arXiv:1505.07137v1 [physics.atom

New Exclusion Limits for the Search of Scalar and Pseudoscalar Axion-Like Particles
from “ Light Shining Through a Wall ”
R. Ballou,1, 2 G. Deferne,3 M. Finger Jr.,4 M. Finger,4 L. Flekova,5 J. Hosek,5 S. Kunc,6 K. Macuchova,5
K. A. Meissner,7 P. Pugnat,8, 2, ∗ M. Schott,9 A. Siemko,3 M. Slunecka,4 M. Sulc,6 C. Weinsheimer,9 and J. Zicha5
(OSQAR Collaboration)
1
CNRS, Institut N´eel, 38042 Grenoble, France
Universit´e Grenoble Alpes, 38042 Grenoble, France
3
CERN, CH-1211 Geneva-23, Switzerland
4
Charles University, Faculty of Mathematics and Physics, Prague, Czech Republic
5
Czech Technical University, Prague, Czech Republic
6
Technical University of Liberec, 46117 Liberec, Czech Republic
7
University of Warsaw, Institute of Theoretical Physics, 00-681 Warsaw, Poland
8
CNRS, LNCMI, F-38042 Grenoble, France
9
University of Mainz, Institute of Physics, 55128 Mainz, Germany
(Dated: June 29, 2015)
arXiv:1506.08082v1 [hep-ex] 26 Jun 2015
2
Physics beyond the Standard Model predicts the possible existence of new particles that can
be searched at the low energy frontier in the sub-eV range. The OSQAR photon regeneration
experiment looks for “Light Shining through a Wall” from the quantum oscillation of optical photons
into “Weakly Interacting Sub-eV Particles”, such as axion or Axion-Like Particles (ALPs), in a 9 T
transverse magnetic field over the unprecedented length of 2 × 14.3 m. In 2014, this experiment has
been run with an outstanding sensitivity, using an 18.5 W continuous wave laser emitting in the
green spectral regime (532 nm). No regenerated photons have been detected after the wall, pushing
the limits for the existence of axions and ALPs down to an unprecedented level for such a type
of laboratory experiment. The di-photon couplings of possible pseudo-scalar and scalar ALPs can
be constrained in the nearly massless limit to be less than 3.5 · 10−8 GeV−1 and 3.2 · 10−8 GeV−1 ,
respectively.
PACS numbers: 14.80.Va, 14.80.-j, 95.35.+d, 14.70.Bh
Keywords: Axion, Scalar and Pseudoscalar Axion-like Particles, Photon Regeneration
Possible extensions of the Standard Model (SM) of particle physics are not restricted to the high energy frontier. Therefore, researchers are increasingly interested
in the search of Weakly Interacting Sub-eV Particles
(WISPs). These are, as the names reveals, particles displaying much weaker interactions and masses below the
eV range. One emblematic example is the axion, a pseudoscalar boson arising from the spontaneous breaking of
a global chiral symmetry U (1)A , postulated to dynamically solve the strong CP problem [1–3]. Axions and
Axion-Like Particles (ALPs) are predicted in supersymmetric theories [4], in string theory [5, 6], and in the
Conformal Standard Model [7]. ALPs can be scalar as
well as pseudoscalar and theorized to couple to the SM
through a variety of mechanisms, giving rise in particular to a two-photons vertex. WISPs also include light
bosons of gauge groups under which the SM particles are
not charged (hidden sectors). These may interact with
the SM through gravity, kinetic mixing or higher order
quantum processes [8]. The interest aroused by WISPs
goes beyond particle physics; similarly to the earlier hypothesis for the axion [9], they provide alternative candidates for dark matter (DM) [10–12]. Moreover, they
might explain a number of astrophysical puzzles, such as
the universe transparency to very high energy photons
(> 100 GeV) [13], the anomalous white dwarf cooling [14]
or the recently discovered gamma ray excesses in galaxy
clusters [15]. In contrast to Weakly Interacting Massive
Particles (WIMPs), which can be searched for at TeV
colliders such as the Large Hadron Collider (LHC) at
CERN, the detection of WISPs requires to have recourse
to dedicated low-energy experiments. Several methodologies exploiting the existence of a di-photon coupling
have been proposed based on lasers, microwave cavities,
strong electromagnetic fields or torsion balances [16, 17].
The OSQAR (Optical Search for QED Vacuum Bifringence, Axions and Photon Regeneration) experiment at
CERN, is at the forefront of this low-energy frontier of
particle/astroparticle physics. It combines the simultaneous use of high magnetic fields with laser beams in
distinct experiments. One of its setups uses the “Light
Shining Through a Wall” (LSW) method for the search
of the ALPs [18]. A pioneering work in this line excluded
ALPs with a di-photon coupling constant gAγγ larger
than 6.7 · 10−7 GeV−1 for masses below 10−3 eV [19].
These exclusion limits were later extended by other LSW
experiments to gAγγ > 6.5 · 10−8 GeV−1 for masses below 5 × 10−4 eV [20] and more recently tightened to
gAγγ > 5.7·10−8 GeV−1 for masses below 2×10−4 eV [21].
This Letter reports novel results, obtained from the 2014
data-taking campaign of the OSQAR LSW experiment.
Compared to the previous experimental setups of OS-
Few-photon imaging at 1550 nm using a lowtiming-jitter superconducting nanowire singlephoton detector
Hui Zhou, Yuhao He, Lixing You, * Sijin Chen, Weijun Zhang,
Junjie Wu, Zhen Wang, and Xiaoming Xie
State Key Laboratory of Functional Materials for Informatics, Shanghai Institute of Microsystem and Information
Technology, Chinese Academy of Sciences, 865 Changning Road, Shanghai 200050, China
*[email protected]
Abstract: We demonstrated a laser depth imaging system based on the
time-correlated single-photon counting technique, which was incorporated
with a low-jitter superconducting nanowire single-photon detector (SNSPD),
operated at the wavelength of 1550 nm. A sub-picosecond time-bin width
was chosen for photon counting, resulting in a discrete noise of less than
one/two counts for each time bin under indoor/outdoor daylight conditions,
with a collection time of 50 ms. Because of the low-jitter SNSPD, the target
signal histogram was significantly distinguishable, even for a fairly low
retro-reflected photon flux. The depth information was determined directly
by the highest bin counts, instead of using any data fitting combined with
complex algorithms. Millimeter resolution depth imaging of a low-signature
object was obtained, and more accurate data than that produced by the
traditional Gaussian fitting method was generated. Combined with the
intensity of the return photons, three-dimensional reconstruction overlaid
with reflectivity data was realized.
References and links
1.
2.
3.
4.
5.
6.
7.
8.
9.
B. Schwarz, "Mapping the world in 3D," Nat. Photonics 4 (7), 429-430 (2010).
R. E. Warburton, A. McCarthy, A. M. Wallace, S. Hernandez-Marin, R. H. Hadfield, S. W. Nam, and G. S.
Buller, "Subcentimeter depth resolution using a single-photon counting time-of-flight laser ranging system
at 1550 nm wavelength," Opt. Lett. 32 (15), 2266-2268 (2007).
G. S. Buller, R. D. Harkins, A. McCarthy, P. A. Hiskett, G. R. MacKinnon, G. R. Smith, R. Sung, A. M.
Wallace, R. A. Lamb, K. D. Ridley, and J. G. Rarity, "Multiple wavelength time-of-flight sensor based on
time-correlated single-photon counting," Rev. Sci. Instrum. 76 (8) (2005).
S. Pellegrini, G. S. Buller, J. M. Smith, A. M. Wallace, and S. Cova, "Laser-based distance measurement
using picosecond resolution time-correlated single-photon counting," Measurement Science & Technology
11 (6), 712-716 (2000).
S. Chen, D. Liu, W. Zhang, L. You, Y. He, W. Zhang, X. Yang, G. Wu, M. Ren, H. Zeng, Z. Wang, X. Xie,
and M. Jiang, "Time-of-flight laser ranging and imaging at 1550 nm using low-jitter superconducting
nanowire single-photon detection system," Appl. Opt. 52 (14), 3241-3245 (2013).
A. McCarthy, X. Ren, A. Della Frera, N. R. Gemmell, N. J. Krichel, C. Scarcella, A. Ruggeri, A. Tosi, and
G. S. Buller, "Kilometer-range depth imaging at 1550 nm wavelength using an InGaAs/InP single-photon
avalanche diode detector," Opt. Express 21 (19), 22098-22113 (2013).
A. McCarthy, N. J. Krichel, N. R. Gemmell, X. Ren, M. G. Tanner, S. N. Dorenbos, V. Zwiller, R. H.
Hadfield, and G. S. Buller, "Kilometer-range, high resolution depth imaging via 1560 nm wavelength
single-photon detection," Opt. Express 21 (7), 8904-8915 (2013).
L. Sjöqvist, M. Henriksson, P. Jonsson, and O. Steinvall, "Time-correlated single-photon counting range
profiling and reflectance tomographic imaging," Adv. Opt. Techn. 3 (2), 187-197 (2014).
N. J. Krichel, A. McCarthy, I. Rech, M. Ghioni, A. Gulinatti, and G. S. Buller, "Cumulative data
acquisition in comparative photon-counting three-dimensional imaging," J. Mod. Opt. 58 (3-4), 244-256
(2011).
Statistical analysis of the temporal single-photon response of
superconducting nanowire single photon detection
He Yu-Hao(何宇昊), Lv Chao-Lin(吕超林), Zhang Wei-Jun(张伟君), Zhang Lu(张露), Wu Jun-Jie(巫
君杰), Chen Si-Jing(陈思井), You Li-Xing(尤立星) †, Wang Zhen(王镇)
State Key Laboratory of Functional Materials for Informatics, Shanghai Institute of Microsystem and
Information Technology, Chinese Academy of Sciences, Shanghai 200050, China
Abstract
Counting rate is a key parameter of superconducting nanowire single photon
detectors (SNSPD) and is determined by the current recovery time of an SNSPD after
a detection event. We propose a new method to study the transient detection efficiency
(DE) and pulse amplitude during the current recovery process by statistically analyzing
the single photon response of an SNSPD under photon illumination with a high
repetition rate. The transient DE results match well with the DEs deduced from the
static current dependence of DE combined with the waveform of a single-photon
detection event. This proves that the static measurement results can be used to analyze
the transient current recovery process after a detection event. The results are relevant
for understanding the current recovery process of SNSPDs after a detection event and
for determining the counting rate of SNSPDs.
Keywords: single photon detector, recovery process, kinetic inductance
1. Introduction
The superconducting nanowire single-photon detector (SNSPD) is a promising
*Project supported by Strategic Priority Research Program (B) of the Chinese Academy of Sciences (XDB04010200)
and National Basic Research Program of China (2011CBA00202) and the National Natural Science Foundation of
China (61401441).
†Corresponding author. E-mail: [email protected]
Particle detection through the quantum counter concept in YAG:Er3+
A. F. Borghesani,1 C. Braggio,2, a) G. Carugno,2 F. Chiossi,2 A. Di Lieto,3 M. Guarise,2 G. Ruoso,4 and M.
Tonelli3
1)
arXiv:1506.07987v1 [physics.ins-det] 26 Jun 2015
CNISM Unit, Dip. di Fisica e Astronomia and INFN, Via F. Marzolo 8, I-35131 Padova,
Italy
2)
Dip. di Fisica e Astronomia and INFN, Via F. Marzolo 8, I-35131 Padova,
Italy
3)
Dip. di Fisica and INFN, Largo Bruno Pontecorvo, 3, I-56127 Pisa, Italy
4)
INFN, Laboratori Nazionali di Legnaro, Viale dell’Universit`
a 2, I-35020 Legnaro,
Italy
We report about a novel scheme for particle detection based on the infrared quantum counter concept. Its
operation consists of a two-step excitation process of a four level system, that can be realized in rare earthdoped crystals when a cw pump laser is tuned to the transition from the second to the fourth level. The
incident particle raises the atoms of the active material into a low lying, metastable energy state, triggering
the absorption of the pump laser to a higher level. Following a rapid non-radiative decay to a fluorescent
level, an optical signal is observed with a conventional detectors. In order to demonstrate the feasibility
of such a scheme, we have investigated the emission from the fluorescent level 4 S3/2 (540 nm band) in an
Er3+ -doped YAG crystal pumped by a tunable titanium sapphire laser when it is irradiated with 60 keV
electrons delivered by an electron gun. We have obtained a clear signature this excitation increases the 4 I13/2
metastable level population that can efficiently be exploited to generate a detectable optical signal.
There is a significant interest in the development of
devices for the detection of low rate, low energy deposition events both for dark matter searches1 and the study
of neutrino interactions in condensed matter2,3 . A very
low value of energy threshold (∼ 0.5 keV) has been reported in semiconductor detectors4 , or in bolometers,
where a few tens of eV events could be detected5 . Unfortunately, their active mass, a crucial parameter in rare
events searches, are also tiniest: in the most sensitive
detector only a few grams of material act as a target.
In the present work an all-optical detection scheme
is proposed to efficiently convert the incident particle
energy into detectable photons. It is based on the infrared quantum counter concept (IRQC), proposed by
N. Bloembergen6 as a way to extend photon detection to
the 1 − 100 µm wavelength range. The incident infrared
photon is upconverted in a material that exhibits a four
energy level system with E2 > E3 > E1 > E0 , as those
determined in wide bandgap materials doped with trivalent rare-earth (RE) ions7 , and kept under the action
of a pump laser source resonant with transition 1 → 2.
In analogy with the infrared quantum counter, detection
of the particle is then accomplished through the fluorescence photons emitted in the transition 3 → 1 as shown
in Fig. 1.
In contrast to narrow-band, selective detection of infrared photons, a particle that interacts in an optical material gives rise to several phenomena, including energy
transfer processes from the host to the RE ion, which
can be viewed as wideband excitation for the present
purposes. The complex chain of events whereby a particle loses its energy in a material has been systematically investigated for the development of state-of-the-art
a) Electronic
mail: [email protected]
2P
3/2
30000
[
E (cm-1)
2
3
PUMP
15000
DETECTION
SIGNAL
PARTICLE
20000
1
0
10000
5000
0
]
4F 4F
4F5/2 3/2
7/2
4S 4H
3/2 11/2
4F
9/2
4I
9/2
4I
11/2
4I
13/2
4I
15/2
FIG. 1. (left) Ideal four-level active material. (right) Energy
level scheme in YAG:Er8 .
scintillators9–12 . Attention has been paid to the transitions in the visible, UV and near infrared in those studies.
Here we want to focus on the fraction of the particle energy that is translated in the excitation of the low energy
metastable level 1 indicated in Fig. 1, that can take place
both through the decay of higher levels and directly from
the ground state. This is motivated by the assumption
that the particle energy loss is a process in which it is
much more probable to increase the population of low
energy atomic levels than highest ones. Such reasoning
is supported by three points:
1. the inelastic scattering of free electrons off bounded
electrons is a major process in particle energy loss
and is described by the Bethe-Bloch formula that
privileges low energy transfer events13 ;
2. another dominant process is the thermalization of
the secondary electrons produced in the interaction
that takes place through optical phonon scattering;
3. infrared scintillation in the range 600 − 900 nm has
been reported12 with an intensity in excess of 105
MAN/HEP/2015/11
Mass Bounds on Light and Heavy Neutrinos from Radiative MFV Leptogenesis
Apostolos Pilaftsis and Daniele Teresi
Consortium for Fundamental Physics, School of Physics and Astronomy,
University of Manchester, Manchester M13 9PL, United Kingdom.
(Dated: June 29, 2015)
arXiv:1506.08124v1 [hep-ph] 26 Jun 2015
We derive novel limits on the masses of the light and heavy Majorana neutrinos by requiring
successful leptogenesis in seesaw models of minimal flavour violation (MFV). Taking properly into
account radiative flavour effects and avoiding the limitations due to a no-go theorem on leptonic
asymmetries, we find that the mass of the lightest of the observable neutrinos must be smaller than
∼ 0.05 eV, whilst the Majorana scale of lepton number violation should be higher than ∼ 1012 GeV.
The latter lower bound enables one to probe the existence of possible new scales of MFV, up to
energies of ∼ 100 TeV, in low-energy experiments, such as µ → eγ and µ → e conversion in nuclei.
Possible realizations of MFV leptogenesis in Grand Unified Theories are briefly discussed.
PACS numbers: 14.60.St, 11.30.Hv, 14.60.Pq, 98.80.Cq
The hypothesis of Minimal Flavour Violation
(MFV) [1] provides an elegant framework, even for
flavour theories of new physics, to naturally implement
the strong constraints from the non-observation of
sizeable flavour changing neutral currents in the quark
sector. Interestingly enough, this MFV hypothesis can
be extended to the lepton sector as well. In particular,
one may consider seesaw models, in which all effects
of Lepton Flavour Violation (LFV), including the
observable light-neutrino masses and mixing, originate
exclusively from the Yukawa interactions to three
right-handed neutrinos NR,α in a flavour basis with
diagonal charged-lepton Yukawa couplings. Instead, the
right-handed Majorana mass matrix MN takes on its
maximally symmetric form: MN = mN 13 , which means
that MN is invariant under O(3)NR rotations [2, 3].
Such seesaw models allow for large hierarchies between
the seesaw scale mN of Lepton Number Violation
(LNV) and possible other scales of LFV, e.g. due to soft
breaking of supersymmetry [4], giving rise to observable
rates for µ → eγ, µ → eee and µ → e conversion in
nuclei [5].
In addition, the presence of heavy Majorana neutrinos
provides, via the well-established mechanism of leptogenesis [6], one of the most plausible explanations for the
origin of the Baryon Asymmetry in the Universe (BAU).
In the MFV seesaw scenario, the exact mass degeneracy of the three right-handed neutrinos leads to vanishing leptonic asymmetries [7]. However, Renormalization
Group (RG) effects due to running from the Grand Unified Theory (GUT) scale µX ≈ 2 × 1016 GeV to a lower
heavy-neutrino scale mN lift the O(3)NR symmetry in the
heavy-neutrino sector, thus potentially offering the possibility to explain the BAU via the so-called mechanism
of Resonant Leptogenesis (RL) [7, 8].
In this radiative framework [9–11] of MFV, it was originally argued [10] that successful leptogenesis requires the
LNV seesaw scale mN to be higher than 1012 GeV, implying sizeable neutrino Yukawa couplings larger than
∼ 0.1, so as to fit the low-energy neutrino data. The latter would allow one to test the possible existence of additional new-physics scales mediating minimal LFV, for
a wide range of energies from 1 to 100 TeV [10], in lowenergy experiments, such as the upgraded MEG experiment [12]. Subsequent elaborate studies [11], however,
which include flavour effects, have suggested that mN
could be as low as 106 GeV, implying neutrino Yukawa
couplings that could be as small as 10−4 . Thus, these results seem to eliminate any prospects of definitely excluding the presence of new MFV scales in the near future.
Most recently, however, a no-go theorem [13] for vanishing leptonic asymmetries to order h4 in the neutrino
Yukawa couplings has been proved, rendering radiative
MFV leptogenesis not viable in the low-mN region.
In this Letter we will show how this no-go theorem
can be circumvented in a MFV framework, provided the
seesaw scale mN > 1012 GeV. The latter reopens the
way of falsifying possible new leptonic MFV scales up to
energies of ∼ 100 TeV. A complete treatment of flavour
effects yields novel limits on the masses of the light neutrinos as well. In particular, we find that the mass of the
lightest neutrino must be smaller than ∼ 0.05 eV, with
the normal hierarchical light-neutrino spectrum favoured
over the inverted one.
In the MFV seesaw scenario, the flavour-symmetric
part of the Lagrangian Lsym at the GUT scale µX reads
Lsym = LSM
sym +
1 C
N [MN ]αβ NR,β + H.c. ,
2 R,α
(1)
where [MN ]αβ = mN δ αβ and LSM
sym is the Standard Model
(SM) Lagrangian, without the neutrino and chargedlepton Yukawa interactions mediated by the couplings
hlα and yl m , respectively. The Lagrangian Lsym is
then invariant under the lepton-flavour symmetry group
GLF = U (3)L ×U (3)eR ×O(3)NR , where L collectively denotes the three lepton doublets, and eR and NR the three
right-handed charged leptons and neutrinos, respectively.
According to the MFV hypothesis, the flavour symmetry
EUROPEAN ORGANIZATION FOR NUCLEAR RESEARCH
CERN-PH-EP-2015-155
22 June 2015
arXiv:1506.08032v1 [nucl-ex] 26 Jun 2015
Forward-central two-particle correlations
√
in p–Pb collisions at sNN = 5.02 TeV
ALICE Collaboration∗
Abstract
Two-particle angular correlations between trigger particles in the forward pseudorapidity
range (2.5 < |η| < 4.0) and associated particles in the central range (|η| < 1.0) are measured with the ALICE detector in p–Pb collisions at a nucleon–nucleon centre-of-mass energy of 5.02 TeV. The trigger particles are reconstructed using the muon spectrometer, and
the associated particles by the central barrel tracking detectors. In high-multiplicity events,
the double-ridge structure, previously discovered in two-particle angular correlations at
midrapidity, is found to persist to the pseudorapidity ranges studied in this Letter. The
second-order Fourier coefficients for muons in high-multiplicity events are extracted after
jet-like correlations from low-multiplicity events have been subtracted. The coefficients are
found to have a similar transverse momentum (pT ) dependence in p-going (p–Pb) and Pbgoing (Pb–p) configurations, with the Pb-going coefficients larger by about 16 ± 6%, rather
independent of pT within the uncertainties of the measurement. The data are compared with
calculations using the AMPT model, which predicts a different pT and η dependence than
observed in the data. The results are sensitive to the parent particle v2 and composition of
reconstructed muon tracks, where the contribution from heavy flavour decays are expected
to dominate at pT > 2 GeV/c.
c 2015 CERN for the benefit of the ALICE Collaboration.
Reproduction of this article or parts of it is allowed as specified in the CC-BY-4.0 license.
∗ See
Appendix A for the list of collaboration members
OUTP-15-06P
DCPT/15/32
IPPP/15/16
Charm production in the forward region: constraints on
the small-x gluon and backgrounds for neutrino astronomy
Rhorry Gauld1 , Juan Rojo2 , Luca Rottoli2 and Jim Talbert2
arXiv:1506.08025v1 [hep-ph] 26 Jun 2015
1
2
Institute for Particle Physics Phenomenology,
Durham University, Durham DH1 3LE, UK
Rudolf Peierls Centre for Theoretical Physics, 1 Keble Road,
University of Oxford, OX1 3NP Oxford, UK
Abstract
The recent observation by the IceCube experiment of cosmic neutrinos at energies up to a
few PeV heralds the beginning of neutrino astronomy. At such high energies, the ‘conventional’
neutrino flux is suppressed and the ‘prompt’ component from charm meson decays is expected to
become the dominant background to astrophysical neutrinos. Charm production at high energies
is however theoretically uncertain, both since the charm mass is at the boundary of applicability
of perturbative QCD, and also because the calculations are sensitive to the poorly-known gluon
PDF at small-x. In this work we provide detailed perturbative QCD predictions for charm and
bottom production in the forward region, and validate them by comparing with recent data
from the LHCb experiment at 7 TeV. Finding good agreement between data and theory, we use
the LHCb measurements to constrain the small-x gluon PDF, achieving a substantial reduction
in its uncertainties. Using these improved PDFs, we provide predictions for charm and bottom
production at LHCb at 13 TeV, as well as for the ratio of cross-sections between 13 and 7 TeV.
The same calculations are used to compute the energy distribution of neutrinos from charm
decays in pA collisions, a key ingredient towards achieving a theoretically robust estimate of
charm-induced backgrounds at neutrino telescopes.
1
Non-minimal UED confronts Bs → µ+µ−
arXiv:1506.08024v1 [hep-ph] 26 Jun 2015
Anindya Datta ∗ , Avirup Shaw
†
Department of Physics, University of Calcutta, 92 Acharya Prafulla Chandra Road,
Kolkata 700009, India
Abstract
Addition of boundary localised kinetic and Yukawa terms to the action of a 5dimensional Standard Model would non-trivially modify the Kaluza-Klein spectra and
some of the interactions among the Kaluza-Klein excitations compared to the minimal
version of this model, in which, these boundary terms are not present. In the minimal
version of this framework known as Universal Extra Dimensional model, special assumptions are made about these unknown, beyond the cut-off contributions to restrict the
number of unknown parameters of the theory to a minimal. We estimate the contribution of Kaluza-Klein modes to the branching ratios of Bs(d) → µ+ µ− in the framework
of non-minimal Universal Extra Dimensional, at one loop level. The results have been
compared to the experimental data to constrain the parameters of this model. From the
measured decay branching ratio of Bs → µ+ µ− (depending on the values of boundary
localised parameters) lower limit on R−1 can be as high as 2 TeV. We have revisited the
bound on R−1 in minimal Universal Extra Dimensional model, which came out to be 1.26
TeV. This limit on R−1 in the minimal framework is comparable with that derived from
the consideration of relic density or Standard Model Higgs boson production and decay to
W + W − . Unfortunately, Bd → µ+ µ− decay branching ratio would not set any significant
limit on R−1 in a minimal or non-minimal Universal Extra Dimensional model.
PACS No: 11.10 Kk, 12.60.-i, 14.70.Hp, 14.80.Rt
I
Introduction
After the discovery of the Higgs boson at the LHC experiment, the new challenge to particle
physics is to provide a framework in which there exists a natural Dark Matter (DM) candidate,
as the Standard Model (SM) itself does not have a sufficiently massive weakly interacting
particle to be a good candidate for DM. Thus one is compelled to look beyond the SM and in
this endeavour, extra dimensional scenarios offer such a paradigm [1]. Some of the variants of
extra dimensional theories offer the solution to the DM puzzle along with many others like gauge
coupling unifications [2] and fermion mass hierarchy [3] to name a few. A particular extension
of the SM needs special attention in this regard. This is known as Universal Extra Dimensional
(UED) Model where all the SM fields can propagate in 4 + 1 dimensional space-time. The extra
∗
†
email: [email protected]
email: [email protected]
1
DESY 15-097, IPPP/15/38, DCPT/15/76, SLAC-PUB-16316
hhjj production at the LHC
Matthew J. Dolan,1, 2, ∗ Christoph Englert,3, † Nicolas Greiner,4, ‡ Karl Nordstrom,3, § and Michael Spannowsky5, ¶
arXiv:1506.08008v1 [hep-ph] 26 Jun 2015
1
Theory Group, SLAC National Accelerator Laboratory,
Menlo Park, CA 94025, USA
2
ARC Centre of Excellence for Particle Physics at the Terascale,
School of Physics, University of Melbourne, 3010, Australia
3
SUPA, School of Physics and Astronomy, University of Glasgow,
Glasgow, G12 8QQ, United Kingdom
4
DESY Theory Group, Notkestr. 85, D-22607 Hamburg, Germany
5
Institute for Particle Physics Phenomenology, Department of Physics,
Durham University, DH1 3LE, United Kingdom
The search for di-Higgs production at the LHC in order to set limits on Higgs trilinear coupling
and constraints on new physics is one of the main motivations for the LHC high luminosity phase.
Recent experimental analyses suggest that such analyses will only be successful if information from
a range of channels is included. We therefore investigate di-Higgs production in association with two
hadronic jets and give a detailed discussion of both the gluon- and weak boson fusion contributions,
with a particular emphasis on the phenomenology with modified Higgs trilinear and quartic gauge
couplings. We perform a detailed investigation of the full hadronic final state and find that hhjj
production should add sensitivity to a di-Higgs search combination at the HL-LHC with 3 ab−1 .
Since the WBF and GF contributions are sensitive to different sources of physics beyond the Standard Model, we devise search strategies to disentangle and isolate these production modes. While
gluon fusion remains non-negligible in WBF-type selections, sizeable new physics contributions to
the latter can still be constrained. As an example of the latter point we investigate the sensitivity
that can be obtained for a measurement of the quartic Higgs-gauge boson couplings.
I.
INTRODUCTION
After the Higgs boson discovery in 2012 [1] and subsequent analyses of its properties [2], evidence for physics
beyond the Standard Model (BSM) remains elusive. Although consistency with SM Higgs properties is expected
in many BSM scenarios, current measurements do not
fully constrain the Higgs sector. One coupling which is
currently unconstrained and has recently been subject of
much interest is the Higgs self-interaction ∼ η, which is
responsible for the spontaneous breaking of electroweak
gauge symmetry in the SM via the potential
V (H † H) = µ2 H † H + η(H † H)2 ,
(1)
√
with µ2 < 0, where H = (0, v + h)T / 2 in unitary
gauge. The Higgs self-coupling manifests itself primarily in a destructive interference in gluon fusion-induced
di-Higgs production [3–5] through feeding into the
p trilinear Higgs interaction with strength λSM = mh η/2 =
gm2h /(4mW ) in the SM. The latter relation can be altered in BSM scenarios, e.g. the SM coupling pattern
can be distorted by the presence of a dimension six operator ∼ (H † H)3 , and di-Higgs production is the only
∗ Electronic
address:
address:
‡ Electronic address:
§ Electronic address:
¶ Electronic address:
† Electronic
[email protected]
[email protected]
[email protected]
[email protected]
[email protected]
channel with direct sensitivity to this interaction [6]. A
modification solely of the Higgs trilinear coupling, which
is typically invoked in di-Higgs feasibility studies, is predicted in models of µ2 -less electroweak symmetry breaking, e.g. [7].
After the Higgs discovery, analyses of the di-Higgs final
state at the high-luminosity LHC and beyond have experienced a renaissance, and di-Higgs final states such as
the b¯bγγ [6, 8–10], b¯bτ + τ − [11–13], b¯bW + W − [11, 12, 14]
and b¯bb¯b [11, 12, 15] channels have been studied phenomenologically, often relying on boosted jet substructure techniques [16] (see also an investigation [17] of rare
decay channels relevant for a 100 TeV collider). Recent
analyses by ATLAS and CMS [18, 19] have highlighted
the complexity of these analyses and the necessity to explore different production mechanisms to formulate constraints on the Higgs self-interactions in the future. This
program has already been initiated by feasibility analyses of the hhj, hhjj and tt¯hh production modes in
Refs. [10, 12, 20–22].
Di-Higgs production in association with two jets is a
particularly important channel in this regard since this
final state receives contributions from the weak boson
fusion (WBF) production mode. The phenomenological
appeal of the WBF mode is twofold. Firstly, the weak
boson fusion component of pp → hhjj is sensitive to
modifications of the gauge-Higgs sector [20, 23], which
can lead to large cross-section enhancements. Secondly,
the QCD uncertainties for the WBF topologies are known
and under theoretical control [24, 25], such that a search
for BSM electroweak-induced deviations is not hampered
June 29, 2015
Determination of CP violation parameter using neutrino pair beam
arXiv:1506.08003v1 [hep-ph] 26 Jun 2015
M. Yoshimura and N. Sasao†
Center of Quantum Universe, Faculty of Science, Okayama University
Tsushima-naka 3-1-1 Kita-ku Okayama 700-8530 Japan
†
Research Core for Extreme Quantum World, Okayama University
Tsushima-naka 3-1-1 Kita-ku Okayama 700-8530 Japan
ABSTRACT
Neutrino oscillation experiments under neutrino pair beam from circulating excited heavy ions are studied. It is found that detection of double weak events has a good sensitivity to measure CP violating parameter
and distinguish mass hierarchy patterns in short baseline experiments in which the earth-induced matter
effect is minimized.
PACS numbers
13.15.+g, 14.60.Pq,
Keywords
CP violation, neutrino oscillation, CP-even neutrino beam, heavy ion synchrotron, earth
matter effect
1
PITT PACC 1508
arXiv:1506.07885v1 [hep-ph] 25 Jun 2015
Revealing Compressed Stops Using
High-Momentum Recoils
Sebastian Macaluso1 , Michael Park1,2 , David Shih1 , and Brock Tweedie3
1
3
NHETC, Department of Physics and Astronomy
Rutgers University, Piscataway, NJ 08854
2
Stanford University,
Stanford, CA 94305
PITT PACC, Department of Physics and Astronomy
University of Pittsburgh, Pittsburgh, PA 15260
Searches for supersymmetric top quarks at the LHC have been making great progress in
pushing sensitivity out to higher mass, but are famously plagued by gaps in coverage around
lower-mass regions where the decay phase space is closing off. Within the common stopNLSP / neutralino-LSP simplified model, the line in the mass plane where there is just
enough phase space to produce an on-shell top quark remains almost completely unconstrained. Here, we show that is possible to define searches capable of probing a large patch
of this difficult region, with S/B ∼ 1 and significances often well beyond 5σ. The basic
strategy is to leverage the large energy gain of LHC Run 2, leading to a sizable population
of stop pair events recoiling against a hard jet. The recoil not only re-establishes a 6ET
signature, but also leads to a distinctive anti-correlation between the 6ET and the recoil jet
transverse vectors when the stops decay all-hadronically. Accounting for jet combinatorics,
backgrounds, and imperfections in 6ET measurements, we estimate that Run 2 will already
start to close the gap in exclusion sensitivity with the first few 10s of fb−1 . By 300 fb−1 ,
exclusion sensitivity may extend from stop masses of 550 GeV on the high side down to
below 200 GeV on the low side, approaching the “stealth” point at mt˜ = mt and potentially
overlapping with limits from tt¯ cross section and spin correlation measurements.
I.
INTRODUCTION
Light stops with mass below a TeV are extremely well-motivated by the supersymmetric
solution to the hierarchy problem. The uniquely important role of these particles has inspired
a growing and increasingly sophisticated set of dedicated searches at the LHC, targeting an
array of different possible decay topologies [1–17] (see also [18, 19]). While these searches
have already probed significant portions of the possible model space below a TeV, sizable
gaps in coverage remain even at O(100 GeV), leaving us to consider: Is it possible that
light stops have already been produced in abundance in LHC Run 1 but have simply been
missed?
In perhaps the most minimalistic benchmark scenario, stops are produced directly in pairs
via QCD, and each stop undergoes a one-step R-parity-conserving cascade into an invisible
neutralino LSP and an on-shell or off-shell top quark:
pp → t˜t˜∗ ,
t˜ → t(∗) + χ˜0
(1)
The visible composition of the final state is then identical to that of tt¯, which serves as
a copious background. The main kinematic handle exploited in most searches has been
the additional injection of 6ET (or more properly 6pT ) from the neutralinos. For mt˜ mχ˜ ,
exclusion limits from tt¯ + 6ET searches at Run 1 extend beyond 700 GeV [10]. However,
such searches face a major challenge when confronted with lower-mass regions in the stopneutralino mass plane where the 6ET is squeezed out. In particular, much attention has
recently been directed at the “top compression line” mt˜ ' mχ˜ + mt , which defines the
boundary between two-body decays into an on-shell top quark and neutralino, and threebody decays via an off-shell top quark into W bχ˜0 . Limits along this compression line are
largely nonexistent over a roughly 20 GeV-wide gap in stop mass.
Proposals to probe this region using the total tt¯ cross section and spin correlations [20,
21] have led to some inroads near the so-called “stealth” point (mt˜, mχ˜ ) = (mt , 0) [1, 4],
but theoretical limitations make it unclear if these searches can be pushed much further.
The relatively long lifetimes of stops very near to the top compression line has led to a
complementary suggestion to use the annihilation-decays of stoponium [22–24], which would
lead to distinctive resonant diboson signatures (including, e.g., γγ and Zγ). Projections for
Run 2 predict sensitivity up to stop masses of several hundred GeV, depending in detail on
the stop chirality admixture. However, these searches become insensitive if the individual
stops decay more quickly than the stoponium, which generally occurs as soon as the stopneutralino mass difference opens up to even O(GeV). Other approaches have sought to use
the small amount of 6ET that is available within the bulk of the produced stop pair events.
Very detailed measurements of the shapes of the tails of 6ET -sensitive observables may be
promising [25], but a careful accounting of theoretical and experimental errors is not always
available, and the one measurement of this type that has been carried out [2] (by ATLAS,
1
EUROPEAN ORGANIZATION FOR NUCLEAR RESEARCH
CERN-PH-EP-2015-106
23 Apr 2015
arXiv:1506.07884v1 [nucl-ex] 25 Jun 2015
One-dimensional pion, kaon, and proton femtoscopy
√
in Pb–Pb collisions at s NN = 2.76 TeV
ALICE Collaboration∗
Abstract
The size of the particle emission region in high-energy collisions can be deduced using the femtoscopic correlations of particle pairs at low relative momentum. Such correlations arise due to quantum statistics and Coulomb and strong final state interactions. In this paper, results are presented
from femtoscopic analyses of π ± π ± , K± K± , K0S K0S , pp, and pp correlations from Pb-Pb collisions
√
at sNN = 2.76 TeV by the ALICE experiment at the LHC. One-dimensional radii of the system
are extracted from correlation functions in terms of the invariant momentum difference of the pair.
The comparison of the measured radii with the predictions from a hydrokinetic model is discussed.
The pion and kaon source radii display a monotonic decrease with increasing average pair transverse
mass mT which is consistent with hydrodynamic model predictions for central collisions. The kaon
and proton source sizes can be reasonably described by approximate mT -scaling.
c 2015 CERN for the benefit of the ALICE Collaboration.
Reproduction of this article or parts of it is allowed as specified in the CC-BY-4.0 license.
∗ See
Appendix A for the list of collaboration members
MIT-CTP 4685
Separated at Birth: Jet Maximization, Axis Minimization, and Stable Cone Finding
Jesse Thaler∗
arXiv:1506.07876v1 [hep-ph] 25 Jun 2015
Center for Theoretical Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA
Jet finding is a type of optimization problem, where hadrons from a high-energy collision event
are grouped into jets based on a clustering criterion. As three interesting examples, one can form
a jet cluster that (1) optimizes the overall jet four-vector, (2) optimizes the jet axis, or (3) aligns
the jet axis with the jet four-vector. In this paper, we show that these three approaches to jet
finding, despite being philosophically quite different, can be regarded as descendants of a mother
optimization problem. For the special case of finding a single cone jet of fixed opening angle,
the three approaches are genuinely identical when defined appropriately. This relationship is only
approximate for cone jets in the rapidity-azimuth plane, as used at the Large Hadron Collider,
though the differences are mild for small radius jets.
I.
INTRODUCTION
Jet algorithms are essential tools for connecting longdistance measurements made on hadrons to shortdistance interpretations based on perturbative quarks
and gluons. Since there is a fundamental mismatch between color-singlet hadrons and color-carrying partons,
there is no way to define an ideal jet finding procedure.
For this reason, a variety of jet algorithms have been introduced with different underlying philosophies and different practical advantages [1, 2].
In this paper, we expose a surprising connection between three seemingly unrelated approaches to jet finding: jet function maximization [3–5], 1-jettiness minimization [6–9], and stable cone finding [10–12], all reviewed in Sec. II. Philosophically, these algorithms are
quite different, so one might think that they would yield
rather different jets. Instead, all three algorithms yield
(approximately) conical jets of (approximately) fixed radius R, with a high degree of correlation between the
methods. As we will show, this correlation is not an accident, since jet function maximization and 1-jettiness
minimization can be viewed as descendants of a mother
optimization problem, whose solution is a stable cone jet.
This relationship is most transparent in electronpositron collisions, where jets are typically defined in
terms of particle energies and angles. Remarkably, with
appropriate definitions, all three algorithms can be made
to yield identical cone jets. We will prove this in Sec. III
by showing how these three methods can be derived from
optimizing a common meta function. In Sec. IV, we explain this relationship in more detail by performing a two
particle case study.
Turning to the Large Hadron Collider (LHC), the intrinsic differences between the three algorithms become
apparent. In proton-proton collisions, jets are typically
defined in terms of particle transverse momenta and
rapidity-azimuth distances. As discussed in Sec. V, the
transverse momentum of a jet is not exactly the same as
the summed transverse momenta of its constituents. For
this reason, there is now a difference between optimizing
a four-vector (as in jet functions), optimizing a light-like
axis (as in 1-jettiness), and aligning the jet axis with the
jet momentum (as in stable cones). Despite these differences, though, we show that the algorithms still give
similar results when the jet radius R is small.
II.
To begin, we briefly review these three approaches to
jet finding, all of which are infrared and collinear (IRC)
safe. For simplicity, in this paper we only consider finding
the single hardest jet in an event, though all of these
methods can be adapted to identify multiple jets. For
consistency of notation, we always use R to refer to the
adjustable jet radius parameter in each algorithm, which
can be identified with the true jet radius only in the small
R limit.
• Jet function maximization [3–5]. Here, the jet finding strategy is to find the subset of particles in an
event that maximizes a jet function J(Pµ ), where
Pµ is the four-vector of the candidate jet.1 The
original paper [3] introduced a jet function appropriate for electron-positron collisions:
Jorig (Pµ ) = E −
[email protected]
1 m2
,
R2 E
(1)
where E and m are the total energy and mass of the
candidate jet, and the original paper used the notation R2 → 1/β. Maximizing this Jorig gives quasiconical jets with nearly fixed radius ≃ R. Since
then, this algorithm has been adapted to collisions
at the LHC [5] to yield nearly conical jets in the
rapidity-azimuth plane.
• 1-jettiness minimization [6–9]. This approach is
based on finding a light-like jet axis n = (1, n
ˆ ) that
1
∗
REVIEW OF JET ALGORITHMS
The name “jet function” and the symbol J should not be confused with earlier usage in the context of factorization and resummation, e.g. [13–15].
On the determination of the leptonic CP phase
Jessica Elevant∗ , Thomas Schwetz† ,
Oskar Klein Centre for Cosmoparticle Physics,
Department of Physics, Stockholm University, SE-10691 Stockholm, Sweden
Abstract
arXiv:1506.07685v1 [hep-ph] 25 Jun 2015
The combination of data from long-baseline and reactor oscillation experiments
leads to a preference of the leptonic CP phase δCP in the range between π and 2π. We
study the statistical significance of this hint by performing a Monte Carlo simulation
of the relevant data. We find that the distribution of the standard test statistic used to
derive confidence intervals for δCP is highly non-Gaussian and depends on the unknown
true values of θ23 and the neutrino mass ordering. Values of δCP around π/2 are
disfavored at between 2σ and 3σ, depending on the unknown true values of θ23 and the
mass ordering. Typically the standard χ2 approximation leads to over-coverage of the
confidence intervals for δCP . For the 2-dimensional confidence region in the (δCP , θ23 )
plane the usual χ2 approximation is better justified. The 2-dimensional region does
not include the value δCP = π/2 up to the 81.8% (83.9%) CL assuming a true normal
(inverted) mass ordering. Furthermore, we study the sensitivity to δCP and θ23 of
an increased exposure of the T2K experiment, roughly a factor 12 larger than the
current exposure and including also anti-neutrino data. Also in this case deviations
from Gaussianity may be significant, especially if the mass ordering is unknown.
∗
†
jessica.elevant AT fysik.su.se
schwetz AT fysik.su.se
Prospects for measurement of the neutrino mass hierarchy
R. B. Patterson1
arXiv:1506.07917v1 [hep-ex] 25 Jun 2015
1
Division of Physics, Mathematics, and Astronomy, California Institute of Technology,
Pasadena, California 91125; email: [email protected]
The unknown neutrino mass hierarchy – whether the ν3 mass eigenstate is the heaviest or the
lightest – represents a major gap in our knowledge of neutrino properties. Determining the hierarchy
is a critical step toward further precision measurements in the neutrino sector. The hierarchy is also
central to interpreting the next generation of neutrinoless double beta decay results, plays a role
in numerous cosmological and astrophysical questions, and serves as a powerful model discriminant
for theories of neutrino mass generation and unification. Various current and planned experiments
claim sensitivity for establishing the neutrino mass hierarchy. We review the most promising of these
here, paying special attention to points of concern and consolidating the projected sensitivities into
an outlook for the years ahead.
INTRODUCTION
In the Standard Model with non-zero neutrino masses, the three neutrino flavor eigenstates νe , νµ , and ντ are
non-trivial linear combinations of the three mass eigenstates ν1 , ν2 , and ν3 . The complex matrix U relating the flavor
and mass eigenstates to each other, called the PMNS matrix for the authors of Refs. [1–3], can be parametrized in
terms of three rotation angles and three complex phases:
  iα /2




e 1
0
0
c13
0 s13 e−iδ
0
0 1
1 0
0
  c12 s12 0   0
0
1
0
(1)
U =  0 c23 s23  
eiα2 /2 0  ,
iδ
−s12 c12 0
0 −s23 c23
−s13 e 0
c13
0
0
1
with cij = cos θij and sij = sin θij . While this neutrino mixing gives rise to several physical phenomena, the most
experimentally fruitful one so far has been neutrino oscillations, whereby the flavor composition of a neutrino can
vary as it propagates. The probability P that a relativistic neutrino produced with flavor l will be detected as a
neutrino of flavor l0 after traveling a distance L through vacuum is given by
!
2
X
∆m
L
ij
P (νl → νl0 ) = δll0 − 4
<(Uli∗ Ul0 i Ulj Ul∗0 j ) sin2
4E
i>j
!
(2)
2
X
∆m
L
ij
+ 2
=(Uli∗ Ul0 i Ulj Ul∗0 j ) sin2
,
2E
i>j
where E is the neutrino’s energy, ∆m2ij ≡ m2i − m2j is the difference between the squares of the masses of the mass
eigenstates νi and νj , δll0 is the Kronecker delta, and the matrix elements Uli are defined by


Ue1 Ue2 Ue3
U =  Uµ1 Uµ2 Uµ3  .
(3)
Uτ 1 Uτ 2 Uτ 3
Our quantitative understanding of the neutrino sector has grown tremendously in the past two decades primarily due
to neutrino oscillation experiments. The “solar” mass splitting ∆m221 [4] and the thirty-times-larger “atmospheric”
mass splitting ∆m232 [5] have been measured with impressive precision:
∆m221 = (7.53 ± 0.18)×10−5 eV2
∆m232
= (2.34 ± 0.09)×10
−3
eV
2
(4)
[NH]
or
−3
(−2.37+0.07
−0.11 )×10
eV
2
[IH] .
(5)
The normal hierarchy (NH) and inverted hierarchy (IH) distinction is defined below.[61] In addition to these mass
splittings, we know at varying levels of precision the values of the three mixing angles θ12 , θ13 , and θ23 [6]. We do
not have good information about the sign of the atmospheric mass splitting, the complex phases in the PMNS matrix
of Eq. (1), the absolute masses {mi } of the neutrinos, or whether neutrinos are Majorana particles. Furthermore,
several experimental anomalies suggest that extensions to the neutrino standard model may be needed [7].
Preprint typeset in JHEP style - HYPER VERSION
DCPT-15/39
arXiv:1506.08130v1 [hep-th] 26 Jun 2015
Contact interactions between particle
worldlines
James P. Edwards
Centre for Particle Theory, Department of Mathematical Sciences,
University of Durham, Durham DH1 3LE, UK
Email: [email protected]
Abstract: We construct contact interactions for bosonic and fermionic point particles. We first relate the resulting theories to classical electrostatics by constructing
functional averages over worldlines whose endpoints are fixed to charged particles.
Counting those paths which pass through a space-time point xµ gives the static electric field at that point, provided we take the limit where the length measured along
the worldlines is large. We also investigate corrections to the classical field that arise
beyond leading order in this limit before constructing a theory of point particles
that interact when their worldlines intersect. We quantise this theory and show that
the partition function contains propagator couplings between the endpoints of the
particles before discussing how this is related to the worldline formalism of quantum
field theory and action at a distance theories.
Keywords: Electromagnetism, Supersymmetry.
CERN-PH-TH-2015-147
Observational Hints of a Pre–Inflationary Scale?
A. Gruppuso a,b and A. Sagnotti c,d
arXiv:1506.08093v1 [astro-ph.CO] 26 Jun 2015
a
INAF-IASF Bologna,
Istituto di Astrofisica Spaziale e Fisica Cosmica di Bologna,
Istituto Nazionale di Astrofisica,
via Gobetti 101, I-40129 Bologna, Italy
b
INFN, Sezione di Bologna,
Via Irnerio 46, I-40126 Bologna, Italy
e-mail: [email protected]
c
Department of Physics, CERN Theory Division
CH - 1211 Geneva 23, Switzerland
d
Scuola Normale Superiore and INFN
Piazza dei Cavalieri 7
I-56126 Pisa Italy
e-mail: [email protected]
Abstract
We argue that the lack of power exhibited by cosmic microwave background (CMB) anisotropies
at large angular scales might be linked to the onset of inflation. We highlight observational
features and theoretical hints that support this view, and present a preliminary estimate of the
physical scale that would underlie the phenomenon.
Essay Written for the 2015 Gravity Research Foundation Awards for Essays on Gravitation.
Selected for Honorable Mention.
Electromagnetic modulation of monochromatic
neutrino beams
arXiv:1506.07883v1 [nucl-th] 25 Jun 2015
2
A. L. Barabanov1,2∗, O. A. Titov1†
1
NRC "Kurchatov Institute" , 123182, Moscow, Russia
Moscow Institute of Physics and Technology, 141700, Dolgoprudny,
Moscow Region, Russia
Abstract
A possibility to produce a modulated monochromatic neutrino beam is discussed.
Monochromatic neutrinos can be obtained in electron capture by nuclei of atoms or
ions, in particular, by nuclei of hydrogen-like ions. It is shown that monochromatic
neutrino beam from such hydrogen-like ions with nuclei of non-zero spin can be
modulated because of different probabilities of electron capture from hyperfine states.
Modulation arises by means of inducing of electromagnetic transitions between the
hyperfine states. Requirements for the hydrogen-like ions with necessary properties
are discussed. A list of the appropriate nuclei for such ions is presented.
1
Introduction
”Clean” neutrino beams (with neutrinos of single flavour) with known intensity and energy
spectrum are of great interest for studying the neutrino properties and the details of
neutrino-involving processes. A comprehensive list of applications for such beams is given,
e.g., in [1]. These applications include measurements of neutrino oscillation parameters
(in particular, for oscillations to sterile states), search for neutrino magnetic moment,
refinement of the weak interaction constants values, study of coherent neutrino scattering
off nuclei, investigation of elastic and inelastic neutrino-nucleon and neutrino-nucleus
scattering processes (including astrophysical problems).
In [2] it was proposed to use accelerated β + - and β − - decaying nuclei (or radioactive
ions), held in a storage ring, as a source for such neutrino beams (β-beams). The neutrinos
are emitted isotropically in the rest frame of the nuclei. However, in the laboratory frame,
half of the neutrinos are emitted in the forward direction within a cone with the opening
angle θ ≃ 1/γ. Therefore, if the nuclei are accelerated to γ ≫ 1, the neutrino beam is
collimated. The neutrinos, which are emitted forwardly with the energy Eν0 in the rest
frame of the nuclei, have much larger energy Eν ≃ 2γEν0 ≫ Eν0 in the laboratory frame.
So, one can control the neutrino spectrum by varying the value of γ.
∗
†
e-mail: [email protected]
e-mail: [email protected]
1
Dark Matter Superfluidity and Galactic Dynamics
Lasha Berezhiani and Justin Khoury1
arXiv:1506.07877v1 [astro-ph.CO] 25 Jun 2015
1
Center for Particle Cosmology, Department of Physics and Astronomy,
University of Pennsylvania, Philadelphia, PA 19104, USA
We propose a unified framework that reconciles the stunning success of MOND on galactic scales
with the triumph of the ΛCDM model on cosmological scales. This is achieved through the physics
of superfluidity. Dark matter consists of self-interacting axion-like particles that thermalize and
condense to form a superfluid in galaxies, with ∼mK critical temperature. The superfluid phonons
mediate a MOND acceleration on baryonic matter. Our framework naturally distinguishes between
galaxies (where MOND is successful) and galaxy clusters (where MOND is not): dark matter has a
higher temperature in clusters, and hence is in a mixture of superfluid and normal phase. The rich
and well-studied physics of superfluidity leads to a number of striking observational signatures.
The standard Λ Cold Dark Matter (ΛCDM) model
does very well at fitting large scale observables. On galactic scales, however, a number of challenges have emerged.
Disc galaxies display a tight correlation between total
baryonic mass and asymptotic velocity, Mb ∼ vc4 , known
as the Baryonic Tully-Fisher Relation (BTFR) [1, 2]. Hydrodynamical simulations can reproduce the BTFR by
tuning baryonic feedback processes, but their stochastic
nature naturally results in a much larger scatter [3]. Furthermore, the mass [4, 5] and phase-space [6–9] distributions of dwarf satellites in the Local Group are puzzling.
A radical alternative is MOdified Newtonian Dynamics (MOND) [10], which replaces dark matter (DM) with
a modification of gravity at low acceleration: a ' aN
√
(aN a0 ); a ' aN a0 (aN a0 ), with best-fit value
a0 ' 1.2 × 10−8 cm/s2 . This empirical force law has
been remarkably successful at explaining a wide range of
galactic phenomena [11]. In the MOND regime, a test
particle orbits an isolated source according to v 2 /r =
p
GN Mb a√0 /r2 . This gives a constant asymptotic velocity, vc2 = GN Mb a0 , which in turn implies the BTFR.
The empirical success of MOND, however, is limited
to galaxies. The predicted temperature profile in galaxy
clusters conflicts with observations [12]. The TensorVector-Scalar (TeVeS) relativistic extension [13] fails to
reproduce the CMB and matter spectra [14, 15]. The
lensing features of merging clusters [16, 17] are problematic [18]. This has motivated various hybrid proposals
that include both DM and MOND, e.g., [19–23].
In this Letter, together with a longer companion paper [24], we propose a novel framework that unifies the
DM and MOND phenomena through the physics of superfluidity. Our central idea is that DM forms a superfluid inside galaxies, with a coherence length of order the
size of galaxies. The critical temperature is ∼ mK, which
intriguingly is comparable to Bose-Einstein condensation
(BEC) critical temperatures for cold atom gases. Indeed,
in many ways our DM behaves like cold dark atoms.
The superfluid nature of DM dramatically changes its
macroscopic behavior in galaxies. Instead of evolving
as independent particles, DM is more aptly described as
collective excitations. Superfluid phonons, in particular, mediate a MOND-like force between baryons. Since
superfluidity only occurs at low enough temperature,
our framework naturally distinguishes between galaxies
(where MOND is successful) and galaxy clusters (where
MOND is not). Due to the larger velocity dispersion in
clusters, DM has a higher temperature and hence is in a
mixture of superfluid and normal phases [25–27].
The superfluid interpretation makes the non-analytic
nature of the MOND scalar action more palatable. The
Unitary Fermi Gas, which has attracted much excitement
in cold atom physics [28], is also governed by a nonanalytic kinetic term [29]. Our equation of state P ∼ ρ3
suggests that the DM superfluid arises through threebody interactions. It would be fascinating to find precise
cold atom systems with the same equation of state, as
this would give important insights on the microphysical
interactions underlying our superfluid. Tantalizingly, this
might allow laboratory simulations of galactic dynamics.
The idea of DM BEC has been studied before [30–34],
with important differences from our work. In BEC DM
galactic dynamics are caused by the condensate density
profile; in our case phonons play a key role in explaining
the BTFR. Moreover, BEC DM has P ∼ ρ2 instead of ∼
ρ3 . This implies a much lower sound speed, which puts
BEC DM in tension with observations [35].
DM Condensation: In order for DM particles to condense in galaxies, their de Broglie wavelength λdB ∼
(mv)−1 must be larger than the interparticle separation
1/3
` ∼ (m/ρvir ) . From standard collapse theory, the density at virialization is ρvir ' (1+zvir )3 5.4×10−28 g/cm3 ,
1/3 √
while the virial velocity is v = 113 M12 1 + zvir km/s,
> ` implies
where M12 ≡ M/1012 M . Thus λdB ∼
3/8
m<
∼ 2.3 (1 + zvir )
−1/4
M12
eV .
(1)
The second condition is that DM thermalizes, with
temperature set by the virial velocity v. The interac(2π)3
σ
vir
tion rate is Γ ∼ N vρvir m
, where N ∼ ρm
4π
3 is the
3 (mv)
Bose enhancement factor. The rate should be larger than
the inverse dynamical time tdyn ∼ √G 1 ρ , such that the
N vir
coherence length will span the halo. This translates into
a bound on the cross section (with meV ≡ m/eV):
> (1 + zvir )
σ/m ∼
−7/2
2/3
m4eV M12 52 cm2 /g .
(2)
2
⌅(⇠)
1.0
m = 0.4 eV
1.0
0.8
Ncond
N
0.8
0.6
m = 0.6 eV
0.6
0.4
0.4
0.2
m = 0.8 eV
1011
1012
1013
M [h
1
1014
0.2
1015
0.5
M ]
1.0
1.5
2.0
2.5
⇠
FIG. 2: Numerical solution of Lane-Emden equation (9).
FIG. 1: Fraction of DM particles in the condensate.
Later on, we will adopt m = 0.6 eV as a fiducial value.
For M12 = 1 and zvir = 2, the inequality becomes
> 0.1 cm2 /g. The lower end is consistent with curσ/m ∼
rent constraints [36–38] on σ/m for self-interacting dark
matter (SIDM) [39], though these constraints must be
carefully revisited in the superfluid context.
The critical temperature, obtained by equipartition
kB Tc = 13 mvc2 , is in the mK range:
−5/3
Tc = 6.5 meV (1 + zvir )2 mK .
(3)
For 0 < T < Tc , the system is a mixture of condensate
and normal components. The fraction of condensed par3/2
ticles, 1 − (T /Tc )
[40], is shown in Fig. 1 as a function
< eV, galaxies
of halo mass assuming zvir = 0. For m ∼
are almost completely condensed while massive clusters
have a significant normal component.
Superfluid Phase: The relevant low-energy degrees of
freedom of a superfluid are phonons, described by a scalar
field θ. In the presence of a gravitational potential Φ,
the non-relativistic effective action is L = P (X), where
˙
~ 2 /2m [29]. We conjecture that DM suX = θ−mΦ−(
∇θ)
perfluid phonons are governed by the MOND action [41]
p
L = 23 Λ(2m)3/2 X |X| − α MΛPl θρb ,
(4)
where Λ is a mass scale, ρb is the baryonic matter density,
and α is a dimensionless constant. This action should
only be trusted away from X = 0, as we will see later.
The matter coupling breaks the shift symmetry at the
1/MPl level and is thus technically natural. Remarkably (4) is strikingly reminiscent of the Unitary Fermi
Gas, LUFG (X) ∼ X 5/2 , which is also non-analytic [29].
The phonon action (4) uniquely fixes the properties of
the condensate through standard thermodynamics. At
finite chemical potential, θ = µt, and ignoring Φ, the
pressure is given by the Lagrangian density:
P (µ) = 23 Λ(2mµ)3/2 .
(5)
This is the grand canonical equation of state P = P (µ)
for the condensate. The number density, n = ∂P/∂µ, is
n = Λ(2m)3/2 µ1/2 .
(6)
This is a polytropic equation of state P ∼ ρ1+1/n with
index n = 1/2. In comparison, BEC DM has P ∼ ρ2 [32].
Including phonons excitations θ = µt+φ, the quadratic
3/2
~ 2
action for φ is Lquad = Λ(2m)
φ˙ 2 − 2µ
m (∇φ) . The
4µ1/2
sound speed can be immediately read off:
p
cs = 2µ/m .
(8)
Using (7), we compute the static, sphericallysymmetric density profile of the DM condensate halo.
Introducing
dimensionless variables ρ = ρ0 Ξ and r =
q
ρ0
32πGN Λ2 m6 ξ, with ρ0 denoting the central density, hydrostatic equilibrium implies the Lane-Emden equation,
0
ξ 2 Ξ0 = −ξ 2 Ξ1/2 ,
(9)
with boundary conditions Ξ(0) = 1 and Ξ0 (0) = 0. The
numerical solution is shown in Fig. 2. It q
vanishes at ξ1 '
2.75, which defines the halo size: R = 32πGρN0Λ2 m6 ξ1 .
Meanwhile the central density is related to the halo mass
ξ1
3M
0
as [42] ρ0 = 4πR
3 |Ξ0 (ξ )| , with Ξ (ξ1 ) ' −0.5. Combining
1
these results, we can solve for ρ0 and R:
2/5
18/5
1/5
−6/5
6/5
ρ0 ' M12 meV ΛmeV 7 × 10−25 g/cm3 ;
ρ3
12Λ2 m6
.
(7)
(10)
where ΛmeV ≡ Λ/meV. Remarkably, with m ∼ eV and
Λ ∼ meV we obtain DM halos of realistic size. Concretely, as fiducial values we will fix
m = 0.6 eV ;
Λ = 0.2 meV .
(11)
This implies R ∼ 125 kpc for MDM = 1012 M . In general, we expect this superfluid core to be surrounded by
a cloud of DM particles in the normal phase, likely described by a Navarro-Frenk-White profile [43].
Including Baryons: We now derive the phonon profile θ = µt + φ(r) in galaxies, modeling baryons as
a static, spherically-symmetric
p
source. The equation
~ ·
~
of motion, ∇
2m|X| ∇φ
= αρb (r)/2MPl , where
X = µ − mΦ(r) − φ02 /2m, can be readily integrated:
Combining these expressions with ρ = mn, we obtain
P =
−2/5
R ' M12 meV ΛmeV 36 kpc ,
p
2m|X| φ0 =
αMb (r)
≡ κ(r) .
8πMPl r2
(12)
arXiv:1506.08185v1 [hep-ph] 26 Jun 2015
SLAC–PUB-16315
June, 2015
On the Trail of the Higgs Boson
Michael E. Peskin1
SLAC, Stanford University, Menlo Park, California 94025 USA
ABSTRACT
I review theoretical issues associated with the Higgs boson and the
mystery of spontaneous breaking of the electroweak gauge symmetry. This
essay is intended as an introduction to the special issue of Annalen der
Physik, “Particle Physics after the Higgs”.
Submitted to Annalen der Physik
1
Work supported by the US Department of Energy, contract DE–AC02–76SF00515.
SI-HEP-2015-15
QFET-2015-18
Inclusive weak decays of heavy hadrons with power suppressed terms at NLO
Thomas Mannel, Alexei A. Pivovarov, Denis Rosenthal
Theoretische Physik 1,Universit¨at Siegen, D-57068 Siegen, Germany
arXiv:1506.08167v1 [hep-ph] 26 Jun 2015
Abstract
Within the heavy quark expansion techniques for the heavy hadron weak decays we analytically
compute the coefficient of the power suppressed dimension five chromo-magnetic operator at
next-to-leading order of QCD perturbation theory with the full dependence on the final state
quark mass. We present explicit expressions for the total width of inclusive semileptonic decays
including the power suppressed terms and for a few moments of decay differential distributions.
One of the important phenomenological applications of our results is precision analysis of the
decays of bottom mesons to charmed final states and extraction of the numerical value for the
CKM matrix entry |Vcb |.
Neutrino Decay and Solar Neutrino Seasonal Effect
R. Picoreti, M. M. Guzzo, and P. C. de Holanda
Instituto de F´ısica Gleb Wataghin - UNICAMP, 13083-859, Campinas SP, Brazil
O. L. G. Peres
arXiv:1506.08158v1 [hep-ph] 26 Jun 2015
Instituto de F´ısica Gleb Wataghin - UNICAMP, 13083-859, Campinas SP, Brazil and
The Abdus Salam International Centre for Physics, Trieste, Italy
We consider the possibility of solar neutrino decay as a sub-leading effect on their propagation
between production and detection. Using current oscillation data, we set a new lower bound to
the ν2 neutrino lifetime at τ2 / m2 ≥ 7.2 × 10−4 s . eV−1 at 99% C.L.. Also, we show how seasonal
variations in the solar neutrino data can give interesting additional information about neutrino
lifetime.
INTRODUCTION
Beyond any reasonable doubt, it is now established
that neutrinos have non-zero, non-degenerate masses
and, thus, it would be possible - if not mandatory - for
them to decay into other particles.
Before the establishment of the LMA-MSW solution [1], decay was studied both by itself and in combination with standard flavor oscillations to explain the
difference between the expected solar neutrino flux from
nuclear fusion processes in the Sun and the detected flux
on Earth - the so-called Solar Neutrino Problem (SNP).
Although it is now ruled out as a leading process [2],
one can investigate neutrino decay as a sub-leading effect
in the propagation of solar neutrinos and set limits to
their lifetime using the most recent experimental data.
Solar neutrinos are produced in the nuclear fusion processes that power the Sun. In such processes, Hydrogen
nuclei are converted into Helium through several intermediate reactions, among which some produce neutrinos in
very particular spectra - both continuous and monochromatic.
Over the years, several experiments were developed
for the detection of solar neutrinos at different energy
ranges. From the pioneer Homestake [3] chlorine experiment - which first hinted at the SNP - through the gallium experiments GALLEX [4], SAGE [5] and GNO [6]
to the water Cherenkov detectors Kamiokande, Super
Kamiokande [7] and SNO [8]. Most recently, the Borexino [9] experiment also measured the 7 Be line.
The LMA-MSW solution, now considered the best explanation for the SNP, in combination with the measurement of the other oscillation parameters by experiments
designed for atmospheric, reactor and long-baseline neutrinos established the scenario of three massive light neutrinos that mix [1]. With such precise measurements of
the standard oscillation parameters, it is possible to investigate new phenomena such as the neutrino decay scenario: ν ′ → ν + X.
For solar neutrinos, the decay of the mass eigenstate
ν2 into the lighter state ν1 is disfavored by the data
and the current bound to ν2 lifetime for invisible nonradiative decays [2] is τ2 /m2 ≥ 8.7 × 10−5 s . eV−1 at
99% C.L.. Similarly, from the combined accelerator
and atmospheric neutrino data, ν3 lifetime is τ3 /m3 ≥
2.9 × 10−10 s . eV−1 at 90% C.L. [10]. Also, an analysis of long-baseline experiments MINOS and T2K give a
combined limit of τ3 /m3 ≥ 2.8 × 10−12 s . eV−1 at 90%
C.L. [11].
In this work, we consider the decay scenario in which
all the final products are invisible. We combine the available solar neutrino data with KamLAND [12] and Daya
Bay [13] data. For both experiments, the effect of neutrino decay is minimum thus allowing us to constrain the
standard neutrino mixing parameters independently of
the decay parameter τ2 /m2 and leading us to obtain a
robust bound on ν2 lifetime. Additionally, we show how
seasonal variations in the solar neutrino data, which are
enhanced by neutrino decay, can give some interesting
information about neutrino lifetime.
FORMALISM
After production in the solar core, neutrinos propagate
outwards undergoing flavor oscillation and resonant flavor transition due to the solar matter potential. After
emerging from the solar matter, they travel across the
interplanetary medium until they reach the Earth’s surface where they can be detected promptly (during the
day) or after traversing Earth’s matter (during the night
- on which they may also be subject to matter effects).
The transition amplitude for an electron neutrino produced in the Sun to be detected on Earth as a neutrino
of flavor α, νe → να , for the standard case of neutrino
oscillations with MSW effect, can be written as [14]:
X
vac ⊕
(1)
Aeα =
A⊙
ei Aii Aiα ,
where A⊙
ei is the transition amplitude of an electron neutrino produced in the solar core to be in a νi state in
the solar surface, Avac
ii is the propagation amplitude between Sun and Earth surfaces, and A⊕
iα is the transition
Gravitational mass of relativistic matter and antimatter
Tigran Kalaydzhyan∗
arXiv:1506.08063v1 [hep-ph] 21 Jun 2015
Department of Physics and Astronomy, Stony Brook University,
Stony Brook, New York 11794-3800, USA
The universality of free fall, the so-called weak equivalence principle (WEP), is a cornerstone of
the general theory of relativity, the most precise theory of gravity confirmed in all experiments up
to date. The WEP states the equivalence of the inertial and gravitational masses and was tested
in numerous occasions with normal matter at relatively low energies. However, there is no proof
for the matter and antimatter at high energies. For the antimatter the situation is even less clear
– current direct observations of trapped antihydrogen suggest the limits −65 < mg /m < 110 not
ruling out antigravity, i.e. repulsion of the antimatter by Earth. Here we demonstrate a bound
1 − 4 × 10−7 < mg /m . 1 + 2 × 10−7 on the gravitational mass of relativistic electrons and
positrons in the potential of the Local Supercluster (LS) coming from the Large Electron-Positron
Collider (LEP) and Tevatron accelerator experiments. By considering annual variations of the solar
gravitational potential instead of the absolute potentials we predict the bounds 0.96 < mg /m < 1.04
for an electron and positron, ruling out the speculated antigravity phenomenon. We also comment
on a possibility of performing complementary tests at the future International Linear Collider (ILC)
and Compact Linear Collider (CLIC).
Introduction.— Since the formulation of the general relativity (GR) by Einstein there were numerous
tests proving validity of the theory with an exceptional
precision1 . The weak equivalence principle (WEP), postulating the universality of the free fall, or equivalence
of the inertial and gravitational masses, was confirmed
in torsion balance experiments2 at the 2 × 10−13 level
for the normal matter. For the antimatter, the direct
observation of cold-trapped antihydrogen3 sets the limits
on the ratio between the inertial m and gravitational mg
masses of the antihydrogen, −65 < mg /m < 110, including systematic errors, at the 5% significance level3 . At
the same time, indirect limits have a long history and
are much stricter, see review4 for the results prior to
1991. At the moment, the most precise bounds on the
difference between the gravitational mass of the matter
and antimatter (to our knowledge) are obtained from the
¯ 0 system5
comparison of decay parameters of the K 0 − K
−9
(1.8 × 10 level with gravitational potential variations
and 1.9 × 10−14 with the LS potential) and from comparison of cyclotron frequencies6 of the p − p¯ system7
(10−6 level with LS potential). Equality of the inertial
masses for the considered (anti)particles is supported by
the CPT -symmetry tested with a much higher precision8 .
These and other indirect limits are, however, not absolute, but relative (between particles and antiparticles)
and for relatively low energies. There is, therefore, no
guarantee that, e.g., the strange matter (kaons) at any
energies, or normal matter and antimatter at high energies (several GeV and higher) will obey WEP. Even
though astrophysical tests of the Lorentz invariance9,10
can be, perhaps, used for the precise tests of the WEP
(mainly for electrons and protons), they rely on certain
models describing the high-energy sources and their dynamics. It is, therefore, desirable to obtain similar or
better constrains in a well-controlled experimental setup.
In this paper, we constrain possible deviations from
WEP for ultrarelativistic electrons and positrons (anti-
matter counterparts of electrons) based on the absence
of the vacuum Cherenkov radiation from 104.5 GeV
electrons and positrons at the LEP at CERN, and on
the absence of the photon decays for 340.5 GeV photons at the Tevatron accelerator at Fermilab. It is
known that the large Lorentz γ-factor for the ultrarelativistic particles reveals certain gravity and Lorentzviolating effects11–13 , and suppresses the ordinary electromagnetic interaction14 otherwise overwhelming gravitational forces15 . This nontrivial fact makes accelerator experiments suitable for the gravitational studies.
In addition, continuous collection of the accelerator data
makes it possible to study changes in the observables (or
exclusion regions in the parameter space) relative to the
periodic variations of the astrophysical potentials. This
gives one an opportunity to avoid assumptions on the
absolute values of the gravitational potentials5 . An additional advantage of the vacuum Cherenkov radiation
for the positron (electron) is its independence on the
gravitational properties of the electron (positron). We
also choose the electron and positron for our studies because of the absence of an additional internal structure
or flavor composition, avoiding possible speculations on,
e.g., undiscovered “strange”, “isotopic” or “hypercharge”
forces16,17 .
Dispersion relations.— Let us begin with a description
of the gravity effects on the high-energy processes. Gravitational field of the Earth (Sun or other distant massive
celestial objects) around the accelerator can be considered homogeneous and described by an isotropic metric
for a static weak field,
ds2 = H2 dt2 − H−2 (dx2 + dy 2 + dz 2 ) ,
(1)
where H2 = 1 + 2Φ, and Φ is the gravitational potential, defining the acceleration of free-falling bodies,
a = −∇Φ(x), taken at the Earth’s surface. Here and
after we work in natural units, c = ~ = 1. For a massive (anti-)particle of inertial mass m and gravitational
TUHEP-TH-151839
Further Investigation on Model-Independent Probe of Heavy Neutral Higgs Bosons at
the LHC Run 2
Yu-Ping Kuang1,2;1
Hong-Yu Ren1;2
Ling-Hao Xia1;3
1
arXiv:1506.08007v1 [hep-ph] 26 Jun 2015
2
Department of Physics, Tsinghua University, Beijing, 100084, China and
Center for High Energy Physics, Tsinghua University, Beijing, 100084, China
(Dated: June 29, 2015)
In our previous paper [1], we provided general effective Higgs interactions for the lightest
Higgs boson h (SM-like) and a heavier neutral Higgs boson H based on the effective Lagrangian
formulation up to the dim-6 interactions, and then proposed two sensitive processes for probing
H. We showed in several examples that the resonance peak of H and its dim-6 effective coupling
constants (ECC) can be detected at the LHC Run 2 with reasonable integrated luminosity. In this
paper, we further perform a more thorough study of the most sensitive process, pp → V H ∗ → V V V ,
on the information about the relations between the 1σ, 3σ, 5σ statistical significance and the
corresponding ranges of the Higgs ECC for an integrated luminosity of 100 fb−1 . These results
have two useful applications in the LHC Run 2: (A) realizing the experimental determination
of the ECC in the dim-6 interactions if H is found and, (B) obtaining the theoretical exclusion
bounds if H is not found. Some alternative processes sensitive for certain ranges of the ECC are
also analyzed.
PACS numbers: 14.80.Ec, 12.60.Fr, 12.15.-y
I.
INTRODUCTION
After the discovery of the 125 GeV Higgs in 2012 at
the CERN LHC [2], the ATLAS and CMS collaborations
have measured its couplings to other particles[3][4]. So
far, to the present experimental precision, they turn out
to be all consistent with the standard model (SM) predictions. However, it does not mean that the SM is the
final theory of fundamental interactions since it has several shortcomings, such as unnaturalness[5], triviality[6],
vacuum instability[7] and its lack of a suitable dark matter candidate. Searching for new physics beyond the SM
is still the main task in the TeV scale particle physics. So
far, there is no evidence of the well-known new physics
models such as supersymmetry, large extra dimensions,
etc.
We know that most new physics models contain several
Higgs bosons, and the lightest one may behave as (or very
close to) the SM Higgs boson, while the masses of other
heavy Higgs are usually in the few hundred GeV to 1 TeV
range. Therefore, the discovered 125 GeV Higgs boson
may actually be the lightest Higgs boson in a new physics
model. So that searching for a heavier Higgs boson may
be a feasible way to find the evidence of new physics.
Heavy Higgs bosons in several most popular models such
as the minimal supersymmetric extension of the standard model (MSSM) and the two-Higgs-doublet model
(2HDM) [8] were searched for during the LHC Run 1, but
no positive evidence has been found. Therefore, a modelindependent probe of the neutral heavy Higgs bosons is a
more efficient way of doing it. In our previous paper [1],
we provided general effective Higgs interactions for the
lightest Higgs boson h (SM-like) and a heavier neutral
Higgs boson H based on the effective Lagrangian formulation up to the dim-6 interactions, and then we proposed
two sensitive processes, namely the weak-boson scattering V V → V V (WBS) and pp → V H ∗ → V V V (VH∗ ),
where V = W, Z, for probing H. We showed in several
examples that the resonance peak of H and its dim-6
effective coupling constants (ECC) can be detected at
the LHC Run 2 with reasonable integrated luminosity.
Experimentally, the CMS collaboration performs a more
general search, which gives the exclusion limit for a neutral heavy Higgs boson with the SM couplings up to an
overall factor C ′ [9].
In this paper, as in Ref. [1], we consider an arbitrary new physics theory containing more than one Higgs
fields Φ1 , Φ2 , . . . without specifying the number of Φi
and their representations. Their interaction potential
V (Φ1 , Φ2 , . . .) may, in general, cause mixing between the
Higgs fields, and form a set of mass eigenstates. We denote the lightest mass eigenstate by Φh , and the second
lightest one by ΦH . The neutral Higgs bosons in Φh and
ΦH will be denoted by h and H, respectively. Here we
identify h with the discovered 125 GeV Higgs boson.
In the language of effective Lagrangian, we expand the
effective interactions up to the dim-6 terms. Since h is
identified with the discovered 125 GeV SM-like Higgs boson with vanishing dim-6 interactions. For H, the effective interactions can be expressed by
L = L(4) + L(6) .
(1)
Since ΦH is a mixture of the original Higgs Fields
Φ1 , Φ2 , . . ., the gauge coupling gH and vacuum expectation value (VEV) vH of H may be different from the
TUM-HEP 1003/15, TTP14-040, SFB/CPP-14-106, 15 January 2015
Non-relativistic high-energy physics: top production and dark matter annihilation
Martin Beneke
Physik Department T31, James-Franck-Straße 1, Technische Universit¨at M¨unchen, 85748 Garching, Germany
Matthias Steinhauser
arXiv:1506.07962v1 [hep-ph] 26 Jun 2015
Institut f¨ur Theoretische Teilchenphysik, Karlsruhe Institute of Technology (KIT), 76128 Karlsruhe, Germany
Abstract
Non-relativistic physics is often associated with atomic physics and low-energy phenomena of the strong interactions
between nuclei and quarks. In this review we cover three topics in contemporary high-energy physics at or close
to the TeV scale, where non-relativistic dynamics plays an important if not defining role. We first discuss in detail
the third-order corrections to top-quark pair production in electron-positron collisions in the threshold region, which
plays a major role at a future high-energy e+ e− collider. Threshold effects are also relevant in the production of heavy
particles in hadronic collisions, where in addition to the Coulomb force soft gluon radiation contributes to enhanced
quantum corrections. We review the joint resummation of non-relativistic and soft gluon effects for pair production of
top quarks and supersymmetric particles to next-to-next-to-leading logarithmic accuracy. The third topic deals with
pair annihilation of dark matter particles within the framework of the Minimal Supersymmetric Standard Model. Here
the electroweak Yukawa force generated by the exchange of gauge and Higgs bosons can cause large “Sommerfeld”
enhancements of the annihilation cross section in some parameter regions.
Keywords: top quark production, perturbative QCD, (P)NRQCD, ILC, LHC, dark matter
1. Introduction
Non-relativistic particle physics is often associated
with atomic physics, the interactions between nuclei,
and the low-energy phenomena of the strong interactions of the charm and bottom quarks, which form
quarkonium bound states. But non-relativistic dynamics also governs the interactions of weak-scale particles
such as the top quark or the hypothetical supersymmetric partners of the Standard Model (SM) particles,
when they are slowly moving. In this review we
cover three topics in contemporary high-energy physics
where the unique non-relativistic dynamics caused by
instantaneous, potential interactions plays an important
if not defining role: the production of top-quark pairs
in electron-positron collisions in the threshold region,
which can provide a measurement of the top-quark mass
with unchallenged precision and could be realized at
a future high-energy e+ e− collider. The hadronic pair
production of top quarks at Tevatron and LHC, or yet
unobserved heavy particles at the LHC, where in addition to the Coulomb force soft gluon radiation contributes to enhanced quantum corrections, which should be
summed. And finally, the pair annihilation of dark matter particles within the framework of the Minimal Supersymmetric Standard Model (MSSM). While the first
two processes are determined by the long-range colour
Coulomb force, the low-velocity annihilation of heavy
neutralinos in the MSSM can be dramatically enhanced
by the short-range electroweak Yukawa force generated
by the exchange of gauge and Higgs bosons.
We discuss the phenomena and results but also emphasize theoretical methods and techniques. Since there
is always a small relative velocity v in the problem, nonrelativistic systems involve (at least) the three scales m
(mass), mv (momentum), mv2 (energy), and it is appropriate to formulate effective Lagrangians to organize the
Upper bound of the N(1440) → N(939) + π decay width obtained from a three-flavor
parity doublet model
Hiroki Nishihara∗1 and Masayasu Harada†1
arXiv:1506.07956v1 [hep-ph] 26 Jun 2015
1
Department of Physics, Nagoya University, Nagoya 464-8602, Japan
(Dated: June 29, 2015)
We study masses and decay widths of positive and negative parity nucleons using a three-flavor
¯ ⊕ (3,
¯ 3)], [(3, 6) ⊕ (6, 3)],
parity doublet model, in which we introduce three representations, [(3, 3)
and [(8, 1) ⊕ (1, 8)] of the chiral U(3)L ×U(3)R symmetry. We find an extended version of the
Goldberger-Treiman relation among the mass differences and the coupling constants for pionic transitions. This relation leads to an upper bound for the decay width of N (1440) → N (939) + π
independently of the model parameters. We perform the numerical fitting of the model parameters
and derive several predictions, which can be tested in future experiments or lattice QCD analysis.
Furthermore, when we use the axial coupling of the excited nucleons obtained from lattice QCD analyses, we also find that the ground state nucleon N (939) consists of about 80% of [(3, 6) ⊕ (6, 3)]
component and about 20% of [(8, 1) ⊕ (1, 8)] component, and that the chiral invariant mass of
N (939) is about 250 MeV.
PACS numbers: 14.20.Gk, 12.39.Fe, 13.30.Eg, 14.20.Dh
I.
INTRODUCTION
There are many baryons in Nature [1], which are described by the flavor symmetry as in Table I. It is considered that the Quantum ChromoDynamics (QCD) describes their properties such as the masses and the interactions as well as the structure under the flavor symmetry. One of the most important features of QCD relevant
to the low-energy hadron physics is the chiral symmetry
and its spontaneous breaking. The spontaneous chiral
symmetry breaking generates mass differences between
the chiral partners, which lead to the mass difference between the parity partners, as well as the mixing among
different chiral representations. It is interesting to study
the role of the chiral symmetry breaking to determine
the properties and structures of baryons. In particular,
asking how much amount of the masses of baryons are
generated by the chiral symmetry breaking is an attractive question.
Two-flavor parity doublet model was proposed by
Table I: Data of spin 1/2 baryons in PDG [1].
JP
1/2+
1/2+
1/2−
1/2−
1/2+
N (939)
N (1440)
N (1535)
N (1650)
N (1710)
Octets
Λ(1116)
Λ(1600)
Λ(1670)
Λ(1800)
Λ(1810)
∗ [email protected][email protected]
Σ(1193) Ξ(1318)
Σ(1660) Ξ(1690)
Σ(1620) Ξ(?)
Σ(1750) Ξ(?)
Σ(1880) Ξ(?)
+
Ref. [2] in which the nucleon N (939) J P = 21
and
−
are introduced as
its parity partner N (1535) J P = 21
a doublet. This model includes a chiral invariant mass
denoted by m0 , which implies that the masses of N (939)
and N (1535) tend to m0 when the chiral symmetry is
restored and their mass splitting is given by the spontaneous chiral symmetry braking. This structure has been
studied by many works [2–10]. One of the key feature of
the structure is a relation among the pionic interactions
and mass difference between the chiral partners, which
is an extended Goldberger-Treiman relation. Using the
relation, one can obtain some information of m0 by comparing the model with experimental data, which seems
to prefer small value of m0 , m0 . 500 MeV (See e.g.
Refs. [6, 10]). On the other hand, parity doublet models
are applied to study the nuclear matter, where large value
of m0 (e.g. m0 & 800 MeV [11, 12], m0 & 500 MeV [13])
seems to be preferred.
References [7–9] gave the extension to the three-flavor
parity doublet model with the chiral U(3)L ×U(3)R symmetry. In two-flavor case, the axial charge gA of the nucleon is constrained as gA ≤ 1 as mentioned in Ref. [2].
However, as is well known, it is empirically determined
as gA = F + D = 1.268 ± 0.006, where F and D are
the axial coupling constants given in Ref. [14], and their
values are 0.475 ± 0.004 and 0.793 ± 0.005 [15]. To solve
this puzzle, the authors in Refs. [7–9] introduced three
parity-even and three parity-odd baryons which belong
¯ ⊕ (3,
¯ 3)], [(3, 6) ⊕ (6, 3)], and [(8, 1) ⊕ (1, 8)]
to [(3, 3)
representations of the chiral U(3)L ×U(3)R group, and
showed that the baryon mass spectra and the axial coupling constants are reasonably agree with experiments
when the number representations are suitably restricted.
In this paper, we make a general analysis using a threeflavor parity doublet model including all possible representations to fit available experimental data. We will
show that there exists an extended Goldberger-Treiman
Heavy quarkonia in strong magnetic field
Claudio Bonati∗
INFN - Sezione di Pisa, Largo Pontecorvo 3, I-56127 Pisa, Italy
Massimo D’Elia†
INFN - Sezione di Pisa, Largo Pontecorvo 3, I-56127 Pisa, Italy and
Dipartimento di Fisica dell’Universit`
a di Pisa, Largo Pontecorvo 3, I-56127 Pisa, Italy
Andrea Rucci‡
arXiv:1506.07890v1 [hep-ph] 25 Jun 2015
Dipartimento di Fisica dell’Universit`
a di Pisa, Largo Pontecorvo 3, I-56127 Pisa, Italy
(Dated: June 29, 2015)
We investigate the influence of a homogeneus and constant strong external magnetic field on
the heavy-meson spectrum. Quarkonium states c¯
c and b¯b are described within a non-relativistic
framework and by means of a suitable potential model based on the Cornell parametrization. In
particular, in this work we propose a model which takes into account the possible anisotropies
emerging at the level of the static quark-antiquark potential, as observed in recent lattice studies.
The investigation is perfomed both with and without taking into account the anisotropy of the static
potential, in order to better clarify its effects.
PACS numbers: 12.39.Jh 12.39.Pn 14.40.Pq 12.38.Gc
I.
INTRODUCTION
In recent times there has been a great interest regarding the physics of strongly interacting matter in the presence of strong external magnetic fields, i.e. such that1
eB ≃ m2π or larger (see, e.g., Refs. [1, 2] for recent reviews). This topic could be relevant to the study of some
dense astrophysical objects, like magnetars [3], and for
cosmology [4, 5]. However, the main interest was triggered by the fact that magnetic fields of this order of
magnitude can be created in a laboratory, in particular
in heavy-ion collisions [6–10], when two relativistic heavy
ions collide with a non-zero impact parameter, producing
a huge field in the collision region. For example, one can
reach |e|B ≃ 0.2 − 0.3 GeV2 in Pb+Pb collisions at the
Large Hadron Collider (LHC).
Such huge magnetic fields are produced in the very
early stages of the collision. It is still not clear how long
and to what extent they survive the thermalization process of the fireball created after the collision. Therefore,
while various theoretical investigations, based both on
model studies and on Lattice QCD simulations (LQCD),
have predicted many interesting phenomena affecting the
properties of strongly interacting matter in the presence
of strong magnetic backgrounds, it is still uncertain to
what extent such phenomena will be detectable in heavy
ion experiments.
In this perspective, effects regarding the physics of
∗ Electronic
address: [email protected]
address: [email protected]
‡ Electronic address: [email protected]
1 For reference, a magnetic field of the order of 1015 Tesla corresponds to eB ≃ 3.3 m2π ≈ 0.06 GeV 2 .
† Electronic
heavy flavors are of particular interest, since they are
more sensitive to the conditions taking place in the early
stages of the collision. Various studies have approached
the issue of quarkonia spectra and production rates in
the presence of magnetic backgrounds [11–17]. Many interesting phenomena have been predicted, including the
emergence of magnetic field induced mixings between different states and of production anisotropies with respect
to the collision plane.
The starting point for most of these investigations is
the coupling of the magnetic field with electric charges
and magnetic moments carried by the valence quarks.
However, the magnetic field is known to induce important modifications also at a non-perturbative level and in
the gluon sector: a natural question is whether that can
change the picture both quantitatively and qualitatively.
A very interesting phenomenon in this respect is the
magnetic field induced modification of the static quark¯ potential. This is a typical pure gauge
antiquark (QQ)
quantity (it is related to the expectation values of Wilson
loops) which represents the starting point for many approaches to the study of heavy quarkonia, tipically within
a non-relativistic approximation. The potential is usually
expressed in terms of the so called Cornell parametrization [18]:
V (r) = −
α
+ σr .
r
(1)
where α is the Coulomb coupling and σ the string tension.
The magnetic background field breaks rotational invariance, hence one may expect the emergence of anisotropies
in the potential, which is central otherwise. This has
been proposed in Ref. [19] and confirmed by a recent
lattice QCD study [20]: the potential gets steeper in
the directions transverse to the magnetic field and flatter in the longitudinal one. In particular one observes
arXiv:1506.08181v1 [nucl-ex] 26 Jun 2015
φ meson production in d+Au collisions at
√
sNN = 200 GeV
A. Adare,13 C. Aidala,42, 43 N.N. Ajitanand,62 Y. Akiba,56, 57 H. Al-Bataineh,50 J. Alexander,62 M. Alfred,23
A. Angerami,14 K. Aoki,31, 34, 56 N. Apadula,28, 63 Y. Aramaki,12, 56 H. Asano,34, 56 E.T. Atomssa,35 R. Averbeck,63
T.C. Awes,52 B. Azmoun,7 V. Babintsev,24 M. Bai,6 G. Baksay,19 L. Baksay,19 N.S. Bandara,42 B. Bannier,63
K.N. Barish,8 B. Bassalleck,49 A.T. Basye,1 S. Bathe,5, 8, 57 V. Baublis,55 C. Baumann,7, 44 A. Bazilevsky,7
M. Beaumier,8 S. Beckman,13 S. Belikov,7, ∗ R. Belmont,43, 67 R. Bennett,63 A. Berdnikov,59 Y. Berdnikov,59
J.H. Bhom,71 D.S. Blau,33 J.S. Bok,50, 71 K. Boyle,57, 63 M.L. Brooks,38 J. Bryslawskyj,5 H. Buesching,7
V. Bumazhnov,24 G. Bunce,7, 57 S. Butsyk,38 S. Campbell,14, 28, 63 A. Caringi,45 C.-H. Chen,57, 63 C.Y. Chi,14
M. Chiu,7 I.J. Choi,25, 71 J.B. Choi,10 R.K. Choudhury,4 P. Christiansen,40 T. Chujo,66 P. Chung,62 O. Chvala,8
V. Cianciolo,52 Z. Citron,63, 69 B.A. Cole,14 Z. Conesa del Valle,35 M. Connors,63 M. Csan´
ad,17 T. Cs¨org˝
o,70
T. Dahms,63 S. Dairaku,34, 56 I. Danchev,67 D. Danley,51 K. Das,20 A. Datta,42, 49 M.S. Daugherity,1 G. David,7
21
49
63
24
57, 63
7
M.K. Dayananda, K. DeBlasio, K. Dehmelt, A. Denisov, A. Deshpande,
E.J. Desmond,
K.V. Dharmawardane,50 O. Dietzsch,60 A. Dion,28, 63 P.B. Diss,41 J.H. Do,71 M. Donadelli,60 L. D’Orazio,41
O. Drapier,35 A. Drees,63 K.A. Drees,6 J.M. Durham,38, 63 A. Durum,24 D. Dutta,4 S. Edwards,20 Y.V. Efremenko,52
F. Ellinghaus,13 T. Engelmore,14 A. Enokizono,52, 56, 58 H. En’yo,56, 57 S. Esumi,66 B. Fadem,45 N. Feege,63
D.E. Fields,49 M. Finger,9 M. Finger, Jr.,9 F. Fleuret,35 S.L. Fokin,33 Z. Fraenkel,69, ∗ J.E. Frantz,51, 63 A. Franz,7
A.D. Frawley,20 K. Fujiwara,56 Y. Fukao,56 T. Fusayasu,47 C. Gal,63 P. Gallus,15 P. Garg,3 I. Garishvili,37, 64
H. Ge,63 F. Giordano,25 A. Glenn,37 H. Gong,63 M. Gonin,35 Y. Goto,56, 57 R. Granier de Cassagnac,35 N. Grau,2, 14
S.V. Greene,67 G. Grim,38 M. Grosse Perdekamp,25 T. Gunji,12 H.-˚
A. Gustafsson,40, ∗ T. Hachiya,56 J.S. Haggerty,7
K.I. Hahn,18 H. Hamagaki,12 J. Hamblen,64 H.F. Hamilton,1 R. Han,54 S.Y. Han,18 J. Hanks,14, 63
29
21
56, 58
S. Hasegawa, T.O.S. Haseler, K. Hashimoto,
E. Haslum,40 R. Hayano,12 X. He,21 M. Heffner,37
T.K. Hemmick,63 T. Hester,8 J.C. Hill,28 M. Hohlmann,19 R.S. Hollis,8 W. Holzmann,14 K. Homma,22 B. Hong,32
T. Horaguchi,22 D. Hornback,64 T. Hoshino,22 N. Hotvedt,28 J. Huang,7 S. Huang,67 T. Ichihara,56, 57
R. Ichimiya,56 Y. Ikeda,66 K. Imai,29, 34, 56 M. Inaba,66 A. Iordanova,8 D. Isenhower,1 M. Ishihara,56 M. Issah,67
D. Ivanishchev,55 Y. Iwanaga,22 B.V. Jacak,63 M. Jezghani,21 J. Jia,7, 62 X. Jiang,38 J. Jin,14 B.M. Johnson,7
T. Jones,1 K.S. Joo,46 D. Jouan,53 D.S. Jumper,1, 25 F. Kajihara,12 J. Kamin,63 S. Kanda,12 J.H. Kang,71
J. Kapustinsky,38 K. Karatsu,34, 56 M. Kasai,56, 58 D. Kawall,42, 57 M. Kawashima,56, 58 A.V. Kazantsev,33
T. Kempel,28 J.A. Key,49 V. Khachatryan,63 A. Khanzadeev,55 K.M. Kijima,22 J. Kikuchi,68 A. Kim,18 B.I. Kim,32
´ Kiss,17
C. Kim,32 D.J. Kim,30 E.-J. Kim,10 G.W. Kim,18 M. Kim,61 Y.-J. Kim,25 B. Kimelman,45 E. Kinney,13 A.
E. Kistenev,7 R. Kitamura,12 J. Klatsky,20 D. Kleinjan,8 P. Kline,63 T. Koblesky,13 L. Kochenda,55 B. Komkov,55
66
25
55, 59
15
14
38
56, 58
M. Konno, J. Koster, D. Kotov,
A. Kr´al, A. Kravitz, G.J. Kunde, K. Kurita,
M. Kurosawa,56, 57
Y. Kwon,71 G.S. Kyle,50 R. Lacey,62 Y.S. Lai,14 J.G. Lajoie,28 A. Lebedev,28 D.M. Lee,38 J. Lee,18 K.B. Lee,32
32
71
63
38
60
11
45
K.S. Lee, S Lee, S.H. Lee, M.J. Leitch, M.A.L. Leite, X. Li, P. Lichtenwalner, P. Liebing,57
S.H. Lim,71 L.A. Linden Levy,13 T. Liˇska,15 H. Liu,38 M.X. Liu,38 B. Love,67 D. Lynch,7 C.F. Maguire,67
Y.I. Makdisi,6 M. Makek,72 M.D. Malik,49 A. Manion,63 V.I. Manko,33 E. Mannel,7, 14 Y. Mao,54, 56 H. Masui,66
F. Matathias,14 M. McCumber,38, 63 P.L. McGaughey,38 D. McGlinchey,13, 20 C. McKinney,25 N. Means,63
A. Meles,50 M. Mendoza,8 B. Meredith,25 Y. Miake,66 T. Mibe,31 A.C. Mignerey,41 K. Miki,56, 66 A. Milov,7, 69
D.K. Mishra,4 J.T. Mitchell,7 S. Miyasaka,56, 65 S. Mizuno,56, 66 A.K. Mohanty,4 P. Montuenga,25 H.J. Moon,46
T. Moon,71 Y. Morino,12 A. Morreale,8 D.P. Morrison,7, † T.V. Moukhanova,33 T. Murakami,34, 56 J. Murata,56, 58
A. Mwai,62 S. Nagamiya,31, 56 K. Nagashima,22 J.L. Nagle,13, ‡ M. Naglis,69 M.I. Nagy,17, 70 I. Nakagawa,56, 57
H. Nakagomi,56, 66 Y. Nakamiya,22 K.R. Nakamura,34, 56 T. Nakamura,56 K. Nakano,56, 65 S. Nam,18 C. Nattrass,64
P.K. Netrakanti,4 J. Newby,37 M. Nguyen,63 M. Nihashi,22 T. Niida,66 S. Nishimura,12 R. Nouicer,7, 57 T. Novak,70
N. Novitzky,30, 63 A.S. Nyanin,33 C. Oakley,21 E. O’Brien,7 S.X. Oda,12 C.A. Ogilvie,28 M. Oka,66 K. Okada,57
Y. Onuki,56 J.D. Orjuela Koop,13 J.D. Osborn,43 A. Oskarsson,40 M. Ouchida,22, 56 K. Ozawa,12, 31 R. Pak,7
V. Pantuev,26, 63 V. Papavassiliou,50 I.H. Park,18 J.S. Park,61 S. Park,61 S.K. Park,32 W.J. Park,32 S.F. Pate,50
M. Patel,28 H. Pei,28 J.-C. Peng,25 H. Pereira,16 D.V. Perepelitsa,7 G.D.N. Perera,50 D.Yu. Peressounko,33
J. Perry,28 R. Petti,7, 63 C. Pinkenburg,7 R. Pinson,1 R.P. Pisani,7 M. Proissl,63 M.L. Purschke,7 H. Qu,21
J. Rak,30 B.J. Ramson,43 I. Ravinovich,69 K.F. Read,52, 64 S. Rembeczki,19 K. Reygers,44 D. Reynolds,62
V. Riabov,48, 55 Y. Riabov,55, 59 E. Richardson,41 T. Rinn,28 D. Roach,67 G. Roche,39, ∗ S.D. Rolnick,8 M. Rosati,28
C.A. Rosen,13 S.S.E. Rosendahl,40 Z. Rowan,5 J.G. Rubin,43 P. Ruˇziˇcka,27 B. Sahlmueller,44, 63 N. Saito,31
T. Sakaguchi,7 K. Sakashita,56, 65 H. Sako,29 V. Samsonov,48, 55 S. Sano,12, 68 M. Sarsour,21 S. Sato,29, 31 T. Sato,66
S. Sawada,31 B. Schaefer,67 B.K. Schmoll,64 K. Sedgwick,8 J. Seele,13 R. Seidl,25, 56, 57 A. Sen,64 R. Seto,8
3
44
Institut f¨
ur Kernphysik, University of Muenster, D-48149 Muenster, Germany
45
Muhlenberg College, Allentown, Pennsylvania 18104%–5586, USA
46
Myongji University, Yongin, Kyonggido 449%–728, Korea
47
Nagasaki Institute of Applied Science, Nagasaki-shi, Nagasaki 851%–0193, Japan
48
National Research Nuclear University, MEPhI, Moscow Engineering Physics Institute, Moscow, 115409, Russia
49
University of New Mexico, Albuquerque, New Mexico 87131, USA
50
New Mexico State University, Las Cruces, New Mexico 88003, USA
51
Department of Physics and Astronomy, Ohio University, Athens, Ohio 45701, USA
52
Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA
53
IPN-Orsay, Universite Paris Sud, CNRS-IN2P3, BP1, F-91406, Orsay, France
54
Peking University, Beijing 100871, P. R. China
55
PNPI, Petersburg Nuclear Physics Institute, Gatchina, Leningrad region, 188300, Russia
56
RIKEN Nishina Center for Accelerator-Based Science, Wako, Saitama 351%–0198, Japan
57
RIKEN BNL Research Center, Brookhaven National Laboratory, Upton, New York 11973%–5000, USA
58
Physics Department, Rikkyo University, 3%–34%–1 Nishi-Ikebukuro, Toshima, Tokyo 171%–8501, Japan
59
Saint Petersburg State Polytechnic University, St. Petersburg, 195251 Russia
60
Universidade de S˜
ao Paulo, Instituto de F´ısica, Caixa Postal 66318, S˜
ao Paulo CEP05315%–970, Brazil
61
Department of Physics and Astronomy, Seoul National University, Seoul 151%–742, Korea
62
Chemistry Department, Stony Brook University, SUNY, Stony Brook, New York 11794%–3400, USA
63
Department of Physics and Astronomy, Stony Brook University, SUNY, Stony Brook, New York 11794%–3800, USA
64
University of Tennessee, Knoxville, Tennessee 37996, USA
65
Department of Physics, Tokyo Institute of Technology, Oh-okayama, Meguro, Tokyo 152%–8551, Japan
66
Institute of Physics, University of Tsukuba, Tsukuba, Ibaraki 305, Japan
67
Vanderbilt University, Nashville, Tennessee 37235, USA
68
Waseda University, Advanced Research Institute for Science and
Engineering, 17 Kikui-cho, Shinjuku-ku, Tokyo 162%–0044, Japan
69
Weizmann Institute, Rehovot 76100, Israel
70
Institute for Particle and Nuclear Physics, Wigner Research Centre for Physics, Hungarian
Academy of Sciences (Wigner RCP, RMKI) H-1525 Budapest 114, POBox 49, Budapest, Hungary
71
Yonsei University, IPAP, Seoul 120-749, Korea
72
University of Zagreb, Faculty of Science, Department of Physics, Bijeniˇcka 32, HR-10002 Zagreb, Croatia
(Dated: June 29, 2015)
√
The PHENIX experiment has measured φ meson production in d+Au collisions at sN N =
200 GeV using the dimuon and dielectron decay channels. The φ meson is measured in the forward
(backward) d-going (Au-going) direction, 1.2 < y < 2.2 (−2.2 < y < −1.2) in the transversemomentum (pT ) range from 1–7 GeV/c, and at midrapidity |y| < 0.35 in the pT range below
7 GeV/c. The φ meson invariant yields and nuclear-modification factors as a function of pT , rapidity,
and centrality are reported. An enhancement of φ meson production is observed in the Au-going
direction, while suppression is seen in the d-going direction, and no modification is observed at
midrapidity relative to the yield in p+p collisions scaled by the number of binary collisions. Similar
behavior was previously observed for inclusive charged hadrons and open heavy flavor indicating
similar cold-nuclear-matter effects.
PACS numbers: 25.75.Dw
I.
INTRODUCTION
Collisions of deuterons with gold nuclei (d+Au) are of
significant interest in the study of the strongly coupled
Quark Gluon Plasma (QGP) produced at the Relativistic Heavy Ion Collider (RHIC) [1, 2]. The highly asymmetric collisions of a small projectile and a large target
nucleus provide a way to investigate the initial state of a
nucleus-nucleus collision experimentally, potentially disentangling effects due to QGP formation from the coldnuclear-matter effects. The latter include modification of
∗
†
‡
Deceased
PHENIX Co-Spokesperson: [email protected]
PHENIX Co-Spokesperson: [email protected]
the parton distribution functions (PDFs) in the nucleus
relative to those in the nucleon [3], initial-state energy
loss [4], and the so-called Cronin effect. The Cronin effect refers to the enhancement of high-pT particle production in p+A collisions relative to that in p+p collisions
scaled by the number of binary collisions and is often
attributed to multiple scattering of the incoming parton
inside the target nucleus [5–8]. In addition, results from
recent p(d)+A collisions at the Large Hadron Collider
and RHIC suggest that long-range correlations, either
present in the initial state or induced by the evolution of
the medium, play a important role even in these small
collision systems [7–15]. Detailed studies of particle production systematics in d+Au at RHIC may inform this
question [16–21].
PHENIX has measured the production of identified
June 29, 2015
Search for critical behavior of strongly interacting matter
at the CERN Super Proton Synchrotron
M. Gazdzicki1, 2 and P. Seyboth3, 2
1
arXiv:1506.08141v1 [nucl-ex] 26 Jun 2015
2
3
Goethe–University, Frankfurt, Germany
Jan Kochanowski University, Kielce, Poland
Max-Planck-Institut fuer Physik, Munich, Germany
Abstract
History, status and plans of the search for critical behavior of strongly interacting matter created
in nucleus-nucleus collisions at the CERN Super Proton Synchrotron is reviewed. In particular, it is
expected that the search should answer the question whether the critical point of strongly interacting
matter exists and, if it does, where it is located.
First, the search strategies are presented and a short introduction is given to expected fluctuation
signals and to the quantities used by experiments to detect them. The most important background
effects are also discussed.
Second, relevant experimental results are summarized and discussed. It is intriguing that both
the fluctuations of quantities integrated over the full experimental acceptance (event multiplicity and
transverse momentum) as well as the bin size dependence of the second factorial moment of pion and
proton multiplicities in medium-sized Si+Si collisions at 158A GeV/c suggest critical behaviour of the
created matter.
These results provide strong motivation for the ongoing systematic scan of the phase diagram by
the NA61/SHINE experiment at the SPS and the continuing search at the Brookhaven Relativistic
Hadron Collider.
PACS numbers: 12.40.-y, 12.40.Ee
Keywords: Critical point, nucleus-nucleus collisions, fluctuations
1