T.P.A. Åkesson 99 , A.V. Akimov 37 , D. Akiyama 169 , N.N. Akolkar 24 ,
S. Aktas 21a , K. Al Khoury 41 , G.L. Alberghi 23b , J. Albert 166 , P. Albicocco 53 , G.L. Albouy 60 ,
S. Alderweireldt 52 , Z.L. Alegria 122 , M. Aleksa 36 , I.N. Aleksandrov 38 , C. Alexa 27b ,
T. Alexopoulos 10 , F. Alfonsi 23b , M. Algren 56 , M. Alhroob 142 , B. Ali 133 , H.M.J. Ali 92 ,
S. Ali 31 , S.W. Alibocus 93 , M. Aliev 33c , G. Alimonti 71a , W. Alkakhi 55 , C. Allaire 66 ,
B.M.M. Allbrooke 147 , J.F. Allen 52 , C.A. Allendes Flores 138f , P.P. Allport 20 , A. Aloisio 72a,72b ,
F. Alonso 91 , C. Alpigiani 139 , Z.M.K. Alsolami 92 , M. Alvarez Estevez 100 ,
A. Alvarez Fernandez 101 , M. Alves Cardoso 56 , M.G. Alviggi 72a,72b , M. Aly 102 ,
Y. Amaral Coutinho 83b , A. Ambler 105 , C. Amelung36 , M. Amerl 102 , C.G. Ames 110 ,
D. Amidei 107 , K.J. Amirie 156 , S.P. Amor Dos Santos 131a , K.R. Amos 164 , S. An84 ,
V. Ananiev 126 , C. Anastopoulos 140 , T. Andeen 11 , J.K. Anders 36 , S.Y. Andrean 47a,47b ,
A. Andreazza 71a,71b , S. Angelidakis 9 , A. Angerami 41,z , A.V. Anisenkov 37 , A. Annovi 74a ,
C. Antel 56 , E. Antipov 146 , M. Antonelli 53 , F. Anulli 75a , M. Aoki 84 , T. Aoki 154 ,
M.A. Aparo 147 , L. Aperio Bella 48 , C. Appelt 18 , A. Apyan 26 , S.J. Arbiol Val 87 ,
C. Arcangeletti 53 , A.T.H. Arce 51 , E. Arena 93 , J-F. Arguin 109 , S. Argyropoulos 54 ,
J.-H. Arling 48 , O. Arnaez 4 , H. Arnold 115 , G. Artoni 75a,75b , H. Asada 112 , K. Asai 119 ,
S. Asai 154 , N.A. Asbah 36 , K. Assamagan 29 , R. Astalos 28a , K.S.V. Astrand 99 , S. Atashi 160 ,
R.J. Atkin 33a , M. Atkinson163 , H. Atmani35f , P.A. Atmasiddha 129 , K. Augsten 133 ,
S. Auricchio 72a,72b , A.D. Auriol 20 , V.A. Austrup 102 , G. Avolio 36 , K. Axiotis 56 ,
G. Azuelos 109,ad , D. Babal 28b , H. Bachacou 136 , K. Bachas 153,o , A. Bachiu 34 ,
F. Backman 47a,47b , A. Badea 39 , T.M. Baer 107 , P. Bagnaia 75a,75b , M. Bahmani 18 ,
D. Bahner 54 , K. Bai 124 , J.T. Baines 135 , L. Baines 95 , O.K. Baker 173 , E. Bakos 15 ,
D. Bakshi Gupta 8 , V. Balakrishnan 121 , R. Balasubramanian 115 , E.M. Baldin 37 , P. Balek 86a ,
E. Ballabene 23b,23a , F. Balli 136 , L.M. Baltes 63a , W.K. Balunas 32 , J. Balz 101 ,
I. Bamwidhi 117b , E. Banas 87 , M. Bandieramonte 130 , A. Bandyopadhyay 24 , S. Bansal 24 ,
L. Barak 152 , M. Barakat 48 , E.L. Barberio 106 , D. Barberis 57b,57a , M. Barbero 103 ,
M.Z. Barel 115 , K.N. Barends 33a , T. Barillari 111 , M-S. Barisits 36 , T. Barklow 144 , P. Baron 123 ,
D.A. Baron Moreno 102 , A. Baroncelli 62a , G. Barone 29 , A.J. Barr 127 , J.D. Barr 97 ,
F. Barreiro 100 , J. Barreiro Guimarães da Costa 14a , U. Barron 152 , M.G. Barros Teixeira 131a ,
S. Barsov 37 , F. Bartels 63a , R. Bartoldus 144 , A.E. Barton 92 , P. Bartos 28a , A. Basan 101 ,
M. Baselga 49 , A. Bassalat 66,b , M.J. Basso 157a , R. Bate 165 , R.L. Bates 59 , S. Batlamous100 ,
B. Batool 142 , M. Battaglia 137 , D. Battulga 18 , M. Bauce 75a,75b , M. Bauer 36 , P. Bauer 24 ,
L.T. Bazzano Hurrell 30 , J.B. Beacham 51 , T. Beau 128 , J.Y. Beaucamp 91 , P.H. Beauchemin 159 ,
P. Bechtle 24 , H.P. Beck 19,n , K. Becker 168 , A.J. Beddall 82 , V.A. Bednyakov 38 , C.P. Bee 146 ,
L.J. Beemster 15 , T.A. Beermann 36 , M. Begalli 83d , M. Begel 29 , A. Behera 146 , J.K. Behr 48 ,
J.F. Beirer 36 , F. Beisiegel 24 , M. Belfkir 117b , G. Bella 152 , L. Bellagamba 23b , A. Bellerive 34 ,
P. Bellos 20 , K. Beloborodov 37 , D. Benchekroun 35a , F. Bendebba 35a , Y. Benhammou 152 ,

71
K.C. Benkendorfer 61 , L. Beresford 48 , M. Beretta 53 , E. Bergeaas Kuutmann 162 , N. Berger 4 ,
B. Bergmann 133 , J. Beringer 17a , G. Bernardi 5 , C. Bernius 144 , F.U. Bernlochner 24 ,
F. Bernon 36,103 , A. Berrocal Guardia 13 , T. Berry 96 , P. Berta 134 , A. Berthold 50 , S. Bethke 111 ,
A. Betti 75a,75b , A.J. Bevan 95 , N.K. Bhalla 54 , M. Bhamjee 33c , S. Bhatta 146 ,
D.S. Bhattacharya 167 , P. Bhattarai 144 , K.D. Bhide 54 , V.S. Bhopatkar 122 , R.M. Bianchi 130 ,
G. Bianco 23b,23a , O. Biebel 110 , R. Bielski 124 , M. Biglietti 77a , C.S. Billingsley44 , M. Bindi 55 ,
A. Bingul 21b , C. Bini 75a,75b , A. Biondini 93 , C.J. Birch-sykes 102 , G.A. Bird 32 , M. Birman 170 ,
M. Biros 134 , S. Biryukov 147 , T. Bisanz 49 , E. Bisceglie 43b,43a , J.P. Biswal 135 , D. Biswas 142 ,
I. Bloch 48 , A. Blue 59 , U. Blumenschein 95 , J. Blumenthal 101 , V.S. Bobrovnikov 37 ,
M. Boehler 54 , B. Boehm 167 , D. Bogavac 36 , A.G. Bogdanchikov 37 , C. Bohm 47a ,
V. Boisvert 96 , P. Bokan 36 , T. Bold 86a , M. Bomben 5 , M. Bona 95 , M. Boonekamp 136 ,
C.D. Booth 96 , A.G. Borbély 59 , I.S. Bordulev 37 , H.M. Borecka-Bielska 109 , G. Borissov 92 ,
D. Bortoletto 127 , D. Boscherini 23b , M. Bosman 13 , J.D. Bossio Sola 36 , K. Bouaouda 35a ,
N. Bouchhar 164 , L. Boudet 4 , J. Boudreau 130 , E.V. Bouhova-Thacker 92 , D. Boumediene 40 ,
R. Bouquet 57b,57a , A. Boveia 120 , J. Boyd 36 , D. Boye 29 , I.R. Boyko 38 , L. Bozianu 56 ,
J. Bracinik 20 , N. Brahimi 4 , G. Brandt 172 , O. Brandt 32 , F. Braren 48 , B. Brau 104 ,
J.E. Brau 124 , R. Brener 170 , L. Brenner 115 , R. Brenner 162 , S. Bressler 170 , D. Britton 59 ,
D. Britzger 111 , I. Brock 24 , G. Brooijmans 41 , E. Brost 29 , L.M. Brown 166 , L.E. Bruce 61 ,
T.L. Bruckler 127 , P.A. Bruckman de Renstrom 87 , B. Brüers 48 , A. Bruni 23b , G. Bruni 23b ,
M. Bruschi 23b , N. Bruscino 75a,75b , T. Buanes 16 , Q. Buat 139 , D. Buchin 111 , A.G. Buckley 59 ,
O. Bulekov 37 , B.A. Bullard 144 , S. Burdin 93 , C.D. Burgard 49 , A.M. Burger 36 ,
B. Burghgrave 8 , O. Burlayenko 54 , J.T.P. Burr 32 , J.C. Burzynski 143 , E.L. Busch 41 ,
V. Büscher 101 , P.J. Bussey 59 , J.M. Butler 25 , C.M. Buttar 59 , J.M. Butterworth 97 ,
W. Buttinger 135 , C.J. Buxo Vazquez 108 , A.R. Buzykaev 37 , S. Cabrera Urbán 164 ,
L. Cadamuro 66 , D. Caforio 58 , H. Cai 130 , Y. Cai 14a,14e , Y. Cai 14c , V.M.M. Cairo 36 ,
O. Cakir 3a , N. Calace 36 , P. Calafiura 17a , G. Calderini 128 , P. Calfayan 68 , G. Callea 59 ,
L.P. Caloba83b , D. Calvet 40 , S. Calvet 40 , M. Calvetti 74a,74b , R. Camacho Toro 128 ,
S. Camarda 36 , D. Camarero Munoz 26 , P. Camarri 76a,76b , M.T. Camerlingo 72a,72b ,
D. Cameron 36 , C. Camincher 166 , M. Campanelli 97 , A. Camplani 42 , V. Canale 72a,72b ,
A.C. Canbay 3a , E. Canonero 96 , J. Cantero 164 , Y. Cao 163 , F. Capocasa 26 , M. Capua 43b,43a ,
A. Carbone 71a,71b , R. Cardarelli 76a , J.C.J. Cardenas 8 , G. Carducci 43b,43a , T. Carli 36 ,
G. Carlino 72a , J.I. Carlotto 13 , B.T. Carlson 130,p , E.M. Carlson 166,157a , J. Carmignani 93 ,
L. Carminati 71a,71b , A. Carnelli 136 , M. Carnesale 75a,75b , S. Caron 114 , E. Carquin 138f ,
S. Carrá 71a , G. Carratta 23b,23a , A.M. Carroll 124 , T.M. Carter 52 , M.P. Casado 13,h ,
M. Caspar 48 , F.L. Castillo 4 , L. Castillo Garcia 13 , V. Castillo Gimenez 164 , N.F. Castro 131a,131e ,
A. Catinaccio 36 , J.R. Catmore 126 , T. Cavaliere 4 , V. Cavaliere 29 , N. Cavalli 23b,23a ,
Y.C. Cekmecelioglu 48 , E. Celebi 21a , S. Cella 36 , F. Celli 127 , M.S. Centonze 70a,70b ,
V. Cepaitis 56 , K. Cerny 123 , A.S. Cerqueira 83a , A. Cerri 147 , L. Cerrito 76a,76b , F. Cerutti 17a ,
B. Cervato 142 , A. Cervelli 23b , G. Cesarini 53 , S.A. Cetin 82 , D. Chakraborty 116 , J. Chan 17a ,
W.Y. Chan 154 , J.D. Chapman 32 , E. Chapon 136 , B. Chargeishvili 150b , D.G. Charlton 20 ,
M. Chatterjee 19 , C. Chauhan 134 , Y. Che 14c , S. Chekanov 6 , S.V. Chekulaev 157a ,
G.A. Chelkov 38,a , A. Chen 107 , B. Chen 152 , B. Chen 166 , H. Chen 14c , H. Chen 29 ,
J. Chen 62c , J. Chen 143 , M. Chen 127 , S. Chen 154 , S.J. Chen 14c , X. Chen 62c,136 ,
X. Chen 14b,ac , Y. Chen 62a , C.L. Cheng 171 , H.C. Cheng 64a , S. Cheong 144 , A. Cheplakov 38 ,
E. Cheremushkina 48 , E. Cherepanova 115 , R. Cherkaoui El Moursli 35e , E. Cheu 7 , K. Cheung 65 ,
L. Chevalier 136 , V. Chiarella 53 , G. Chiarelli 74a , N. Chiedde 103 , G. Chiodini 70a ,
A.S. Chisholm 20 , A. Chitan 27b , M. Chitishvili 164 , M.V. Chizhov 38 , K. Choi 11 , Y. Chou 139 ,

72
E.Y.S. Chow 114 , K.L. Chu 170 , M.C. Chu 64a , X. Chu 14a,14e , J. Chudoba 132 ,
J.J. Chwastowski 87 , D. Cieri 111 , K.M. Ciesla 86a , V. Cindro 94 , A. Ciocio 17a , F. Cirotto 72a,72b ,
Z.H. Citron 170 , M. Citterio 71a , D.A. Ciubotaru27b , A. Clark 56 , P.J. Clark 52 , C. Clarry 156 ,
J.M. Clavijo Columbie 48 , S.E. Clawson 48 , C. Clement 47a,47b , J. Clercx 48 , Y. Coadou 103 ,
M. Cobal 69a,69c , A. Coccaro 57b , R.F. Coelho Barrue 131a , R. Coelho Lopes De Sa 104 ,
S. Coelli 71a , B. Cole 41 , J. Collot 60 , P. Conde Muiño 131a,131g , M.P. Connell 33c ,
S.H. Connell 33c , E.I. Conroy 127 , F. Conventi 72a,ae , H.G. Cooke 20 , A.M. Cooper-Sarkar 127 ,
F.A. Corchia 23b,23a , A. Cordeiro Oudot Choi 128 , L.D. Corpe 40 , M. Corradi 75a,75b ,
F. Corriveau 105,v , A. Cortes-Gonzalez 18 , M.J. Costa 164 , F. Costanza 4 , D. Costanzo 140 ,
B.M. Cote 120 , J. Couthures4 , G. Cowan 96 , K. Cranmer 171 , D. Cremonini 23b,23a ,
S. Crépé-Renaudin 60 , F. Crescioli 128 , M. Cristinziani 142 , M. Cristoforetti 78a,78b , V. Croft 115 ,
J.E. Crosby 122 , G. Crosetti 43b,43a , A. Cueto 100 , Z. Cui 7 , W.R. Cunningham 59 , F. Curcio 164 ,
J.R. Curran 52 , P. Czodrowski 36 , M.M. Czurylo 36 , M.J. Da Cunha Sargedas De Sousa 57b,57a ,
J.V. Da Fonseca Pinto 83b , C. Da Via 102 , W. Dabrowski 86a , T. Dado 49 , S. Dahbi 149 ,
T. Dai 107 , D. Dal Santo 19 , C. Dallapiccola 104 , M. Dam 42 , G. D’amen 29 , V. D’Amico 110 ,
J. Damp 101 , J.R. Dandoy 34 , D. Dannheim 36 , M. Danninger 143 , V. Dao 36 , G. Darbo 57b ,
S.J. Das 29,af , F. Dattola 48 , S. D’Auria 71a,71b , A. D’avanzo 72a,72b , C. David 33a , T. Davidek 134 ,
I. Dawson 95 , H.A. Day-hall 133 , K. De 8 , R. De Asmundis 72a , N. De Biase 48 ,
S. De Castro 23b,23a , N. De Groot 114 , P. de Jong 115 , H. De la Torre 116 , A. De Maria 14c ,
A. De Salvo 75a , U. De Sanctis 76a,76b , F. De Santis 70a,70b , A. De Santo 147 ,
J.B. De Vivie De Regie 60 , D.V. Dedovich38 , J. Degens 93 , A.M. Deiana 44 , F. Del Corso 23b,23a ,
J. Del Peso 100 , F. Del Rio 63a , L. Delagrange 128 , F. Deliot 136 , C.M. Delitzsch 49 ,
M. Della Pietra 72a,72b , D. Della Volpe 56 , A. Dell’Acqua 36 , L. Dell’Asta 71a,71b , M. Delmastro 4 ,
P.A. Delsart 60 , S. Demers 173 , M. Demichev 38 , S.P. Denisov 37 , L. D’Eramo 40 ,
D. Derendarz 87 , F. Derue 128 , P. Dervan 93 , K. Desch 24 , C. Deutsch 24 , F.A. Di Bello 57b,57a ,
A. Di Ciaccio 76a,76b , L. Di Ciaccio 4 , A. Di Domenico 75a,75b , C. Di Donato 72a,72b ,
A. Di Girolamo 36 , G. Di Gregorio 36 , A. Di Luca 78a,78b , B. Di Micco 77a,77b , R. Di Nardo 77a,77b ,
M. Diamantopoulou 34 , F.A. Dias 115 , T. Dias Do Vale 143 , M.A. Diaz 138a,138b ,
F.G. Diaz Capriles 24 , M. Didenko 164 , E.B. Diehl 107 , S. Díez Cornell 48 , C. Diez Pardos 142 ,
C. Dimitriadi 162,24 , A. Dimitrievska 20 , J. Dingfelder 24 , I-M. Dinu 27b , S.J. Dittmeier 63b ,
F. Dittus 36 , M. Divisek 134 , F. Djama 103 , T. Djobava 150b , C. Doglioni 102,99 ,
A. Dohnalova 28a , J. Dolejsi 134 , Z. Dolezal 134 , K. Domijan 86a , K.M. Dona 39 ,
M. Donadelli 83c , B. Dong 108 , J. Donini 40 , A. D’Onofrio 72a,72b , M. D’Onofrio 93 ,
J. Dopke 135 , A. Doria 72a , N. Dos Santos Fernandes 131a , P. Dougan 102 , M.T. Dova 91 ,
A.T. Doyle 59 , M.A. Draguet 127 , E. Dreyer 170 , I. Drivas-koulouris 10 , M. Drnevich 118 ,
M. Drozdova 56 , D. Du 62a , T.A. du Pree 115 , F. Dubinin 37 , M. Dubovsky 28a , E. Duchovni 170 ,
G. Duckeck 110 , O.A. Ducu 27b , D. Duda 52 , A. Dudarev 36 , E.R. Duden 26 , M. D’uffizi 102 ,
L. Duflot 66 , M. Dührssen 36 , I. Duminica 27g , A.E. Dumitriu 27b , M. Dunford 63a , S. Dungs 49 ,
K. Dunne 47a,47b , A. Duperrin 103 , H. Duran Yildiz 3a , M. Düren 58 , A. Durglishvili 150b ,
B.L. Dwyer 116 , G.I. Dyckes 17a , M. Dyndal 86a , B.S. Dziedzic 36 , Z.O. Earnshaw 147 ,
G.H. Eberwein 127 , B. Eckerova 28a , S. Eggebrecht 55 , E. Egidio Purcino De Souza 128 ,
L.F. Ehrke 56 , G. Eigen 16 , K. Einsweiler 17a , T. Ekelof 162 , P.A. Ekman 99 , S. El Farkh 35b ,
Y. El Ghazali 35b , H. El Jarrari 36 , A. El Moussaouy 109 , V. Ellajosyula 162 , M. Ellert 162 ,
F. Ellinghaus 172 , N. Ellis 36 , J. Elmsheuser 29 , M. Elsawy 117a , M. Elsing 36 ,
D. Emeliyanov 135 , Y. Enari 154 , I. Ene 17a , S. Epari 13 , P.A. Erland 87 , M. Errenst 172 ,
M. Escalier 66 , C. Escobar 164 , E. Etzion 152 , G. Evans 131a , H. Evans 68 , L.S. Evans 96 ,
A. Ezhilov 37 , S. Ezzarqtouni 35a , F. Fabbri 23b,23a , L. Fabbri 23b,23a , G. Facini 97 ,

73
V. Fadeyev 137 , R.M. Fakhrutdinov 37 , D. Fakoudis 101 , S. Falciano 75a ,
L.F. Falda Ulhoa Coelho 36 , F. Fallavollita 111 , J. Faltova 134 , C. Fan 163 , Y. Fan 14a ,
Y. Fang 14a,14e , M. Fanti 71a,71b , M. Faraj 69a,69b , Z. Farazpay 98 , A. Farbin 8 , A. Farilla 77a ,
T. Farooque 108 , S.M. Farrington 52 , F. Fassi 35e , D. Fassouliotis 9 , M. Faucci Giannelli 76a,76b ,
W.J. Fawcett 32 , L. Fayard 66 , P. Federic 134 , P. Federicova 132 , O.L. Fedin 37,a , M. Feickert 171 ,
L. Feligioni 103 , D.E. Fellers 124 , C. Feng 62b , M. Feng 14b , Z. Feng 115 , M.J. Fenton 160 ,
L. Ferencz 48 , R.A.M. Ferguson 92 , S.I. Fernandez Luengo 138f , P. Fernandez Martinez 13 ,
M.J.V. Fernoux 103 , J. Ferrando 92 , A. Ferrari 162 , P. Ferrari 115,114 , R. Ferrari 73a , D. Ferrere 56 ,
C. Ferretti 107 , F. Fiedler 101 , P. Fiedler 133 , A. Filipčič 94 , E.K. Filmer 1 , F. Filthaut 114 ,
M.C.N. Fiolhais 131a,131c,c , L. Fiorini 164 , W.C. Fisher 108 , T. Fitschen 102 , P.M. Fitzhugh136 ,
I. Fleck 142 , P. Fleischmann 107 , T. Flick 172 , M. Flores 33d,aa , L.R. Flores Castillo 64a ,
L. Flores Sanz De Acedo 36 , F.M. Follega 78a,78b , N. Fomin 16 , J.H. Foo 156 , A. Formica 136 ,
A.C. Forti 102 , E. Fortin 36 , A.W. Fortman 17a , M.G. Foti 17a , L. Fountas 9,i , D. Fournier 66 ,
H. Fox 92 , P. Francavilla 74a,74b , S. Francescato 61 , S. Franchellucci 56 , M. Franchini 23b,23a ,
S. Franchino 63a , D. Francis36 , L. Franco 114 , V. Franco Lima 36 , L. Franconi 48 , M. Franklin 61 ,
G. Frattari 26 , Y.Y. Frid 152 , J. Friend 59 , N. Fritzsche 50 , A. Froch 54 , D. Froidevaux 36 ,
J.A. Frost 127 , Y. Fu 62a , S. Fuenzalida Garrido 138f , M. Fujimoto 103 , K.Y. Fung 64a ,
E. Furtado De Simas Filho 83e , M. Furukawa 154 , J. Fuster 164 , A. Gabrielli 23b,23a ,
A. Gabrielli 156 , P. Gadow 36 , G. Gagliardi 57b,57a , L.G. Gagnon 17a , S. Gaid 161 ,
S. Galantzan 152 , E.J. Gallas 127 , B.J. Gallop 135 , K.K. Gan 120 , S. Ganguly 154 , Y. Gao 52 ,
F.M. Garay Walls 138a,138b , B. Garcia29 , C. García 164 , A. Garcia Alonso 115 ,
A.G. Garcia Caffaro 173 , J.E. García Navarro 164 , M. Garcia-Sciveres 17a , G.L. Gardner 129 ,
R.W. Gardner 39 , N. Garelli 159 , D. Garg 80 , R.B. Garg 144 , J.M. Gargan52 , C.A. Garner156 ,
C.M. Garvey 33a , V.K. Gassmann159 , G. Gaudio 73a , V. Gautam13 , P. Gauzzi 75a,75b ,
I.L. Gavrilenko 37 , A. Gavrilyuk 37 , C. Gay 165 , G. Gaycken 48 , E.N. Gazis 10 , A.A. Geanta 27b ,
C.M. Gee 137 , A. Gekow120 , C. Gemme 57b , M.H. Genest 60 , A.D. Gentry 113 , S. George 96 ,
W.F. George 20 , T. Geralis 46 , P. Gessinger-Befurt 36 , M.E. Geyik 172 , M. Ghani 168 ,
K. Ghorbanian 95 , A. Ghosal 142 , A. Ghosh 160 , A. Ghosh 7 , B. Giacobbe 23b , S. Giagu 75a,75b ,
T. Giani 115 , P. Giannetti 74a , A. Giannini 62a , S.M. Gibson 96 , M. Gignac 137 , D.T. Gil 86b ,
A.K. Gilbert 86a , B.J. Gilbert 41 , D. Gillberg 34 , G. Gilles 115 , L. Ginabat 128 ,
D.M. Gingrich 2,ad , M.P. Giordani 69a,69c , P.F. Giraud 136 , G. Giugliarelli 69a,69c , D. Giugni 71a ,
F. Giuli 36 , I. Gkialas 9,i , L.K. Gladilin 37 , C. Glasman 100 , G.R. Gledhill 124 , G. Glemža 48 ,
M. Glisic124 , I. Gnesi 43b,e , Y. Go 29 , M. Goblirsch-Kolb 36 , B. Gocke 49 , D. Godin109 ,
B. Gokturk 21a , S. Goldfarb 106 , T. Golling 56 , M.G.D. Gololo 33g , D. Golubkov 37 ,
J.P. Gombas 108 , A. Gomes 131a,131b , G. Gomes Da Silva 142 , A.J. Gomez Delegido 164 ,
R. Gonçalo 131a , L. Gonella 20 , A. Gongadze 150c , F. Gonnella 20 , J.L. Gonski 144 ,
R.Y. González Andana 52 , S. González de la Hoz 164 , R. Gonzalez Lopez 93 ,
C. Gonzalez Renteria 17a , M.V. Gonzalez Rodrigues 48 , R. Gonzalez Suarez 162 ,
S. Gonzalez-Sevilla 56 , L. Goossens 36 , B. Gorini 36 , E. Gorini 70a,70b , A. Gorišek 94 ,
T.C. Gosart 129 , A.T. Goshaw 51 , M.I. Gostkin 38 , S. Goswami 122 , C.A. Gottardo 36 ,
S.A. Gotz 110 , M. Gouighri 35b , V. Goumarre 48 , A.G. Goussiou 139 , N. Govender 33c ,
I. Grabowska-Bold 86a , K. Graham 34 , E. Gramstad 126 , S. Grancagnolo 70a,70b , C.M. Grant1,136 ,
P.M. Gravila 27f , F.G. Gravili 70a,70b , H.M. Gray 17a , M. Greco 70a,70b , C. Grefe 24 ,
I.M. Gregor 48 , K.T. Greif 160 , P. Grenier 144 , S.G. Grewe111 , A.A. Grillo 137 , K. Grimm 31 ,
S. Grinstein 13,r , J.-F. Grivaz 66 , E. Gross 170 , J. Grosse-Knetter 55 , J.C. Grundy 127 ,
L. Guan 107 , J.G.R. Guerrero Rojas 164 , G. Guerrieri 69a,69c , F. Guescini 111 , R. Gugel 101 ,
J.A.M. Guhit 107 , A. Guida 18 , E. Guilloton 168 , S. Guindon 36 , F. Guo 14a,14e , J. Guo 62c ,

74
L. Guo 48 , Y. Guo 107 , R. Gupta 130 , S. Gurbuz 24 , S.S. Gurdasani 54 , G. Gustavino 36 ,
M. Guth 56 , P. Gutierrez 121 , L.F. Gutierrez Zagazeta 129 , M. Gutsche 50 , C. Gutschow 97 ,
C. Gwenlan 127 , C.B. Gwilliam 93 , E.S. Haaland 126 , A. Haas 118 , M. Habedank 48 ,
C. Haber 17a , H.K. Hadavand 8 , A. Hadef 50 , S. Hadzic 111 , A.I. Hagan 92 , J.J. Hahn 142 ,
E.H. Haines 97 , M. Haleem 167 , J. Haley 122 , J.J. Hall 140 , G.D. Hallewell 103 , L. Halser 19 ,
K. Hamano 166 , M. Hamer 24 , G.N. Hamity 52 , E.J. Hampshire 96 , J. Han 62b , K. Han 62a ,
L. Han 14c , L. Han 62a , S. Han 17a , Y.F. Han 156 , K. Hanagaki 84 , M. Hance 137 ,
D.A. Hangal 41 , H. Hanif 143 , M.D. Hank 129 , J.B. Hansen 42 , P.H. Hansen 42 , K. Hara 158 ,
D. Harada 56 , T. Harenberg 172 , S. Harkusha 37 , M.L. Harris 104 , Y.T. Harris 127 , J. Harrison 13 ,
N.M. Harrison 120 , P.F. Harrison168 , N.M. Hartman 111 , N.M. Hartmann 110 , R.Z. Hasan 96,135 ,
Y. Hasegawa 141 , S. Hassan 16 , R. Hauser 108 , C.M. Hawkes 20 , R.J. Hawkings 36 ,
Y. Hayashi 154 , S. Hayashida 112 , D. Hayden 108 , C. Hayes 107 , R.L. Hayes 115 , C.P. Hays 127 ,
J.M. Hays 95 , H.S. Hayward 93 , F. He 62a , M. He 14a,14e , Y. He 155 , Y. He 48 , Y. He 97 ,
N.B. Heatley 95 , V. Hedberg 99 , A.L. Heggelund 126 , N.D. Hehir 95,* , C. Heidegger 54 ,
K.K. Heidegger 54 , W.D. Heidorn 81 , J. Heilman 34 , S. Heim 48 , T. Heim 17a , J.G. Heinlein 129 ,
J.J. Heinrich 124 , L. Heinrich 111,ab , J. Hejbal 132 , A. Held 171 , S. Hellesund 16 ,
C.M. Helling 165 , S. Hellman 47a,47b , R.C.W. Henderson92 , L. Henkelmann 32 ,
A.M. Henriques Correia36 , H. Herde 99 , Y. Hernández Jiménez 146 , L.M. Herrmann 24 ,
T. Herrmann 50 , G. Herten 54 , R. Hertenberger 110 , L. Hervas 36 , M.E. Hesping 101 ,
N.P. Hessey 157a , M. Hidaoui 35b , E. Hill 156 , S.J. Hillier 20 , J.R. Hinds 108 , F. Hinterkeuser 24 ,
M. Hirose 125 , S. Hirose 158 , D. Hirschbuehl 172 , T.G. Hitchings 102 , B. Hiti 94 , J. Hobbs 146 ,
R. Hobincu 27e , N. Hod 170 , M.C. Hodgkinson 140 , B.H. Hodkinson 127 , A. Hoecker 36 ,
D.D. Hofer 107 , J. Hofer 48 , T. Holm 24 , M. Holzbock 111 , L.B.A.H. Hommels 32 ,
B.P. Honan 102 , J.J. Hong 68 , J. Hong 62c , T.M. Hong 130 , B.H. Hooberman 163 ,
W.H. Hopkins 6 , M.C. Hoppesch 163 , Y. Horii 112 , S. Hou 149 , A.S. Howard 94 , J. Howarth 59 ,
J. Hoya 6 , M. Hrabovsky 123 , A. Hrynevich 48 , T. Hryn’ova 4 , P.J. Hsu 65 , S.-C. Hsu 139 ,
T. Hsu 66 , M. Hu 17a , Q. Hu 62a , S. Huang 64b , X. Huang 14a,14e , Y. Huang 140 , Y. Huang 101 ,
Y. Huang 14a , Z. Huang 102 , Z. Hubacek 133 , M. Huebner 24 , F. Huegging 24 , T.B. Huffman 127 ,
C.A. Hugli 48 , M. Huhtinen 36 , S.K. Huiberts 16 , R. Hulsken 105 , N. Huseynov 12 , J. Huston 108 ,
J. Huth 61 , R. Hyneman 144 , G. Iacobucci 56 , G. Iakovidis 29 , L. Iconomidou-Fayard 66 ,
J.P. Iddon 36 , P. Iengo 72a,72b , R. Iguchi 154 , T. Iizawa 127 , Y. Ikegami 84 , N. Ilic 156 ,
H. Imam 35a , M. Ince Lezki 56 , T. Ingebretsen Carlson 47a,47b , G. Introzzi 73a,73b , M. Iodice 77a ,
V. Ippolito 75a,75b , R.K. Irwin 93 , M. Ishino 154 , W. Islam 171 , C. Issever 18,48 , S. Istin 21a,ah ,
H. Ito 169 , R. Iuppa 78a,78b , A. Ivina 170 , J.M. Izen 45 , V. Izzo 72a , P. Jacka 132 , P. Jackson 1 ,
C.S. Jagfeld 110 , G. Jain 157a , P. Jain 48 , K. Jakobs 54 , T. Jakoubek 170 , J. Jamieson 59 ,
M. Javurkova 104 , L. Jeanty 124 , J. Jejelava 150a,y , P. Jenni 54,f , C.E. Jessiman 34 , C. Jia62b ,
J. Jia 146 , X. Jia 61 , X. Jia 14a,14e , Z. Jia 14c , C. Jiang 52 , S. Jiggins 48 , J. Jimenez Pena 13 ,
S. Jin 14c , A. Jinaru 27b , O. Jinnouchi 155 , P. Johansson 140 , K.A. Johns 7 , J.W. Johnson 137 ,
D.M. Jones 147 , E. Jones 48 , P. Jones 32 , R.W.L. Jones 92 , T.J. Jones 93 , H.L. Joos 55,36 ,
R. Joshi 120 , J. Jovicevic 15 , X. Ju 17a , J.J. Junggeburth 104 , T. Junkermann 63a ,
A. Juste Rozas 13,r , M.K. Juzek 87 , S. Kabana 138e , A. Kaczmarska 87 , M. Kado 111 ,
H. Kagan 120 , M. Kagan 144 , A. Kahn 129 , C. Kahra 101 , T. Kaji 154 , E. Kajomovitz 151 ,
N. Kakati 170 , I. Kalaitzidou 54 , C.W. Kalderon 29 , N.J. Kang 137 , D. Kar 33g , K. Karava 127 ,
M.J. Kareem 157b , E. Karentzos 54 , O. Karkout 115 , S.N. Karpov 38 , Z.M. Karpova 38 ,
V. Kartvelishvili 92 , A.N. Karyukhin 37 , E. Kasimi 153 , J. Katzy 48 , S. Kaur 34 , K. Kawade 141 ,
M.P. Kawale 121 , C. Kawamoto 88 , T. Kawamoto 62a , E.F. Kay 36 , F.I. Kaya 159 , S. Kazakos 108 ,
V.F. Kazanin 37 , Y. Ke 146 , J.M. Keaveney 33a , R. Keeler 166 , G.V. Kehris 61 , J.S. Keller 34 ,

75
A.S. Kelly97 , J.J. Kempster 147 , P.D. Kennedy 101 , O. Kepka 132 , B.P. Kerridge 135 , S. Kersten 172 ,
B.P. Kerševan 94 , L. Keszeghova 28a , S. Ketabchi Haghighat 156 , R.A. Khan 130 , A. Khanov 122 ,
A.G. Kharlamov 37 , T. Kharlamova 37 , E.E. Khoda 139 , M. Kholodenko 37 , T.J. Khoo 18 ,
G. Khoriauli 167 , J. Khubua 150b , Y.A.R. Khwaira 128 , B. Kibirige33g , D.W. Kim 47a,47b ,
Y.K. Kim 39 , N. Kimura 97 , M.K. Kingston 55 , A. Kirchhoff 55 , C. Kirfel 24 , F. Kirfel 24 ,
J. Kirk 135 , A.E. Kiryunin 111 , C. Kitsaki 10 , O. Kivernyk 24 , M. Klassen 159 , C. Klein 34 ,
L. Klein 167 , M.H. Klein 44 , S.B. Klein 56 , U. Klein 93 , P. Klimek 36 , A. Klimentov 29 ,
T. Klioutchnikova 36 , P. Kluit 115 , S. Kluth 111 , E. Kneringer 79 , T.M. Knight 156 , A. Knue 49 ,
R. Kobayashi 88 , D. Kobylianskii 170 , S.F. Koch 127 , M. Kocian 144 , P. Kodyš 134 ,
D.M. Koeck 124 , P.T. Koenig 24 , T. Koffas 34 , O. Kolay 50 , I. Koletsou 4 , T. Komarek 123 ,
K. Köneke 54 , A.X.Y. Kong 1 , T. Kono 119 , N. Konstantinidis 97 , P. Kontaxakis 56 , B. Konya 99 ,
R. Kopeliansky 41 , S. Koperny 86a , K. Korcyl 87 , K. Kordas 153,d , A. Korn 97 , S. Korn 55 ,
I. Korolkov 13 , N. Korotkova 37 , B. Kortman 115 , O. Kortner 111 , S. Kortner 111 ,
W.H. Kostecka 116 , V.V. Kostyukhin 142 , A. Kotsokechagia 136 , A. Kotwal 51 , A. Koulouris 36 ,
A. Kourkoumeli-Charalampidi 73a,73b , C. Kourkoumelis 9 , E. Kourlitis 111,ab , O. Kovanda 124 ,
R. Kowalewski 166 , W. Kozanecki 136 , A.S. Kozhin 37 , V.A. Kramarenko 37 , G. Kramberger 94 ,
P. Kramer 101 , M.W. Krasny 128 , A. Krasznahorkay 36 , J.W. Kraus 172 , J.A. Kremer 48 ,
T. Kresse 50 , J. Kretzschmar 93 , K. Kreul 18 , P. Krieger 156 , S. Krishnamurthy 104 ,
M. Krivos 134 , K. Krizka 20 , K. Kroeninger 49 , H. Kroha 111 , J. Kroll 132 , J. Kroll 129 ,
K.S. Krowpman 108 , U. Kruchonak 38 , H. Krüger 24 , N. Krumnack81 , M.C. Kruse 51 ,
O. Kuchinskaia 37 , S. Kuday 3a , S. Kuehn 36 , R. Kuesters 54 , T. Kuhl 48 , V. Kukhtin 38 ,
Y. Kulchitsky 37,a , S. Kuleshov 138d,138b , M. Kumar 33g , N. Kumari 48 , P. Kumari 157b ,
A. Kupco 132 , T. Kupfer49 , A. Kupich 37 , O. Kuprash 54 , H. Kurashige 85 , L.L. Kurchaninov 157a ,
O. Kurdysh 66 , Y.A. Kurochkin 37 , A. Kurova 37 , M. Kuze 155 , A.K. Kvam 104 , J. Kvita 123 ,
T. Kwan 105 , N.G. Kyriacou 107 , L.A.O. Laatu 103 , C. Lacasta 164 , F. Lacava 75a,75b ,
H. Lacker 18 , D. Lacour 128 , N.N. Lad 97 , E. Ladygin 38 , A. Lafarge 40 , B. Laforge 128 ,
T. Lagouri 173 , F.Z. Lahbabi 35a , S. Lai 55 , J.E. Lambert 166 , S. Lammers 68 , W. Lampl 7 ,
C. Lampoudis 153,d , G. Lamprinoudis101 , A.N. Lancaster 116 , E. Lançon 29 , U. Landgraf 54 ,
M.P.J. Landon 95 , V.S. Lang 54 , O.K.B. Langrekken 126 , A.J. Lankford 160 , F. Lanni 36 ,
K. Lantzsch 24 , A. Lanza 73a , J.F. Laporte 136 , T. Lari 71a , F. Lasagni Manghi 23b , M. Lassnig 36 ,
V. Latonova 132 , A. Laudrain 101 , A. Laurier 151 , S.D. Lawlor 140 , Z. Lawrence 102 ,
R. Lazaridou168 , M. Lazzaroni 71a,71b , B. Le102 , E.M. Le Boulicaut 51 , L.T. Le Pottier 17a ,
B. Leban 23b,23a , A. Lebedev 81 , M. LeBlanc 102 , F. Ledroit-Guillon 60 , S.C. Lee 149 ,
S. Lee 47a,47b , T.F. Lee 93 , L.L. Leeuw 33c , H.P. Lefebvre 96 , M. Lefebvre 166 , C. Leggett 17a ,
G. Lehmann Miotto 36 , M. Leigh 56 , W.A. Leight 104 , W. Leinonen 114 , A. Leisos 153,q ,
M.A.L. Leite 83c , C.E. Leitgeb 18 , R. Leitner 134 , K.J.C. Leney 44 , T. Lenz 24 , S. Leone 74a ,
C. Leonidopoulos 52 , A. Leopold 145 , C. Leroy 109 , R. Les 108 , C.G. Lester 32 , M. Levchenko 37 ,
J. Levêque 4 , L.J. Levinson 170 , G. Levrini 23b,23a , M.P. Lewicki 87 , C. Lewis 139 , D.J. Lewis 4 ,
A. Li 5 , B. Li 62b , C. Li62a , C-Q. Li 111 , H. Li 62a , H. Li 62b , H. Li 14c , H. Li 14b , H. Li 62b ,
J. Li 62c , K. Li 139 , L. Li 62c , M. Li 14a,14e , S. Li 14a,14e , S. Li 62d,62c , T. Li 5 , X. Li 105 ,
Z. Li 127 , Z. Li 154 , Z. Li 14a,14e , S. Liang14a,14e , Z. Liang 14a , M. Liberatore 136 , B. Liberti 76a ,
K. Lie 64c , J. Lieber Marin 83e , H. Lien 68 , K. Lin 108 , R.E. Lindley 7 , J.H. Lindon 2 ,
E. Lipeles 129 , A. Lipniacka 16 , A. Lister 165 , J.D. Little 4 , B. Liu 14a , B.X. Liu 14d ,
D. Liu 62d,62c , E.H.L. Liu 20 , J.B. Liu 62a , J.K.K. Liu 32 , K. Liu 62d , K. Liu 62d,62c , M. Liu 62a ,
M.Y. Liu 62a , P. Liu 14a , Q. Liu 62d,139,62c , X. Liu 62a , X. Liu 62b , Y. Liu 14d,14e , Y.L. Liu 62b ,
Y.W. Liu 62a , J. Llorente Merino 143 , S.L. Lloyd 95 , E.M. Lobodzinska 48 , P. Loch 7 ,
T. Lohse 18 , K. Lohwasser 140 , E. Loiacono 48 , M. Lokajicek 132,* , J.D. Lomas 20 ,

76
J.D. Long 163 , I. Longarini 160 , R. Longo 163 , I. Lopez Paz 67 , A. Lopez Solis 48 ,
N. Lorenzo Martinez 4 , A.M. Lory 110 , M. Losada 117a , G. Löschcke Centeno 147 , O. Loseva 37 ,
X. Lou 47a,47b , X. Lou 14a,14e , A. Lounis 66 , P.A. Love 92 , G. Lu 14a,14e , M. Lu 66 , S. Lu 129 ,
Y.J. Lu 65 , H.J. Lubatti 139 , C. Luci 75a,75b , F.L. Lucio Alves 14c , F. Luehring 68 , I. Luise 146 ,
O. Lukianchuk 66 , O. Lundberg 145 , B. Lund-Jensen 145 , N.A. Luongo 6 , M.S. Lutz 36 ,
A.B. Lux 25 , D. Lynn 29 , R. Lysak 132 , E. Lytken 99 , V. Lyubushkin 38 , T. Lyubushkina 38 ,
M.M. Lyukova 146 , M.Firdaus M. Soberi 52 , H. Ma 29 , K. Ma62a , L.L. Ma 62b , W. Ma 62a ,
Y. Ma 122 , G. Maccarrone 53 , J.C. MacDonald 101 , P.C. Machado De Abreu Farias 83e ,
R. Madar 40 , T. Madula 97 , J. Maeda 85 , T. Maeno 29 , H. Maguire 140 , V. Maiboroda 136 ,
A. Maio 131a,131b,131d , K. Maj 86a , O. Majersky 48 , S. Majewski 124 , N. Makovec 66 ,
V. Maksimovic 15 , B. Malaescu 128 , Pa. Malecki 87 , V.P. Maleev 37 , F. Malek 60,m , M. Mali 94 ,
D. Malito 96 , U. Mallik 80 , S. Maltezos10 , S. Malyukov38 , J. Mamuzic 13 , G. Mancini 53 ,
M.N. Mancini 26 , G. Manco 73a,73b , J.P. Mandalia 95 , I. Mandić 94 ,
L. Manhaes de Andrade Filho 83a , I.M. Maniatis 170 , J. Manjarres Ramos 90 , D.C. Mankad 170 ,
A. Mann 110 , S. Manzoni 36 , L. Mao 62c , X. Mapekula 33c , A. Marantis 153,q , G. Marchiori 5 ,
M. Marcisovsky 132 , C. Marcon 71a , M. Marinescu 20 , S. Marium 48 , M. Marjanovic 121 ,
A. Markhoos 54 , M. Markovitch 66 , E.J. Marshall 92 , Z. Marshall 17a , S. Marti-Garcia 164 ,
T.A. Martin 135 , V.J. Martin 52 , B. Martin dit Latour 16 , L. Martinelli 75a,75b , M. Martinez 13,r ,
P. Martinez Agullo 164 , V.I. Martinez Outschoorn 104 , P. Martinez Suarez 13 , S. Martin-Haugh 135 ,
G. Martinovicova 134 , V.S. Martoiu 27b , A.C. Martyniuk 97 , A. Marzin 36 , D. Mascione 78a,78b ,
L. Masetti 101 , T. Mashimo 154 , J. Masik 102 , A.L. Maslennikov 37 , P. Massarotti 72a,72b ,
P. Mastrandrea 74a,74b , A. Mastroberardino 43b,43a , T. Masubuchi 154 , T. Mathisen 162 ,
J. Matousek 134 , N. Matsuzawa154 , J. Maurer 27b , A.J. Maury 66 , B. Maček 94 , D.A. Maximov 37 ,
A.E. May 102 , R. Mazini 149 , I. Maznas 116 , M. Mazza 108 , S.M. Mazza 137 , E. Mazzeo 71a,71b ,
C. Mc Ginn 29 , J.P. Mc Gowan 166 , S.P. Mc Kee 107 , C.C. McCracken 165 , E.F. McDonald 106 ,
A.E. McDougall 115 , J.A. Mcfayden 147 , R.P. McGovern 129 , G. Mchedlidze 150b ,
R.P. Mckenzie 33g , T.C. Mclachlan 48 , D.J. Mclaughlin 97 , S.J. McMahon 135 ,
C.M. Mcpartland 93 , R.A. McPherson 166,v , S. Mehlhase 110 , A. Mehta 93 , D. Melini 164 ,
B.R. Mellado Garcia 33g , A.H. Melo 55 , F. Meloni 48 , A.M. Mendes Jacques Da Costa 102 ,
H.Y. Meng 156 , L. Meng 92 , S. Menke 111 , M. Mentink 36 , E. Meoni 43b,43a , G. Mercado 116 ,
S. Merianos 153 , C. Merlassino 69a,69c , L. Merola 72a,72b , C. Meroni 71a,71b , J. Metcalfe 6 ,
A.S. Mete 6 , E. Meuser 101 , C. Meyer 68 , J-P. Meyer 136 , R.P. Middleton 135 , L. Mijović 52 ,
G. Mikenberg 170 , M. Mikestikova 132 , M. Mikuž 94 , H. Mildner 101 , A. Milic 36 ,
D.W. Miller 39 , E.H. Miller 144 , L.S. Miller 34 , A. Milov 170 , D.A. Milstead47a,47b , T. Min14c ,
A.A. Minaenko 37 , I.A. Minashvili 150b , L. Mince 59 , A.I. Mincer 118 , B. Mindur 86a ,
M. Mineev 38 , Y. Mino 88 , L.M. Mir 13 , M. Miralles Lopez 59 , M. Mironova 17a , A. Mishima154 ,
M.C. Missio 114 , A. Mitra 168 , V.A. Mitsou 164 , Y. Mitsumori 112 , O. Miu 156 ,
P.S. Miyagawa 95 , T. Mkrtchyan 63a , M. Mlinarevic 97 , T. Mlinarevic 97 , M. Mlynarikova 36 ,
S. Mobius 19 , P. Mogg 110 , M.H. Mohamed Farook 113 , A.F. Mohammed 14a,14e , S. Mohapatra 41 ,
G. Mokgatitswane 33g , L. Moleri 170 , B. Mondal 142 , S. Mondal 133 , K. Mönig 48 ,
E. Monnier 103 , L. Monsonis Romero164 , J. Montejo Berlingen 13 , M. Montella 120 ,
F. Montereali 77a,77b , F. Monticelli 91 , S. Monzani 69a,69c , N. Morange 66 ,
A.L. Moreira De Carvalho 48 , M. Moreno Llácer 164 , C. Moreno Martinez 56 , P. Morettini 57b ,
S. Morgenstern 36 , M. Morii 61 , M. Morinaga 154 , F. Morodei 75a,75b , L. Morvaj 36 ,
P. Moschovakos 36 , B. Moser 36 , M. Mosidze 150b , T. Moskalets 54 , P. Moskvitina 114 ,
J. Moss 31,j , P. Moszkowicz 86a , A. Moussa 35d , E.J.W. Moyse 104 , O. Mtintsilana 33g ,
S. Muanza 103 , J. Mueller 130 , D. Muenstermann 92 , R. Müller 19 , G.A. Mullier 162 ,

77
A.J. Mullin32 , J.J. Mullin129 , D.P. Mungo 156 , D. Munoz Perez 164 , F.J. Munoz Sanchez 102 ,
M. Murin 102 , W.J. Murray 168,135 , M. Muškinja 94 , C. Mwewa 29 , A.G. Myagkov 37,a ,
A.J. Myers 8 , G. Myers 107 , M. Myska 133 , B.P. Nachman 17a , O. Nackenhorst 49 , K. Nagai 127 ,
K. Nagano 84 , J.L. Nagle 29,af , E. Nagy 103 , A.M. Nairz 36 , Y. Nakahama 84 , K. Nakamura 84 ,
K. Nakkalil 5 , H. Nanjo 125 , E.A. Narayanan 113 , I. Naryshkin 37 , L. Nasella 71a,71b ,
M. Naseri 34 , S. Nasri 117b , C. Nass 24 , G. Navarro 22a , J. Navarro-Gonzalez 164 , R. Nayak 152 ,
A. Nayaz 18 , P.Y. Nechaeva 37 , S. Nechaeva 23b,23a , F. Nechansky 48 , L. Nedic 127 , T.J. Neep 20 ,
A. Negri 73a,73b , M. Negrini 23b , C. Nellist 115 , C. Nelson 105 , K. Nelson 107 , S. Nemecek 132 ,
M. Nessi 36,g , M.S. Neubauer 163 , F. Neuhaus 101 , J. Neundorf 48 , P.R. Newman 20 ,
C.W. Ng 130 , Y.W.Y. Ng 48 , B. Ngair 117a , H.D.N. Nguyen 109 , R.B. Nickerson 127 ,
R. Nicolaidou 136 , J. Nielsen 137 , M. Niemeyer 55 , J. Niermann 55 , N. Nikiforou 36 ,
V. Nikolaenko 37,a , I. Nikolic-Audit 128 , K. Nikolopoulos 20 , P. Nilsson 29 , I. Ninca 48 ,
G. Ninio 152 , A. Nisati 75a , N. Nishu 2 , R. Nisius 111 , J-E. Nitschke 50 , E.K. Nkadimeng 33g ,
T. Nobe 154 , T. Nommensen 148 , M.B. Norfolk 140 , R.R.B. Norisam 97 , B.J. Norman 34 ,
M. Noury 35a , J. Novak 94 , T. Novak 94 , L. Novotny 133 , R. Novotny 113 , L. Nozka 123 ,
K. Ntekas 160 , N.M.J. Nunes De Moura Junior 83b , J. Ocariz 128 , A. Ochi 85 , I. Ochoa 131a ,
S. Oerdek 48,s , J.T. Offermann 39 , A. Ogrodnik 134 , A. Oh 102 , C.C. Ohm 145 , H. Oide 84 ,
R. Oishi 154 , M.L. Ojeda 48 , Y. Okumura 154 , L.F. Oleiro Seabra 131a , S.A. Olivares Pino 138d ,
G. Oliveira Correa 13 , D. Oliveira Damazio 29 , D. Oliveira Goncalves 83a , J.L. Oliver 160 ,
Ö.O. Öncel 54 , A.P. O’Neill 19 , A. Onofre 131a,131e , P.U.E. Onyisi 11 , M.J. Oreglia 39 ,
G.E. Orellana 91 , D. Orestano 77a,77b , N. Orlando 13 , R.S. Orr 156 , V. O’Shea 59 ,
L.M. Osojnak 129 , R. Ospanov 62a , G. Otero y Garzon 30 , H. Otono 89 , P.S. Ott 63a ,
G.J. Ottino 17a , M. Ouchrif 35d , F. Ould-Saada 126 , T. Ovsiannikova 139 , M. Owen 59 ,
R.E. Owen 135 , V.E. Ozcan 21a , F. Ozturk 87 , N. Ozturk 8 , S. Ozturk 82 , H.A. Pacey 127 ,
A. Pacheco Pages 13 , C. Padilla Aranda 13 , G. Padovano 75a,75b , S. Pagan Griso 17a ,
G. Palacino 68 , A. Palazzo 70a,70b , J. Pampel 24 , J. Pan 173 , T. Pan 64a , D.K. Panchal 11 ,
C.E. Pandini 115 , J.G. Panduro Vazquez 135 , H.D. Pandya 1 , H. Pang 14b , P. Pani 48 ,
G. Panizzo 69a,69c , L. Panwar 128 , L. Paolozzi 56 , S. Parajuli 163 , A. Paramonov 6 ,
C. Paraskevopoulos 53 , D. Paredes Hernandez 64b , A. Pareti 73a,73b , K.R. Park 41 , T.H. Park 156 ,
M.A. Parker 32 , F. Parodi 57b,57a , E.W. Parrish 116 , V.A. Parrish 52 , J.A. Parsons 41 ,
U. Parzefall 54 , B. Pascual Dias 109 , L. Pascual Dominguez 100 , E. Pasqualucci 75a ,
S. Passaggio 57b , F. Pastore 96 , P. Patel 87 , U.M. Patel 51 , J.R. Pater 102 , T. Pauly 36 ,
C.I. Pazos 159 , J. Pearkes 144 , M. Pedersen 126 , R. Pedro 131a , S.V. Peleganchuk 37 , O. Penc 36 ,
E.A. Pender 52 , G.D. Penn 173 , K.E. Penski 110 , M. Penzin 37 , B.S. Peralva 83d ,
A.P. Pereira Peixoto 139 , L. Pereira Sanchez 144 , D.V. Perepelitsa 29,af , E. Perez Codina 157a ,
M. Perganti 10 , H. Pernegger 36 , O. Perrin 40 , K. Peters 48 , R.F.Y. Peters 102 , B.A. Petersen 36 ,
T.C. Petersen 42 , E. Petit 103 , V. Petousis 133 , C. Petridou 153,d , T. Petru 134 , A. Petrukhin 142 ,
M. Pettee 17a , N.E. Pettersson 36 , A. Petukhov 37 , K. Petukhova 134 , R. Pezoa 138f ,
L. Pezzotti 36 , G. Pezzullo 173 , T.M. Pham 171 , T. Pham 106 , P.W. Phillips 135 , G. Piacquadio 146 ,
E. Pianori 17a , F. Piazza 124 , R. Piegaia 30 , D. Pietreanu 27b , A.D. Pilkington 102 ,
M. Pinamonti 69a,69c , J.L. Pinfold 2 , B.C. Pinheiro Pereira 131a , A.E. Pinto Pinoargote 136,136 ,
L. Pintucci 69a,69c , K.M. Piper 147 , A. Pirttikoski 56 , D.A. Pizzi 34 , L. Pizzimento 64b ,
A. Pizzini 115 , M.-A. Pleier 29 , V. Pleskot 134 , E. Plotnikova38 , G. Poddar 95 , R. Poettgen 99 ,
L. Poggioli 128 , I. Pokharel 55 , S. Polacek 134 , G. Polesello 73a , A. Poley 143,157a , A. Polini 23b ,
C.S. Pollard 168 , Z.B. Pollock 120 , E. Pompa Pacchi 75a,75b , N.I. Pond 97 , D. Ponomarenko 114 ,
L. Pontecorvo 36 , S. Popa 27a , G.A. Popeneciu 27d , A. Poreba 36 , D.M. Portillo Quintero 157a ,
S. Pospisil 133 , M.A. Postill 140 , P. Postolache 27c , K. Potamianos 168 , P.A. Potepa 86a ,

78
I.N. Potrap 38 , C.J. Potter 32 , H. Potti 1 , J. Poveda 164 , M.E. Pozo Astigarraga 36 ,
A. Prades Ibanez 164 , J. Pretel 54 , D. Price 102 , M. Primavera 70a , M.A. Principe Martin 100 ,
R. Privara 123 , T. Procter 59 , M.L. Proffitt 139 , N. Proklova 129 , K. Prokofiev 64c , G. Proto 111 ,
J. Proudfoot 6 , M. Przybycien 86a , W.W. Przygoda 86b , A. Psallidas 46 , J.E. Puddefoot 140 ,
D. Pudzha 37 , D. Pyatiizbyantseva 37 , J. Qian 107 , D. Qichen 102 , Y. Qin 13 , T. Qiu 52 ,
A. Quadt 55 , M. Queitsch-Maitland 102 , G. Quetant 56 , R.P. Quinn 165 , G. Rabanal Bolanos 61 ,
D. Rafanoharana 54 , F. Raffaeli 76a,76b , F. Ragusa 71a,71b , J.L. Rainbolt 39 , J.A. Raine 56 ,
S. Rajagopalan 29 , E. Ramakoti 37 , I.A. Ramirez-Berend 34 , K. Ran 48,14e , N.P. Rapheeha 33g ,
H. Rasheed 27b , V. Raskina 128 , D.F. Rassloff 63a , A. Rastogi 17a , S. Rave 101 , B. Ravina 55 ,
I. Ravinovich 170 , M. Raymond 36 , A.L. Read 126 , N.P. Readioff 140 , D.M. Rebuzzi 73a,73b ,
G. Redlinger 29 , A.S. Reed 111 , K. Reeves 26 , J.A. Reidelsturz 172 , D. Reikher 152 , A. Rej 49 ,
C. Rembser 36 , M. Renda 27b , M.B. Rendel111 , F. Renner 48 , A.G. Rennie 160 , A.L. Rescia 48 ,
S. Resconi 71a , M. Ressegotti 57b,57a , S. Rettie 36 , J.G. Reyes Rivera 108 , E. Reynolds 17a ,
O.L. Rezanova 37 , P. Reznicek 134 , H. Riani 35d , N. Ribaric 92 , E. Ricci 78a,78b , R. Richter 111 ,
S. Richter 47a,47b , E. Richter-Was 86b , M. Ridel 128 , S. Ridouani 35d , P. Rieck 118 , P. Riedler 36 ,
E.M. Riefel 47a,47b , J.O. Rieger 115 , M. Rijssenbeek 146 , M. Rimoldi 36 , L. Rinaldi 23b,23a ,
T.T. Rinn 29 , M.P. Rinnagel 110 , G. Ripellino 162 , I. Riu 13 , J.C. Rivera Vergara 166 ,
F. Rizatdinova 122 , E. Rizvi 95 , B.R. Roberts 17a , S.H. Robertson 105,v , D. Robinson 32 ,
C.M. Robles Gajardo138f , M. Robles Manzano 101 , A. Robson 59 , A. Rocchi 76a,76b , C. Roda 74a,74b ,
S. Rodriguez Bosca 36 , Y. Rodriguez Garcia 22a , A. Rodriguez Rodriguez 54 ,
A.M. Rodríguez Vera 116 , S. Roe36 , J.T. Roemer 160 , A.R. Roepe-Gier 137 , J. Roggel 172 ,
O. Røhne 126 , R.A. Rojas 104 , C.P.A. Roland 128 , J. Roloff 29 , A. Romaniouk 37 ,
E. Romano 73a,73b , M. Romano 23b , A.C. Romero Hernandez 163 , N. Rompotis 93 , L. Roos 128 ,
S. Rosati 75a , B.J. Rosser 39 , E. Rossi 127 , E. Rossi 72a,72b , L.P. Rossi 61 , L. Rossini 54 ,
R. Rosten 120 , M. Rotaru 27b , B. Rottler 54 , C. Rougier 90 , D. Rousseau 66 , D. Rousso 48 ,
A. Roy 163 , S. Roy-Garand 156 , A. Rozanov 103 , Z.M.A. Rozario 59 , Y. Rozen 151 ,
A. Rubio Jimenez 164 , A.J. Ruby 93 , V.H. Ruelas Rivera 18 , T.A. Ruggeri 1 , A. Ruggiero 127 ,
A. Ruiz-Martinez 164 , A. Rummler 36 , Z. Rurikova 54 , N.A. Rusakovich 38 , H.L. Russell 166 ,
G. Russo 75a,75b , J.P. Rutherfoord 7 , S. Rutherford Colmenares 32 , M. Rybar 134 , E.B. Rye 126 ,
A. Ryzhov 44 , J.A. Sabater Iglesias 56 , P. Sabatini 164 , H.F-W. Sadrozinski 137 ,
F. Safai Tehrani 75a , B. Safarzadeh Samani 135 , S. Saha 1 , M. Sahinsoy 111 , A. Saibel 164 ,
M. Saimpert 136 , M. Saito 154 , T. Saito 154 , A. Sala 71a,71b , D. Salamani 36 , A. Salnikov 144 ,
J. Salt 164 , A. Salvador Salas 152 , D. Salvatore 43b,43a , F. Salvatore 147 , A. Salzburger 36 ,
D. Sammel 54 , E. Sampson 92 , D. Sampsonidis 153,d , D. Sampsonidou 124 , J. Sánchez 164 ,
V. Sanchez Sebastian 164 , H. Sandaker 126 , C.O. Sander 48 , J.A. Sandesara 104 , M. Sandhoff 172 ,
C. Sandoval 22b , L. Sanfilippo 63a , D.P.C. Sankey 135 , T. Sano 88 , A. Sansoni 53 , L. Santi 75a,75b ,
C. Santoni 40 , H. Santos 131a,131b , A. Santra 170 , E. Sanzani 23b,23a , K.A. Saoucha 161 ,
J.G. Saraiva 131a,131d , J. Sardain 7 , O. Sasaki 84 , K. Sato 158 , C. Sauer63b , E. Sauvan 4 ,
P. Savard 156,ad , R. Sawada 154 , C. Sawyer 135 , L. Sawyer 98 , C. Sbarra 23b , A. Sbrizzi 23b,23a ,
T. Scanlon 97 , J. Schaarschmidt 139 , U. Schäfer 101 , A.C. Schaffer 66,44 , D. Schaile 110 ,
R.D. Schamberger 146 , C. Scharf 18 , M.M. Schefer 19 , V.A. Schegelsky 37 , D. Scheirich 134 ,
M. Schernau 160 , C. Scheulen 55 , C. Schiavi 57b,57a , M. Schioppa 43b,43a , B. Schlag 144,l ,
K.E. Schleicher 54 , S. Schlenker 36 , J. Schmeing 172 , M.A. Schmidt 172 , K. Schmieden 101 ,
C. Schmitt 101 , N. Schmitt 101 , S. Schmitt 48 , L. Schoeffel 136 , A. Schoening 63b ,
P.G. Scholer 34 , E. Schopf 127 , M. Schott 24 , J. Schovancova 36 , S. Schramm 56 , T. Schroer 56 ,
H-C. Schultz-Coulon 63a , M. Schumacher 54 , B.A. Schumm 137 , Ph. Schune 136 , A.J. Schuy 139 ,
H.R. Schwartz 137 , A. Schwartzman 144 , T.A. Schwarz 107 , Ph. Schwemling 136 ,

79
R. Schwienhorst 108 , A. Sciandra 29 , G. Sciolla 26 , F. Scuri 74a , C.D. Sebastiani 93 ,
K. Sedlaczek 116 , S.C. Seidel 113 , A. Seiden 137 , B.D. Seidlitz 41 , C. Seitz 48 , J.M. Seixas 83b ,
G. Sekhniaidze 72a , L. Selem 60 , N. Semprini-Cesari 23b,23a , D. Sengupta 56 , V. Senthilkumar 164 ,
L. Serin 66 , M. Sessa 76a,76b , H. Severini 121 , F. Sforza 57b,57a , A. Sfyrla 56 , Q. Sha 14a ,
E. Shabalina 55 , A.H. Shah 32 , R. Shaheen 145 , J.D. Shahinian 129 , D. Shaked Renous 170 ,
L.Y. Shan 14a , M. Shapiro 17a , A. Sharma 36 , A.S. Sharma 165 , P. Sharma 80 , P.B. Shatalov 37 ,
K. Shaw 147 , S.M. Shaw 102 , Q. Shen 62c,5 , D.J. Sheppard 143 , P. Sherwood 97 , L. Shi 97 ,
X. Shi 14a , C.O. Shimmin 173 , J.D. Shinner 96 , I.P.J. Shipsey 127 , S. Shirabe 89 ,
M. Shiyakova 38,t , M.J. Shochet 39 , J. Shojaii 106 , D.R. Shope 126 , B. Shrestha 121 ,
S. Shrestha 120,ag , M.J. Shroff 166 , P. Sicho 132 , A.M. Sickles 163 , E. Sideras Haddad 33g ,
A.C. Sidley 115 , A. Sidoti 23b , F. Siegert 50 , Dj. Sijacki 15 , F. Sili 91 , J.M. Silva 52 ,
M.V. Silva Oliveira 29 , S.B. Silverstein 47a , S. Simion66 , R. Simoniello 36 , E.L. Simpson 102 ,
H. Simpson 147 , L.R. Simpson 107 , N.D. Simpson99 , S. Simsek 82 , S. Sindhu 55 , P. Sinervo 156 ,
S. Singh 156 , S. Sinha 48 , S. Sinha 102 , M. Sioli 23b,23a , I. Siral 36 , E. Sitnikova 48 ,
J. Sjölin 47a,47b , A. Skaf 55 , E. Skorda 20 , P. Skubic 121 , M. Slawinska 87 , V. Smakhtin170 ,
B.H. Smart 135 , S.Yu. Smirnov 37 , Y. Smirnov 37 , L.N. Smirnova 37,a , O. Smirnova 99 ,
A.C. Smith 41 , D.R. Smith160 , E.A. Smith 39 , H.A. Smith 127 , J.L. Smith 102 , R. Smith144 ,
M. Smizanska 92 , K. Smolek 133 , A.A. Snesarev 37 , S.R. Snider 156 , H.L. Snoek 115 ,
S. Snyder 29 , R. Sobie 166,v , A. Soffer 152 , C.A. Solans Sanchez 36 , E.Yu. Soldatov 37 ,
U. Soldevila 164 , A.A. Solodkov 37 , S. Solomon 26 , A. Soloshenko 38 , K. Solovieva 54 ,
O.V. Solovyanov 40 , P. Sommer 36 , A. Sonay 13 , W.Y. Song 157b , A. Sopczak 133 , A.L. Sopio 97 ,
F. Sopkova 28b , J.D. Sorenson 113 , I.R. Sotarriva Alvarez 155 , V. Sothilingam63a ,
O.J. Soto Sandoval 138c,138b , S. Sottocornola 68 , R. Soualah 161 , Z. Soumaimi 35e , D. South 48 ,
N. Soybelman 170 , S. Spagnolo 70a,70b , M. Spalla 111 , D. Sperlich 54 , G. Spigo 36 , S. Spinali 92 ,
D.P. Spiteri 59 , M. Spousta 134 , E.J. Staats 34 , R. Stamen 63a , A. Stampekis 20 , M. Standke 24 ,
E. Stanecka 87 , W. Stanek-Maslouska 48 , M.V. Stange 50 , B. Stanislaus 17a , M.M. Stanitzki 48 ,
B. Stapf 48 , E.A. Starchenko 37 , G.H. Stark 137 , J. Stark 90 , P. Staroba 132 , P. Starovoitov 63a ,
S. Stärz 105 , R. Staszewski 87 , G. Stavropoulos 46 , J. Steentoft 162 , P. Steinberg 29 ,
B. Stelzer 143,157a , H.J. Stelzer 130 , O. Stelzer-Chilton 157a , H. Stenzel 58 , T.J. Stevenson 147 ,
G.A. Stewart 36 , J.R. Stewart 122 , M.C. Stockton 36 , G. Stoicea 27b , M. Stolarski 131a ,
S. Stonjek 111 , A. Straessner 50 , J. Strandberg 145 , S. Strandberg 47a,47b , M. Stratmann 172 ,
M. Strauss 121 , T. Strebler 103 , P. Strizenec 28b , R. Ströhmer 167 , D.M. Strom 124 ,
R. Stroynowski 44 , A. Strubig 47a,47b , S.A. Stucci 29 , B. Stugu 16 , J. Stupak 121 , N.A. Styles 48 ,
D. Su 144 , S. Su 62a , W. Su 62d , X. Su 62a , D. Suchy 28a , K. Sugizaki 154 , V.V. Sulin 37 ,
M.J. Sullivan 93 , D.M.S. Sultan 127 , L. Sultanaliyeva 37 , S. Sultansoy 3b , T. Sumida 88 ,
S. Sun 107 , S. Sun 171 , O. Sunneborn Gudnadottir 162 , N. Sur 103 , M.R. Sutton 147 ,
H. Suzuki 158 , M. Svatos 132 , M. Swiatlowski 157a , T. Swirski 167 , I. Sykora 28a , M. Sykora 134 ,
T. Sykora 134 , D. Ta 101 , K. Tackmann 48,s , A. Taffard 160 , R. Tafirout 157a ,
J.S. Tafoya Vargas 66 , Y. Takubo 84 , M. Talby 103 , A.A. Talyshev 37 , K.C. Tam 64b ,
N.M. Tamir152 , A. Tanaka 154 , J. Tanaka 154 , R. Tanaka 66 , M. Tanasini 146 , Z. Tao 165 ,
S. Tapia Araya 138f , S. Tapprogge 101 , A. Tarek Abouelfadl Mohamed 108 , S. Tarem 151 ,
K. Tariq 14a , G. Tarna 27b , G.F. Tartarelli 71a , M.J. Tartarin 90 , P. Tas 134 , M. Tasevsky 132 ,
E. Tassi 43b,43a , A.C. Tate 163 , G. Tateno 154 , Y. Tayalati 35e,u , G.N. Taylor 106 , W. Taylor 157b ,
A.S. Tee 171 , R. Teixeira De Lima 144 , P. Teixeira-Dias 96 , J.J. Teoh 156 , K. Terashi 154 ,
J. Terron 100 , S. Terzo 13 , M. Testa 53 , R.J. Teuscher 156,v , A. Thaler 79 , O. Theiner 56 ,
N. Themistokleous 52 , T. Theveneaux-Pelzer 103 , O. Thielmann 172 , D.W. Thomas96 ,
J.P. Thomas 20 , E.A. Thompson 17a , P.D. Thompson 20 , E. Thomson 129 , R.E. Thornberry 44 ,

80
C. Tian 62a , Y. Tian 55 , V. Tikhomirov 37,a , Yu.A. Tikhonov 37 , S. Timoshenko37 ,
D. Timoshyn 134 , E.X.L. Ting 1 , P. Tipton 173 , A. Tishelman-Charny 29 , S.H. Tlou 33g ,
K. Todome 155 , S. Todorova-Nova 134 , S. Todt50 , L. Toffolin 69a,69c , M. Togawa 84 , J. Tojo 89 ,
S. Tokár 28a , K. Tokushuku 84 , O. Toldaiev 68 , R. Tombs 32 , M. Tomoto 84,112 ,
L. Tompkins 144,l , K.W. Topolnicki 86b , E. Torrence 124 , H. Torres 90 , E. Torró Pastor 164 ,
M. Toscani 30 , C. Tosciri 39 , M. Tost 11 , D.R. Tovey 140 , A. Traeet16 , I.S. Trandafir 27b ,
T. Trefzger 167 , A. Tricoli 29 , I.M. Trigger 157a , S. Trincaz-Duvoid 128 , D.A. Trischuk 26 ,
B. Trocmé 60 , L. Truong 33c , M. Trzebinski 87 , A. Trzupek 87 , F. Tsai 146 , M. Tsai 107 ,
A. Tsiamis 153,d , P.V. Tsiareshka37 , S. Tsigaridas 157a , A. Tsirigotis 153,q , V. Tsiskaridze 156 ,
E.G. Tskhadadze 150a , M. Tsopoulou 153 , Y. Tsujikawa 88 , I.I. Tsukerman 37 , V. Tsulaia 17a ,
S. Tsuno 84 , K. Tsuri 119 , D. Tsybychev 146 , Y. Tu 64b , A. Tudorache 27b , V. Tudorache 27b ,
A.N. Tuna 61 , S. Turchikhin 57b,57a , I. Turk Cakir 3a , R. Turra 71a , T. Turtuvshin 38,w ,
P.M. Tuts 41 , S. Tzamarias 153,d , E. Tzovara 101 , F. Ukegawa 158 , P.A. Ulloa Poblete 138c,138b ,
E.N. Umaka 29 , G. Unal 36 , A. Undrus 29 , G. Unel 160 , J. Urban 28b , P. Urrejola 138a ,
G. Usai 8 , R. Ushioda 155 , M. Usman 109 , Z. Uysal 82 , V. Vacek 133 , B. Vachon 105 ,
T. Vafeiadis 36 , A. Vaitkus 97 , C. Valderanis 110 , E. Valdes Santurio 47a,47b , M. Valente 157a ,
S. Valentinetti 23b,23a , A. Valero 164 , E. Valiente Moreno 164 , A. Vallier 90 , J.A. Valls Ferrer 164 ,
D.R. Van Arneman 115 , T.R. Van Daalen 139 , A. Van Der Graaf 49 , P. Van Gemmeren 6 ,
M. Van Rijnbach 36 , S. Van Stroud 97 , I. Van Vulpen 115 , P. Vana 134 , M. Vanadia 76a,76b ,
W. Vandelli 36 , E.R. Vandewall 122 , D. Vannicola 152 , L. Vannoli 53 , R. Vari 75a , E.W. Varnes 7 ,
C. Varni 17b , T. Varol 149 , D. Varouchas 66 , L. Varriale 164 , K.E. Varvell 148 , M.E. Vasile 27b ,
L. Vaslin84 , G.A. Vasquez 166 , A. Vasyukov 38 , R. Vavricka101 , T. Vazquez Schroeder 36 ,
J. Veatch 31 , V. Vecchio 102 , M.J. Veen 104 , I. Veliscek 29 , L.M. Veloce 156 , F. Veloso 131a,131c ,
S. Veneziano 75a , A. Ventura 70a,70b , S. Ventura Gonzalez 136 , A. Verbytskyi 111 ,
M. Verducci 74a,74b , C. Vergis 95 , M. Verissimo De Araujo 83b , W. Verkerke 115 ,
J.C. Vermeulen 115 , C. Vernieri 144 , M. Vessella 104 , M.C. Vetterli 143,ad , A. Vgenopoulos 153,d ,
N. Viaux Maira 138f , T. Vickey 140 , O.E. Vickey Boeriu 140 , G.H.A. Viehhauser 127 , L. Vigani 63b ,
M. Villa 23b,23a , M. Villaplana Perez 164 , E.M. Villhauer52 , E. Vilucchi 53 , M.G. Vincter 34 ,
A. Visibile115 , C. Vittori 36 , I. Vivarelli 23b,23a , E. Voevodina 111 , F. Vogel 110 , J.C. Voigt 50 ,
P. Vokac 133 , Yu. Volkotrub 86b , J. Von Ahnen 48 , E. Von Toerne 24 , B. Vormwald 36 ,
V. Vorobel 134 , K. Vorobev 37 , M. Vos 164 , K. Voss 142 , M. Vozak 115 , L. Vozdecky 121 ,
N. Vranjes 15 , M. Vranjes Milosavljevic 15 , M. Vreeswijk 115 , N.K. Vu 62d,62c , R. Vuillermet 36 ,
O. Vujinovic 101 , I. Vukotic 39 , S. Wada 158 , C. Wagner104 , J.M. Wagner 17a , W. Wagner 172 ,
S. Wahdan 172 , H. Wahlberg 91 , M. Wakida 112 , J. Walder 135 , R. Walker 110 , W. Walkowiak 142 ,
A. Wall 129 , E.J. Wallin 99 , T. Wamorkar 6 , A.Z. Wang 137 , C. Wang 101 , C. Wang 11 ,
H. Wang 17a , J. Wang 64c , R. Wang 61 , R. Wang 6 , S.M. Wang 149 , S. Wang 62b , S. Wang 14a ,
T. Wang 62a , W.T. Wang 80 , W. Wang 14a , X. Wang 14c , X. Wang 163 , X. Wang 62c ,
Y. Wang 62d , Y. Wang 14c , Z. Wang 107 , Z. Wang 62d,51,62c , Z. Wang 107 , A. Warburton 105 ,
R.J. Ward 20 , N. Warrack 59 , S. Waterhouse 96 , A.T. Watson 20 , H. Watson 59 , M.F. Watson 20 ,
E. Watton 59,135 , G. Watts 139 , B.M. Waugh 97 , J.M. Webb 54 , C. Weber 29 , H.A. Weber 18 ,
M.S. Weber 19 , S.M. Weber 63a , C. Wei 62a , Y. Wei 54 , A.R. Weidberg 127 , E.J. Weik 118 ,
J. Weingarten 49 , C. Weiser 54 , C.J. Wells 48 , T. Wenaus 29 , B. Wendland 49 , T. Wengler 36 ,
N.S. Wenke111 , N. Wermes 24 , M. Wessels 63a , A.M. Wharton 92 , A.S. White 61 , A. White 8 ,
M.J. White 1 , D. Whiteson 160 , L. Wickremasinghe 125 , W. Wiedenmann 171 , M. Wielers 135 ,
C. Wiglesworth 42 , D.J. Wilbern121 , H.G. Wilkens 36 , J.J.H. Wilkinson 32 , D.M. Williams 41 ,
H.H. Williams129 , S. Williams 32 , S. Willocq 104 , B.J. Wilson 102 , P.J. Windischhofer 39 ,
F.I. Winkel 30 , F. Winklmeier 124 , B.T. Winter 54 , J.K. Winter 102 , M. Wittgen144 , M. Wobisch 98 ,

81
T. Wojtkowski60 , Z. Wolffs 115 , J. Wollrath160 , M.W. Wolter 87 , H. Wolters 131a,131c , M.C. Wong137 ,
E.L. Woodward 41 , S.D. Worm 48 , B.K. Wosiek 87 , K.W. Woźniak 87 , S. Wozniewski 55 ,
K. Wraight 59 , C. Wu 20 , M. Wu 14d , M. Wu 114 , S.L. Wu 171 , X. Wu 56 , Y. Wu 62a , Z. Wu 4 ,
J. Wuerzinger 111,ab , T.R. Wyatt 102 , B.M. Wynne 52 , S. Xella 42 , L. Xia 14c , M. Xia 14b ,
J. Xiang 64c , M. Xie 62a , S. Xin 14a,14e , A. Xiong 124 , J. Xiong 17a , D. Xu 14a , H. Xu 62a ,
L. Xu 62a , R. Xu 129 , T. Xu 107 , Y. Xu 14b , Z. Xu 52 , Z. Xu14c , B. Yabsley 148 , S. Yacoob 33a ,
Y. Yamaguchi 155 , E. Yamashita 154 , H. Yamauchi 158 , T. Yamazaki 17a , Y. Yamazaki 85 ,
J. Yan62c , S. Yan 59 , Z. Yan 104 , H.J. Yang 62c,62d , H.T. Yang 62a , S. Yang 62a , T. Yang 64c ,
X. Yang 36 , X. Yang 14a , Y. Yang 44 , Y. Yang62a , Z. Yang 62a , W-M. Yao 17a , H. Ye 14c ,
H. Ye 55 , J. Ye 14a , S. Ye 29 , X. Ye 62a , Y. Yeh 97 , I. Yeletskikh 38 , B.K. Yeo 17b ,
M.R. Yexley 97 , T.P. Yildirim 127 , P. Yin 41 , K. Yorita 169 , S. Younas 27b , C.J.S. Young 36 ,
C. Young 144 , C. Yu 14a,14e , Y. Yu 62a , M. Yuan 107 , R. Yuan 62d,62c , L. Yue 97 ,
M. Zaazoua 62a , B. Zabinski 87 , E. Zaid52 , Z.K. Zak 87 , T. Zakareishvili 164 , N. Zakharchuk 34 ,
S. Zambito 56 , J.A. Zamora Saa 138d,138b , J. Zang 154 , D. Zanzi 54 , O. Zaplatilek 133 ,
C. Zeitnitz 172 , H. Zeng 14a , J.C. Zeng 163 , D.T. Zenger Jr 26 , O. Zenin 37 , T. Ženiš 28a ,
S. Zenz 95 , S. Zerradi 35a , D. Zerwas 66 , M. Zhai 14a,14e , D.F. Zhang 140 , J. Zhang 62b ,
J. Zhang 6 , K. Zhang 14a,14e , L. Zhang 62a , L. Zhang 14c , P. Zhang 14a,14e , R. Zhang 171 ,
S. Zhang 107 , S. Zhang 90 , T. Zhang 154 , X. Zhang 62c , X. Zhang 62b , Y. Zhang 62c ,
Y. Zhang 97 , Y. Zhang 14c , Z. Zhang 17a , Z. Zhang 62b , Z. Zhang 66 , H. Zhao 139 , T. Zhao 62b ,
Y. Zhao 137 , Z. Zhao 62a , Z. Zhao 62a , A. Zhemchugov 38 , J. Zheng 14c , K. Zheng 163 ,
X. Zheng 62a , Z. Zheng 144 , D. Zhong 163 , B. Zhou 107 , H. Zhou 7 , N. Zhou 62c , Y. Zhou14b ,
Y. Zhou 14c , Y. Zhou7 , C.G. Zhu 62b , J. Zhu 107 , X. Zhu62d , Y. Zhu 62c , Y. Zhu 62a ,
X. Zhuang 14a , K. Zhukov 37 , N.I. Zimine 38 , J. Zinsser 63b , M. Ziolkowski 142 , L. Živković 15 ,
A. Zoccoli 23b,23a , K. Zoch 61 , T.G. Zorbas 140 , O. Zormpa 46 , W. Zou 41 , L. Zwalinski 36 .
1
Department of Physics, University of Adelaide, Adelaide; Australia.
2
Department of Physics, University of Alberta, Edmonton AB; Canada.
3 (𝑎)
Department of Physics, Ankara University, Ankara; (𝑏) Division of Physics, TOBB University of
Economics and Technology, Ankara; Türkiye.
4
LAPP, Université Savoie Mont Blanc, CNRS/IN2P3, Annecy; France.
5
APC, Université Paris Cité, CNRS/IN2P3, Paris; France.
6
High Energy Physics Division, Argonne National Laboratory, Argonne IL; United States of America.
7
Department of Physics, University of Arizona, Tucson AZ; United States of America.
8
Department of Physics, University of Texas at Arlington, Arlington TX; United States of America.
9
Physics Department, National and Kapodistrian University of Athens, Athens; Greece.
10
Physics Department, National Technical University of Athens, Zografou; Greece.
11
Department of Physics, University of Texas at Austin, Austin TX; United States of America.
12
Institute of Physics, Azerbaijan Academy of Sciences, Baku; Azerbaijan.
13
Institut de Física d’Altes Energies (IFAE), Barcelona Institute of Science and Technology, Barcelona;
Spain.
14 (𝑎)
Institute of High Energy Physics, Chinese Academy of Sciences, Beijing; (𝑏) Physics Department,
Tsinghua University, Beijing; (𝑐) Department of Physics, Nanjing University, Nanjing; (𝑑) School of Science,
Shenzhen Campus of Sun Yat-sen University; (𝑒) University of Chinese Academy of Science (UCAS),
Beijing; China.
15
Institute of Physics, University of Belgrade, Belgrade; Serbia.
16
Department for Physics and Technology, University of Bergen, Bergen; Norway.
17 (𝑎)
Physics Division, Lawrence Berkeley National Laboratory, Berkeley CA; (𝑏) University of California,

82
Berkeley CA; United States of America.
18
Institut für Physik, Humboldt Universität zu Berlin, Berlin; Germany.
19
Albert Einstein Center for Fundamental Physics and Laboratory for High Energy Physics, University of
Bern, Bern; Switzerland.
20
School of Physics and Astronomy, University of Birmingham, Birmingham; United Kingdom.
21 (𝑎)
Department of Physics, Bogazici University, Istanbul; (𝑏) Department of Physics Engineering,
Gaziantep University, Gaziantep; (𝑐) Department of Physics, Istanbul University, Istanbul; Türkiye.
22 (𝑎)
Facultad de Ciencias y Centro de Investigaciónes, Universidad Antonio Nariño,
Bogotá; (𝑏) Departamento de Física, Universidad Nacional de Colombia, Bogotá; Colombia.
23 (𝑎)
Dipartimento di Fisica e Astronomia A. Righi, Università di Bologna, Bologna; (𝑏) INFN Sezione di
Bologna; Italy.
24
Physikalisches Institut, Universität Bonn, Bonn; Germany.
25
Department of Physics, Boston University, Boston MA; United States of America.
26
Department of Physics, Brandeis University, Waltham MA; United States of America.
27 (𝑎)
Transilvania University of Brasov, Brasov; (𝑏) Horia Hulubei National Institute of Physics and Nuclear
Engineering, Bucharest; (𝑐) Department of Physics, Alexandru Ioan Cuza University of Iasi, Iasi; (𝑑) National
Institute for Research and Development of Isotopic and Molecular Technologies, Physics Department,
Cluj-Napoca; (𝑒) National University of Science and Technology Politechnica, Bucharest; ( 𝑓 ) West
University in Timisoara, Timisoara; (𝑔) Faculty of Physics, University of Bucharest, Bucharest; Romania.
28 (𝑎)
Faculty of Mathematics, Physics and Informatics, Comenius University, Bratislava; (𝑏) Department of
Subnuclear Physics, Institute of Experimental Physics of the Slovak Academy of Sciences, Kosice; Slovak
Republic.
29
Physics Department, Brookhaven National Laboratory, Upton NY; United States of America.
30
Universidad de Buenos Aires, Facultad de Ciencias Exactas y Naturales, Departamento de Física, y
CONICET, Instituto de Física de Buenos Aires (IFIBA), Buenos Aires; Argentina.
31
California State University, CA; United States of America.
32
Cavendish Laboratory, University of Cambridge, Cambridge; United Kingdom.
33 (𝑎)
Department of Physics, University of Cape Town, Cape Town; (𝑏) iThemba Labs, Western
Cape; (𝑐) Department of Mechanical Engineering Science, University of Johannesburg,
Johannesburg; (𝑑) National Institute of Physics, University of the Philippines Diliman
(Philippines); (𝑒) University of South Africa, Department of Physics, Pretoria; ( 𝑓 ) University of Zululand,
KwaDlangezwa; (𝑔) School of Physics, University of the Witwatersrand, Johannesburg; South Africa.
34
Department of Physics, Carleton University, Ottawa ON; Canada.
35 (𝑎)
Faculté des Sciences Ain Chock, Réseau Universitaire de Physique des Hautes Energies - Université
Hassan II, Casablanca; (𝑏) Faculté des Sciences, Université Ibn-Tofail, Kénitra; (𝑐) Faculté des Sciences
Semlalia, Université Cadi Ayyad, LPHEA-Marrakech; (𝑑) LPMR, Faculté des Sciences, Université
Mohamed Premier, Oujda; (𝑒) Faculté des sciences, Université Mohammed V, Rabat; ( 𝑓 ) Institute of Applied
Physics, Mohammed VI Polytechnic University, Ben Guerir; Morocco.
36
CERN, Geneva; Switzerland.
37
Affiliated with an institute covered by a cooperation agreement with CERN.
38
Affiliated with an international laboratory covered by a cooperation agreement with CERN.
39
Enrico Fermi Institute, University of Chicago, Chicago IL; United States of America.
40
LPC, Université Clermont Auvergne, CNRS/IN2P3, Clermont-Ferrand; France.
41
Nevis Laboratory, Columbia University, Irvington NY; United States of America.
42
Niels Bohr Institute, University of Copenhagen, Copenhagen; Denmark.
43 (𝑎)
Dipartimento di Fisica, Università della Calabria, Rende; (𝑏) INFN Gruppo Collegato di Cosenza,
Laboratori Nazionali di Frascati; Italy.

83
44
Physics Department, Southern Methodist University, Dallas TX; United States of America.
45
Physics Department, University of Texas at Dallas, Richardson TX; United States of America.
46
National Centre for Scientific Research "Demokritos", Agia Paraskevi; Greece.
47 (𝑎)
Department of Physics, Stockholm University; (𝑏) Oskar Klein Centre, Stockholm; Sweden.
48
Deutsches Elektronen-Synchrotron DESY, Hamburg and Zeuthen; Germany.
49
Fakultät Physik , Technische Universität Dortmund, Dortmund; Germany.
50
Institut für Kern- und Teilchenphysik, Technische Universität Dresden, Dresden; Germany.
51
Department of Physics, Duke University, Durham NC; United States of America.
52
SUPA - School of Physics and Astronomy, University of Edinburgh, Edinburgh; United Kingdom.
53
INFN e Laboratori Nazionali di Frascati, Frascati; Italy.
54
Physikalisches Institut, Albert-Ludwigs-Universität Freiburg, Freiburg; Germany.
55
II. Physikalisches Institut, Georg-August-Universität Göttingen, Göttingen; Germany.
56
Département de Physique Nucléaire et Corpusculaire, Université de Genève, Genève; Switzerland.
57 (𝑎)
Dipartimento di Fisica, Università di Genova, Genova; (𝑏) INFN Sezione di Genova; Italy.
58
II. Physikalisches Institut, Justus-Liebig-Universität Giessen, Giessen; Germany.
59
SUPA - School of Physics and Astronomy, University of Glasgow, Glasgow; United Kingdom.
60
LPSC, Université Grenoble Alpes, CNRS/IN2P3, Grenoble INP, Grenoble; France.
61
Laboratory for Particle Physics and Cosmology, Harvard University, Cambridge MA; United States of
America.
62 (𝑎)
Department of Modern Physics and State Key Laboratory of Particle Detection and Electronics,
University of Science and Technology of China, Hefei; (𝑏) Institute of Frontier and Interdisciplinary
Science and Key Laboratory of Particle Physics and Particle Irradiation (MOE), Shandong University,
Qingdao; (𝑐) School of Physics and Astronomy, Shanghai Jiao Tong University, Key Laboratory for Particle
Astrophysics and Cosmology (MOE), SKLPPC, Shanghai; (𝑑) Tsung-Dao Lee Institute, Shanghai; (𝑒) School
of Physics and Microelectronics, Zhengzhou University; China.
63 (𝑎)
Kirchhoff-Institut für Physik, Ruprecht-Karls-Universität Heidelberg, Heidelberg; (𝑏) Physikalisches
Institut, Ruprecht-Karls-Universität Heidelberg, Heidelberg; Germany.
64 (𝑎)
Department of Physics, Chinese University of Hong Kong, Shatin, N.T., Hong Kong; (𝑏) Department
of Physics, University of Hong Kong, Hong Kong; (𝑐) Department of Physics and Institute for Advanced
Study, Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong; China.
65
Department of Physics, National Tsing Hua University, Hsinchu; Taiwan.
66
IJCLab, Université Paris-Saclay, CNRS/IN2P3, 91405, Orsay; France.
67
Centro Nacional de Microelectrónica (IMB-CNM-CSIC), Barcelona; Spain.
68
Department of Physics, Indiana University, Bloomington IN; United States of America.
69 (𝑎)
INFN Gruppo Collegato di Udine, Sezione di Trieste, Udine; (𝑏) ICTP, Trieste; (𝑐) Dipartimento
Politecnico di Ingegneria e Architettura, Università di Udine, Udine; Italy.
70 (𝑎)
INFN Sezione di Lecce; (𝑏) Dipartimento di Matematica e Fisica, Università del Salento, Lecce; Italy.
71 (𝑎)
INFN Sezione di Milano; (𝑏) Dipartimento di Fisica, Università di Milano, Milano; Italy.
72 (𝑎)
INFN Sezione di Napoli; (𝑏) Dipartimento di Fisica, Università di Napoli, Napoli; Italy.
73 (𝑎)
INFN Sezione di Pavia; (𝑏) Dipartimento di Fisica, Università di Pavia, Pavia; Italy.
74 (𝑎)
INFN Sezione di Pisa; (𝑏) Dipartimento di Fisica E. Fermi, Università di Pisa, Pisa; Italy.
75 (𝑎)
INFN Sezione di Roma; (𝑏) Dipartimento di Fisica, Sapienza Università di Roma, Roma; Italy.
76 (𝑎)
INFN Sezione di Roma Tor Vergata; (𝑏) Dipartimento di Fisica, Università di Roma Tor Vergata,
Roma; Italy.
77 (𝑎)
INFN Sezione di Roma Tre; (𝑏) Dipartimento di Matematica e Fisica, Università Roma Tre, Roma;
Italy.
78 (𝑎)
INFN-TIFPA; (𝑏) Università degli Studi di Trento, Trento; Italy.

84
79
Universität Innsbruck, Department of Astro and Particle Physics, Innsbruck; Austria.
80
University of Iowa, Iowa City IA; United States of America.
81
Department of Physics and Astronomy, Iowa State University, Ames IA; United States of America.
82
Istinye University, Sariyer, Istanbul; Türkiye.
83 (𝑎)
Departamento de Engenharia Elétrica, Universidade Federal de Juiz de Fora (UFJF), Juiz de
Fora; (𝑏) Universidade Federal do Rio De Janeiro COPPE/EE/IF, Rio de Janeiro; (𝑐) Instituto de Física,
Universidade de São Paulo, São Paulo; (𝑑) Rio de Janeiro State University, Rio de Janeiro; (𝑒) Federal
University of Bahia, Bahia; Brazil.
84
KEK, High Energy Accelerator Research Organization, Tsukuba; Japan.
85
Graduate School of Science, Kobe University, Kobe; Japan.
86 (𝑎)
AGH University of Krakow, Faculty of Physics and Applied Computer Science, Krakow; (𝑏) Marian
Smoluchowski Institute of Physics, Jagiellonian University, Krakow; Poland.
87
Institute of Nuclear Physics Polish Academy of Sciences, Krakow; Poland.
88
Faculty of Science, Kyoto University, Kyoto; Japan.
89
Research Center for Advanced Particle Physics and Department of Physics, Kyushu University, Fukuoka ;
Japan.
90
L2IT, Université de Toulouse, CNRS/IN2P3, UPS, Toulouse; France.
91
Instituto de Física La Plata, Universidad Nacional de La Plata and CONICET, La Plata; Argentina.
92
Physics Department, Lancaster University, Lancaster; United Kingdom.
93
Oliver Lodge Laboratory, University of Liverpool, Liverpool; United Kingdom.
94
Department of Experimental Particle Physics, Jožef Stefan Institute and Department of Physics,
University of Ljubljana, Ljubljana; Slovenia.
95
School of Physics and Astronomy, Queen Mary University of London, London; United Kingdom.
96
Department of Physics, Royal Holloway University of London, Egham; United Kingdom.
97
Department of Physics and Astronomy, University College London, London; United Kingdom.
98
Louisiana Tech University, Ruston LA; United States of America.
99
Fysiska institutionen, Lunds universitet, Lund; Sweden.
100
Departamento de Física Teorica C-15 and CIAFF, Universidad Autónoma de Madrid, Madrid; Spain.
101
Institut für Physik, Universität Mainz, Mainz; Germany.
102
School of Physics and Astronomy, University of Manchester, Manchester; United Kingdom.
103
CPPM, Aix-Marseille Université, CNRS/IN2P3, Marseille; France.
104
Department of Physics, University of Massachusetts, Amherst MA; United States of America.
105
Department of Physics, McGill University, Montreal QC; Canada.
106
School of Physics, University of Melbourne, Victoria; Australia.
107
Department of Physics, University of Michigan, Ann Arbor MI; United States of America.
108
Department of Physics and Astronomy, Michigan State University, East Lansing MI; United States of
America.
109
Group of Particle Physics, University of Montreal, Montreal QC; Canada.
110
Fakultät für Physik, Ludwig-Maximilians-Universität München, München; Germany.
111
Max-Planck-Institut für Physik (Werner-Heisenberg-Institut), München; Germany.
112
Graduate School of Science and Kobayashi-Maskawa Institute, Nagoya University, Nagoya; Japan.
113
Department of Physics and Astronomy, University of New Mexico, Albuquerque NM; United States of
America.
114
Institute for Mathematics, Astrophysics and Particle Physics, Radboud University/Nikhef, Nijmegen;
Netherlands.
115
Nikhef National Institute for Subatomic Physics and University of Amsterdam, Amsterdam;
Netherlands.

85
116
Department of Physics, Northern Illinois University, DeKalb IL; United States of America.
117 (𝑎)
New York University Abu Dhabi, Abu Dhabi; (𝑏) United Arab Emirates University, Al Ain; United
Arab Emirates.
118
Department of Physics, New York University, New York NY; United States of America.
119
Ochanomizu University, Otsuka, Bunkyo-ku, Tokyo; Japan.
120
Ohio State University, Columbus OH; United States of America.
121
Homer L. Dodge Department of Physics and Astronomy, University of Oklahoma, Norman OK; United
States of America.
122
Department of Physics, Oklahoma State University, Stillwater OK; United States of America.
123
Palacký University, Joint Laboratory of Optics, Olomouc; Czech Republic.
124
Institute for Fundamental Science, University of Oregon, Eugene, OR; United States of America.
125
Graduate School of Science, Osaka University, Osaka; Japan.
126
Department of Physics, University of Oslo, Oslo; Norway.
127
Department of Physics, Oxford University, Oxford; United Kingdom.
128
LPNHE, Sorbonne Université, Université Paris Cité, CNRS/IN2P3, Paris; France.
129
Department of Physics, University of Pennsylvania, Philadelphia PA; United States of America.
130
Department of Physics and Astronomy, University of Pittsburgh, Pittsburgh PA; United States of
America.
131 (𝑎)
Laboratório de Instrumentação e Física Experimental de Partículas - LIP, Lisboa; (𝑏) Departamento
de Física, Faculdade de Ciências, Universidade de Lisboa, Lisboa; (𝑐) Departamento de Física,
Universidade de Coimbra, Coimbra; (𝑑) Centro de Física Nuclear da Universidade de Lisboa,
Lisboa; (𝑒) Departamento de Física, Universidade do Minho, Braga; ( 𝑓 ) Departamento de Física Teórica y
del Cosmos, Universidad de Granada, Granada (Spain); (𝑔) Departamento de Física, Instituto Superior
Técnico, Universidade de Lisboa, Lisboa; Portugal.
132
Institute of Physics of the Czech Academy of Sciences, Prague; Czech Republic.
133
Czech Technical University in Prague, Prague; Czech Republic.
134
Charles University, Faculty of Mathematics and Physics, Prague; Czech Republic.
135
Particle Physics Department, Rutherford Appleton Laboratory, Didcot; United Kingdom.
136
IRFU, CEA, Université Paris-Saclay, Gif-sur-Yvette; France.
137
Santa Cruz Institute for Particle Physics, University of California Santa Cruz, Santa Cruz CA; United
States of America.
138 (𝑎)
Departamento de Física, Pontificia Universidad Católica de Chile, Santiago; (𝑏) Millennium Institute
for Subatomic physics at high energy frontier (SAPHIR), Santiago; (𝑐) Instituto de Investigación
Multidisciplinario en Ciencia y Tecnología, y Departamento de Física, Universidad de La
Serena; (𝑑) Universidad Andres Bello, Department of Physics, Santiago; (𝑒) Instituto de Alta Investigación,
Universidad de Tarapacá, Arica; ( 𝑓 ) Departamento de Física, Universidad Técnica Federico Santa María,
Valparaíso; Chile.
139
Department of Physics, University of Washington, Seattle WA; United States of America.
140
Department of Physics and Astronomy, University of Sheffield, Sheffield; United Kingdom.
141
Department of Physics, Shinshu University, Nagano; Japan.
142
Department Physik, Universität Siegen, Siegen; Germany.
143
Department of Physics, Simon Fraser University, Burnaby BC; Canada.
144
SLAC National Accelerator Laboratory, Stanford CA; United States of America.
145
Department of Physics, Royal Institute of Technology, Stockholm; Sweden.
146
Departments of Physics and Astronomy, Stony Brook University, Stony Brook NY; United States of
America.
147
Department of Physics and Astronomy, University of Sussex, Brighton; United Kingdom.

86
148
School of Physics, University of Sydney, Sydney; Australia.
149
Institute of Physics, Academia Sinica, Taipei; Taiwan.
150 (𝑎)
E. Andronikashvili Institute of Physics, Iv. Javakhishvili Tbilisi State University, Tbilisi; (𝑏) High
Energy Physics Institute, Tbilisi State University, Tbilisi; (𝑐) University of Georgia, Tbilisi; Georgia.
151
Department of Physics, Technion, Israel Institute of Technology, Haifa; Israel.
152
Raymond and Beverly Sackler School of Physics and Astronomy, Tel Aviv University, Tel Aviv; Israel.
153
Department of Physics, Aristotle University of Thessaloniki, Thessaloniki; Greece.
154
International Center for Elementary Particle Physics and Department of Physics, University of Tokyo,
Tokyo; Japan.
155
Department of Physics, Tokyo Institute of Technology, Tokyo; Japan.
156
Department of Physics, University of Toronto, Toronto ON; Canada.
157 (𝑎)
TRIUMF, Vancouver BC; (𝑏) Department of Physics and Astronomy, York University, Toronto ON;
Canada.
158
Division of Physics and Tomonaga Center for the History of the Universe, Faculty of Pure and Applied
Sciences, University of Tsukuba, Tsukuba; Japan.
159
Department of Physics and Astronomy, Tufts University, Medford MA; United States of America.
160
Department of Physics and Astronomy, University of California Irvine, Irvine CA; United States of
America.
161
University of Sharjah, Sharjah; United Arab Emirates.
162
Department of Physics and Astronomy, University of Uppsala, Uppsala; Sweden.
163
Department of Physics, University of Illinois, Urbana IL; United States of America.
164
Instituto de Física Corpuscular (IFIC), Centro Mixto Universidad de Valencia - CSIC, Valencia; Spain.
165
Department of Physics, University of British Columbia, Vancouver BC; Canada.
166
Department of Physics and Astronomy, University of Victoria, Victoria BC; Canada.
167
Fakultät für Physik und Astronomie, Julius-Maximilians-Universität Würzburg, Würzburg; Germany.
168
Department of Physics, University of Warwick, Coventry; United Kingdom.
169
Waseda University, Tokyo; Japan.
170
Department of Particle Physics and Astrophysics, Weizmann Institute of Science, Rehovot; Israel.
171
Department of Physics, University of Wisconsin, Madison WI; United States of America.
172
Fakultät für Mathematik und Naturwissenschaften, Fachgruppe Physik, Bergische Universität
Wuppertal, Wuppertal; Germany.
173
Department of Physics, Yale University, New Haven CT; United States of America.
𝑎
Also Affiliated with an institute covered by a cooperation agreement with CERN.
𝑏
Also at An-Najah National University, Nablus; Palestine.
𝑐
Also at Borough of Manhattan Community College, City University of New York, New York NY; United
States of America.
𝑑
Also at Center for Interdisciplinary Research and Innovation (CIRI-AUTH), Thessaloniki; Greece.
𝑒
Also at Centro Studi e Ricerche Enrico Fermi; Italy.
𝑓
Also at CERN, Geneva; Switzerland.
𝑔
Also at Département de Physique Nucléaire et Corpusculaire, Université de Genève, Genève;
Switzerland.

Also at Departament de Fisica de la Universitat Autonoma de Barcelona, Barcelona; Spain.
𝑖
Also at Department of Financial and Management Engineering, University of the Aegean, Chios; Greece.
𝑗
Also at Department of Physics, California State University, Sacramento; United States of America.
𝑘
Also at Department of Physics, King’s College London, London; United Kingdom.
𝑙
Also at Department of Physics, Stanford University, Stanford CA; United States of America.
𝑚
Also at Department of Physics, Stellenbosch University; South Africa.

87
𝑛
Also at Department of Physics, University of Fribourg, Fribourg; Switzerland.
𝑜
Also at Department of Physics, University of Thessaly; Greece.
𝑝
Also at Department of Physics, Westmont College, Santa Barbara; United States of America.
𝑞
Also at Hellenic Open University, Patras; Greece.
𝑟
Also at Institucio Catalana de Recerca i Estudis Avancats, ICREA, Barcelona; Spain.
𝑠
Also at Institut für Experimentalphysik, Universität Hamburg, Hamburg; Germany.
𝑡
Also at Institute for Nuclear Research and Nuclear Energy (INRNE) of the Bulgarian Academy of
Sciences, Sofia; Bulgaria.
𝑢
Also at Institute of Applied Physics, Mohammed VI Polytechnic University, Ben Guerir; Morocco.
𝑣
Also at Institute of Particle Physics (IPP); Canada.
𝑤
Also at Institute of Physics and Technology, Mongolian Academy of Sciences, Ulaanbaatar; Mongolia.
𝑥
Also at Institute of Physics, Azerbaijan Academy of Sciences, Baku; Azerbaijan.
𝑦
Also at Institute of Theoretical Physics, Ilia State University, Tbilisi; Georgia.
𝑧
Also at Lawrence Livermore National Laboratory, Livermore; United States of America.
𝑎𝑎
Also at National Institute of Physics, University of the Philippines Diliman (Philippines); Philippines.
𝑎𝑏
Also at Technical University of Munich, Munich; Germany.
𝑎𝑐
Also at The Collaborative Innovation Center of Quantum Matter (CICQM), Beijing; China.
𝑎𝑑
Also at TRIUMF, Vancouver BC; Canada.
𝑎𝑒
Also at Università di Napoli Parthenope, Napoli; Italy.
𝑎𝑓
Also at University of Colorado Boulder, Department of Physics, Colorado; United States of America.
𝑎𝑔
Also at Washington College, Chestertown, MD; United States of America.
𝑎ℎ
Also at Yeditepe University, Physics Department, Istanbul; Türkiye.

Deceased

88

You might also like