Download arXiv:1501.03089v1 [nucl

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Quantum vacuum thruster wikipedia , lookup

Scalar field theory wikipedia , lookup

Theory of everything wikipedia , lookup

Atomic nucleus wikipedia , lookup

Higgs boson wikipedia , lookup

Super-Kamiokande wikipedia , lookup

DESY wikipedia , lookup

Renormalization wikipedia , lookup

Theoretical and experimental justification for the Schrödinger equation wikipedia , lookup

Cross section (physics) wikipedia , lookup

Supersymmetry wikipedia , lookup

Antimatter wikipedia , lookup

Neutrino oscillation wikipedia , lookup

Renormalization group wikipedia , lookup

Quantum chromodynamics wikipedia , lookup

Peter Kalmus wikipedia , lookup

Higgs mechanism wikipedia , lookup

Monte Carlo methods for electron transport wikipedia , lookup

Nuclear structure wikipedia , lookup

Strangeness production wikipedia , lookup

Relativistic quantum mechanics wikipedia , lookup

Search for the Higgs boson wikipedia , lookup

Mathematical formulation of the Standard Model wikipedia , lookup

Lepton wikipedia , lookup

ALICE experiment wikipedia , lookup

Technicolor (physics) wikipedia , lookup

Compact Muon Solenoid wikipedia , lookup

Electron scattering wikipedia , lookup

ATLAS experiment wikipedia , lookup

Large Hadron Collider wikipedia , lookup

Minimal Supersymmetric Standard Model wikipedia , lookup

Weakly-interacting massive particles wikipedia , lookup

Elementary particle wikipedia , lookup

Grand Unified Theory wikipedia , lookup

Future Circular Collider wikipedia , lookup

Standard Model wikipedia , lookup

Transcript
Noise characterization of an atomic magnetometer at sub-millihertz frequencies
I. Mateos∗
Institut de Ci`encies de l’Espai (CSIC-IEEC), 08034 Barcelona, Spain
B. Patton1 , E. Zhivun, D. Budker2
arXiv:1501.03829v1 [physics.ins-det] 15 Jan 2015
Department of Physics, University of California, Berkeley, California 94720-7300
D. Wurm
Technische Universit¨at M¨unchen, 85748 Garching, Germany
J. Ramos-Castro3
Departament d’Enginyeria Electr`onica, Universitat Polit`ecnica de Catalunya,
08034 Barcelona, Spain
Abstract
Noise measurements have been carried out in the LISA bandwidth (0.1 mHz to 100 mHz) to characterize an alloptical atomic magnetometer based on nonlinear magneto-optical rotation. This was done in order to assess if the
technology can be used for space missions with demanding low-frequency requirements like the LISA concept. Magnetometry for low-frequency applications is usually limited by 1/ f noise and thermal drifts, which become the dominant contributions at sub-millihertz frequencies. Magnetic field measurements with atomic magnetometers are not
immune to low-frequency fluctuations and significant excess noise may arise due to external elements, such as temperature fluctuations or intrinsic noise in the electronics. In addition, low-frequency drifts in the applied magnetic field
have been identified in order to distinguish their noise contribution from that of the sensor. We have found the technology suitable for LISA in terms of sensitivity, although further work must be done to characterize the low-frequency
noise in a miniaturized setup suitable for space missions.
Keywords: eLISA, LISA Pathfinder, atomic magnetometer, low-frequency noise
1. Introduction
The evolved Laser Interferometer Space Antenna
(eLISA) is a mission concept proposed as a large (Lclass) space mission of the European Space Agency
(ESA) designed to detect low-frequency gravitational
radiation. It will be formed by three drag-free spacecraft
in triangular configuration with one-million-kilometer
∗ Corresponding
author
Email address: [email protected] (I. Mateos∗ )
1 Also at Physik-Department, Technische Universit¨
at M¨unchen,
85748 Garching, Germany
2 Also at Lawrence Berkeley National Laboratory, Berkeley, California 94720 and Helmholtz Institute, Johannes Gutenberg University,
55099 Mainz, Germany
3 Also at Institut d’Estudis Espacials de Catalunya, 08034,
Barcelona, Spain
Preprint submitted to Sensors and Actuators A: Physical
arms, with each spacecraft containing one or two “freefalling” macroscopic bodies called test masses (TMs)
[1]. Gravitational wave (GW) detection requires interferometry measurement at the picometer level between
two test masses along an arm due to the tidal deformation of spacetime caused by GWs. For this reason, the environment where such bodies will be located
must be very quiet (in relation to disturbances exerting
forces in the bodies), otherwise the motion provoked
by the different noise sources perturbing the free floating bodies would conceal the GW signal. The LISA
top-level requirement in terms of acceleration noise
1/2
(m s−2 Hz−1/2 ) in the frequency band between
S δa,LISA
0.1 mHz ≤ ω/2π ≤ 100 mHz is plotted in Fig. 1. At
frequencies below 1 mHz, the √
noise is dominated by the
residual acceleration noise of 2·3 fm s−2 Hz−1/2 caused
January 19, 2015
Antineutrino induced Λ(1405) production off the proton
Xiu-Lei Ren,1, 2, ∗ E. Oset,2, † L. Alvarez-Ruso,3, ‡ and M. J. Vicente Vacas2, §
1
arXiv:1501.04073v1 [hep-ph] 16 Jan 2015
School of Physics and Nuclear Energy Engineering & International Research Center
for Nuclei and Particles in the Cosmos, Beihang University, Beijing 100191, China
2
Departamento de F´ısica Te´orica and IFIC, Centro Mixto Universidad de Valencia-CSIC,
Institutos de Investigaci´on de Paterna, Apartado 22085, 46071 Valencia, Spain
3
Instituto de F´ısica Corpuscular (IFIC), Centro Mixto Universidad de Valencia-CSIC,
Institutos de Investigaci´on de Paterna, Apartado 22085, 46071 Valencia, Spain
We have studied the strangeness changing antineutrino induced reactions ν¯l p → l+ φB, with φB = K − p,
¯ 0 n, π 0 Λ, π 0 Σ0 , ηΛ, ηΣ0 , π + Σ− , π − Σ+ , K + Ξ− and K 0 Ξ0 , using a chiral unitary approach. These ten
K
coupled channels are allowed to interact strongly, using a kernel derived from the chiral Lagrangians. This
interaction generates two Λ(1405) poles, leading to a clear single peak in the πΣ invariant mass distributions.
At backward scattering angles in the center of mass frame, ν¯µ p → µ+ π 0 Σ0 is dominated by the Λ(1405) state
at around 1420 MeV while the lighter state becomes relevant as the angle decreases, leading to an asymmetric
line shape. In addition, there are substantial differences in the shape of πΣ invariant mass distributions for the
three charge channels. If observed, these differences would provide valuable information on a claimed isospin
I = 1, strangeness S = −1 baryonic state around 1400 MeV. Integrated cross sections have been obtained for
¯ channels, investigating the impact of unitarization in the results. The number of events with
the πΣ and KN
Λ(1405) excitation in ν¯µ p collisions in the recent antineutrino run at the MINERνA experiment has also been
obtained. We find that this reaction channel is relevant enough to be investigated experimentally and to be taken
into account in the simulation models of future experiments with antineutrino beams.
I.
INTRODUCTION
The Λ(1405) resonance is a cornerstone in hadron physics, challenging the standard view of baryons made of three quarks.
Long ago it was already suggested that the Λ(1405) could be a kind of molecular state arising from the interaction of the πΣ and
¯ channels [1, 2]. This view has been recurrent [3], but only after the advent of unitary chiral perturbation theory (UChPT) it
KN
has taken a more assertive tone [4–9]. In this framework, a kernel (potential) derived from the chiral Lagrangians is the input into
the Bethe-Salpeter equation in coupled channels. Sometimes the interaction is strong enough to generate poles, denominated as
dynamically generated states, which can be interpreted as hadronic molecules with components on the different channels (see
Ref. [10] for a review).
It came as a surprise that UChPT predicts two Λ(1405) states [6], studied in detail in Ref. [8]. Two poles appear, one
around 1420 MeV with a width of about 40 MeV and another one around 1385 MeV with a larger width of about 150 MeV.
These findings have been reconfirmed in more recent studies with potentials that include higher order terms of the chiral Lagrangians [11–17]. From the experimental perspective, the old experiments [18, 19] produced πΣ invariant mass distributions
where a single Λ(1405) peak is seen around 1405 MeV. According to Ref. [20], this single peak results from the overlap of the
two pole contributions. It has also been suggested that reactions induced by K − p pairs show a peak around 1420 MeV because
¯ , while the one at 1385 MeV does it more strongly to πΣ. This would be the case
the pole at 1420 MeV couples mostly to KN
of K − p → γπΣ [21] and K − p → π 0 π 0 Σ0 . The latter one, measured at Crystal Ball [22] and analyzed in Ref. [23] confirmed
the existence of the state at 1420 MeV. Another reaction that has proved its existence is K − d → nπΣ [24], which was studied
in Ref. [25]. The issues raised in Ref. [26] were addressed in detail in Ref. [27] reconfirming the findings of Ref. [25].
It is somewhat surprising that the two poles emerge in the theory even when only data on K − p scattering and K − p atoms [28],
which are above the Λ(1405) pole masses, are fitted. Nevertheless, it is clear that the best information on the Λ(1405) properties
should come from processes where the Λ(1405) is produced close to its pole masses. In this sense, the abundant Λ(1405)
photoproduction data obtained by CLAS with the γp → K + π + Σ− , K + π 0 Σ0 , K + π − Σ+ reactions [29] add much information
¯ channels
to the earlier data of Ref. [30], bringing new light into the subject. A fit to these data imposing unitarity in the πΣ, KN
and allowing only small variations in the kernel of the chiral Lagrangians [31, 32] has reconfirmed the existence of the two poles,
in agreement with the UChPT predictions. The wide range of energies investigated and the simultaneous measurement of the
three πΣ charged channels were the key to the solutions found in Refs. [31, 32] and, more recently, in Ref. [33].
Studies of p p → p K + Λ(1405) performed at ANKE show again a superposition of the contributions from the two poles [34],
and can be explained with the theoretical framework of UChPT [35]. More recent measurements [36, 37] show the Λ(1405) peak
∗
†
‡
§
E-mail:
E-mail:
E-mail:
E-mail:
[email protected]
[email protected]
[email protected]
[email protected]
Study of B → K0∗ (1430)K (∗) decays in QCD Factorization Approach
Ying Li∗ , Hong-Yan Zhang, Ye Xing, Zuo-Hong Li
Department of Physics, Yantai University, Yantai 264005, China
Cai-Dian L¨
u
arXiv:1501.03865v1 [hep-ph] 16 Jan 2015
Institute of High Energy Physics and Theoretical Physics Center for Science Facilities,CAS, Beijing 100049, China
(Dated: January 19, 2015)
Within the QCD factorization approach, we calculate the branching fractions and CP asymmetry parameters of 12 B → K0∗ (1430)K (∗) decay modes under the assumption that the scalar meson
K0∗ (1430) is the first excited state or the lowest lying ground state in the quark model. We find
that the decay modes with the scalar meson emitted, have large branching fractions due to the
K∗
enhancement of large chiral factor rχ 0 . The branching fractions of decays with the vector meson
∗
emitted, become much smaller owing to the smaller factor rχK . Moreover, the annihilation type
diagram will induce large uncertainties because of the extra free parameters dealing with the endpoint singularity. For the pure annihilation type decays, our predictions are smaller than that from
PQCD approach by 2-3 orders of magnitudes. These results will be tested by the ongoing LHCb
experiment, forthcoming Belle-II experiment and the proposing circular electron-positron collider.
I.
INTRODUCTION
Although the quark model has made great success in
describing most of the hadronic states, the lowest lying
scalar mesons are too light to fit in the quark model.
Two possible scenarios have been proposed on the basis of whether the scalars lower than 1 GeV belong to
the four-quark states or the classical two quark states.
They are controversial for decades [1]. In scenario-1
(S1), the scalars such as κ(800), a0 (980) and f0 (980) are
naively seen as the lowest lying q q¯ states, and K0∗ (1430),
a0 (1450), and f0 (1500) are the first excited states, correspondingly. In contrast, K0∗ (1430), a0 (1450), f0 (1500)
are treated as the the q q¯ ground states and their first
excited states are about (2.0 ∼ 2.3) GeV in scenario-2
(S2), in which the lightest scalar mesons are the fourquark bound states.
In the past decade, many B decay modes involving
scalar mesons have been reported by BaBar, Belle and
large hadron collider-b (LHCb) experiments. This may
provide a unique feature to distinguish these two scenarios by the B meson tag [2]. It is thus hoped that the
combination of the precise experimental measurements
and the accurate theoretical predictions might provide
us valuable information on the nature of scalar mesons.
To achieve this goal, some hadronic Bq (q = u, d, s) decays to scalar mesons have been studied in detail in the
framework of the QCD factorization (QCDF) [1, 3, 4]
and the perturbative QCD approach (PQCD) [5, 6]. Using the QCDF approach, H-Y Cheng et.al had calculated
the branching fractions and direct CP asymmetries of
most decay modes in refs. [3, 4], such as B → f0 K,
B → K0∗ (1430)φ(ρ) and B → K0∗ (1430)π, where most
results accommodate the data with large uncertainties.
∗ Email:
[email protected]
Very recently, the decays B → K0∗ (1430)K (∗) dominated
by b → d penguin operators, have been calculated within
the PQCD approach [6], however they have not been
touched in QCDF frame. To complete, we therefore shall
calculate the branching fractions of B → K0∗ (1430)K (∗)
decays in QCDF, as well as their CP asymmetry parameters. In the experimental side, these observables are
too small to be measured at current experiments, but
they are hopeful to be measured in the future experiments such as the updated LHCb experiment, the high
luminosity Belle-II experiment and the proposing high
energy circular electron-positron collider.
The paper is organized as follows. In Section 2, we
give the analytic formula including the effective Hamiltonian, the form factor and all corrections to the amplitudes. Presentation of results and discussions are given
in Section 3. At last, we summarize this work in Section
4.
II.
ANALYTIC FORMULA
In this section, we shall start from the weak effective
Hamiltonian responsible for B → K0∗ (1430)K (∗) decays.
In the standard model, it could be written as [7]
GF
∗
(C1 O1u + C2 O2u )
Heff = √ Vub Vud
2
10
X
∗
Ci Oi + C7γ O7γ + C8g O8g +h.c. (1)
− Vtb Vtd
i=3
The explicit form of the operators Oi and the corresponding Wilson Coefficients Ci at different scale µ could
be found in ref.[7]. Vu(t)b and Vu(t)d are the CabibboKabayashi-Maskawa (CKM) matrix elements. Note that
O1,2 are tree operators and others O3−10,7γ,8g are penguin ones.
EUROPEAN ORGANISATION FOR NUCLEAR RESEARCH (CERN)
CERN-PH-EP-2014-228
arXiv:1501.04020v1 [hep-ex] 16 Jan 2015
Submitted to: Physics Letters B
Search for pair-produced long-lived neutral particles √
decaying in
the ATLAS hadronic calorimeter in pp collisions at s = 8 TeV
The ATLAS Collaboration
Abstract
The ATLAS detector at the Large Hadron Collider at CERN is used to search for the decay of
a scalar boson to a pair of long-lived particles, neutral under the Standard Model gauge group, in
√
20.3 fb−1 of data collected in proton–proton collisions at s = 8 TeV. This search is sensitive to longlived particles that decay to Standard Model particles producing jets at the outer edge of the ATLAS
electromagnetic calorimeter or inside the hadronic calorimeter. No significant excess of events is
observed. Limits are reported on the product of the scalar boson production cross section times
branching ratio into long-lived neutral particles as a function of the proper lifetime of the particles.
Limits are reported for boson masses from 100 GeV to 900 GeV, and a long-lived neutral particle
mass from 10 GeV to 150 GeV.
c 2015 CERN for the benefit of the ATLAS Collaboration.
Reproduction of this article or parts of it is allowed as specified in the CC-BY-3.0 license.
Three-loop Correction to the Instanton Density for the Double
Well Potential
M.A. Escobar-Ruiz1 ,∗ E. Shuryak2 ,† and A.V. Turbiner1,2‡
1
Instituto de Ciencias Nucleares, Universidad Nacional Aut´onoma de M´exico,
Apartado Postal 70-543, 04510 M´exico, D.F., M´exico and
arXiv:1501.03993v1 [hep-th] 16 Jan 2015
2
Department of Physics and Astronomy,
Stony Brook University, Stony Brook, NY 11794-3800, USA
(Dated: January 15, 2015)
Abstract
This paper deals with quantum fluctuations near the classical instanton configuration. Feynman
diagrams in the instanton background are used for the calculation of the tunneling amplitude (the
instanton density) in the three-loop order for quartic double-well potential. The result for the threeloop contribution coincides in five significant figures with one given long ago by J. Zinn-Justin.
Unlike the two-loop contribution where all involved Feynman integrals are rational numbers, in
the three-loop case Feynman diagrams can contain irrational contributions.
∗
Electronic address: [email protected]
†
Electronic address: [email protected]
‡
Electronic address: [email protected], [email protected]
1
January 19, 2015
1:21
WSPC Proceedings - 9.75in x 6.5in
otqgp
1
Origin of Temperature of Quark-Gluon Plasma in Heavy Ion Collisions
arXiv:1501.03970v1 [nucl-th] 16 Jan 2015
Xiao-Ming Xu∗
Department of Physics, Shanghai University,
Baoshan, Shanghai 200444, China
∗ E-mail: [email protected]
Initially produced quark-gluon matter at RHIC and LHC does not have a temperature. A
quark-gluon plasma has a high temperature. From this quark-gluon matter to the quarkgluon plasma is the early thermalization or the rapid creation of temperature. Elastic
three-parton scattering plays a key role in the process. The temperature originates from
the two-parton scattering, the three-parton scattering, the four-parton scattering and so
forth in quark-gluon matter.
Keywords: Origin of temperature; Early thermalization; Elastic three-parton scattering;
Transport equation, Quark-gluon matter.
Temperature is common in daily life, and people believe that temperature exists
no matter how cold matter is. In gas and liquid temperature reflects the existence
of thermal states where all particles randomly move and obey isotropic momentum
distributions. The average momentum that a massless particle possesses is proportional to temperature. When temperature equals zero, the average momentum is
zero, and vice versa. However, there is one system where the average momentum is
large but particles do not obey isotropic momentum distributions, i.e., the system
does not have temperature. The system is quark-gluon matter that was produced in
initial gold-gold nuclear collisions performed at the Brookhaven National Laboratory Relativistic Heavy Ion Collider (RHIC) in the last fourteen years. In the parton
system the parton distribution functions in the direction of the incoming nucleus
beam are much larger than those in the direction perpendicular to the beam direction. The distributions are anisotropic in momentum space, and do not correspond
to a thermal state. Now we must wonder if this quark-gluon matter always has no
temperature. The answer is no, and this matter is exactly a place for exploring the
origin of temperature.
The reaction gg → gg was considered by Shuryak 1 to cause gluon matter to
thermalize. He assumed that every gluon suffers from scattering once in the same
period of time. In order to guarantee validity of the assumption, the cross section
for the gluon-gluon scattering must be very large (empirically be larger than 40
mb). This is not achievable! Even though a thermalization time of the order of 0.3
fm/c was obtained, the assumption makes the thermalization time unreliable. To
explore the kinetic equilibration of gluon momenta, inelastic scattering gg → ggg
was taken into account in Refs. 2,3 . gg → gg and gg → ggg lead to a gluon-matter
thermalization time larger than 1 fm/c. Overpopulation is found to be a key feature
page 1
Possible observables for chiral electric separation effect in Cu + Au collisions
Guo-Liang Ma1, ∗ and Xu-Guang Huang2, †
arXiv:1501.03903v1 [nucl-th] 16 Jan 2015
2
1
Shanghai Institute of Applied Physics, Chinese Academy of Sciences, Shanghai 201800, China
Physics Department and Center for Particle Physics and Field Theory, Fudan University, Shanghai 200433, China
The quark-gluon plasma (QGP) generated in relativistic heavy-ion collisions could be locally P- and CP-odd.
In P- and CP-odd QGP, the electric field may induce a chiral current which is called chiral electric separation
effect (CESE). We propose two possible observables for CESE in Cu + Au collisions: The first one is the
correlation ζαβ = hcos[2(φα + φβ − 2ΨRP )]i; the second one is the charge-dependent event-plane angle Ψq2 with
q = ± being charge. Nonzero ∆ζ = ζopp − ζsame and ∆Ψ = hΨ+2 − Ψ−2 i may signal the CESE in Cu + Au collisions.
Within a multiphase transport (AMPT) model, we study how the final state interaction affects these observables.
We find that the correlation γαβ = hcos(φα + φβ − ΨRP )i is sensitive to the out-of-plane charge separation caused
by chiral magnetic effect (CME) and the in-plane charge separation caused by the in-plane electric field but not
sensitive to CESE. On the other hand, ∆ζ and ∆Ψ are sensitive to the CESE. Therefore, we suggest the future
experiments to measure the above observables in Cu+Au collisions in order to disentangle different chiral and
charge separation mechanisms.
I.
INTRODUCTION
Relativistic heavy-ion collisions generate not only extremely hot quark-gluon plasma (QGP) but also extremely
large magnetic fields due to the fast motion of the colliding
ions. The recent detail calculations revealed that the maximum magnetic fields in Au + Au collisions at RHIC can reach
5m2π ∼ 1018 Gauss while in Pb + Pb collisions at LHC they can
reach 60m2π ∼ 1019 Gauss [1–5]. Under such large magnetic
fields, some novel quantum phenomena can possibly happen.
The most intriguing ones are the so-called chiral magnetic effect (CME) [6–10] and chiral separation effect [11, 12]: They
occur in P- and CP-odd regions in QGP and can result in
charge and chirality separations along the B direction. The recent experimental measurements about the charge azimuthal
correlation [13],
γαβ = hcos(φα + φβ − 2ΨRP )i,
(1)
where φα and φβ are the emission azimuthal angles of particles with charges α and β and ΨRP is the reaction plane angle,
showed some consistent features with the expectation of CME
at both RHIC and LHC energies [14–16]. Another important
experimental test of CME and CSE is the observation of the
charge asymmetry in the elliptic flow of pions [17, 18] that
is consistent with the expectation of a chiral magnetic wave
(CMW) [19, 20] — a collective mode arising due to the interplay between CME and CSE in the presence of the magnetic
field. However, one must notice that there are other backgrounds that contribute to the experimental observables, see
Refs. [21–29] for discussions.
Heavy-ion collisions can also generate strong electric fields
due to the event-by-event fluctuation of the proton positions
in the ions [3, 4] or due to asymmetric colliding geometry (for
example, in Cu + Au collision [30–32]). It was proposed that
the electric fields may also induce chiral current and chiral
∗
[email protected][email protected]
separation in QGP, which is called chiral electric separation
effect (CESE) [33–36]. It was also proposed that the Cu+Au
collisions may provide us a good chance to detect CESE because there is a strong E field directing from Au to Cu due to
the charge asymmetry between Au and Cu nuclei [33]. Recently, the electromagnetic fields in Cu + Au collisions were
studied in details, and it was found that the large in-plane
charge separation effect (in-plane CSE) induced by the strong
in-plane electric fields could strongly suppress or even reverse
signs of γαβ [30]. We in this paper will propose two observables that are designed solely for the detection of CESE. In
addition, it is crucial to take the final state interactions into
account for any model calculations in order to link the initial
anomalous transports to the experimental data since heavy-ion
collisions undergo complicated dynamical evolutions which
involve many final interactions. We study the effects of the final state interactions by using a multi-phase transport (AMPT)
model which successfully describes the main evolution stages
of heavy-ion collisions. By introducing appropriate initial
dipolar or quadrupole charge distributions, the AMPT model
can successfully describe both the charge azimuthal correlation γαβ [37, 38] and the charge asymmetry of the pion elliptic
flow [39] in Au + Au collisions at the top RHIC energy. In
this work, we introduce different kinds of initial charge separations which are used to mimic different initial chiral effects into the initial condition of the AMPT model, and predict some observables which can be used to test whether these
effects can be observed in the final state of Cu+Au collisions
√
at sNN = 200 GeV.
The paper is organized as follows. In Sec. II we setup the
numerical simulations, in Sec. III we show our main results.
Finally we summarize and discuss in Sec. IV. Throughout this
paper, we use the natural units ~ = kB = c = 1.
arXiv:1501.03492v1 [astro-ph.CO] 14 Jan 2015
EPJ Web of Conferences will be set by the publisher
DOI: will be set by the publisher
c Owned by the authors, published by EDP Sciences, 2015
Direct Dark Matter Search with XENON100
S.E.A. Orrigo1 , a , b
on behalf of the XENON Collaboration
1
Department of Physics, University of Coimbra, Coimbra, Portugal
Abstract. The XENON100 experiment is the second phase of the XENON program for
the direct detection of the dark matter in the universe. The XENON100 detector is a
two-phase Time Projection Chamber filled with 161 kg of ultra pure liquid xenon. The
results from 224.6 live days of dark matter search with XENON100 are presented. No
evidence for dark matter in the form of WIMPs is found, excluding spin-independent
WIMP-nucleon scattering cross sections above 2 × 10−45 cm2 for a 55 GeV/c2 WIMP
at 90% confidence level (C.L.). The most stringent limit is established on the spindependent WIMP-neutron interaction for WIMP masses above 6 GeV/c2 , with a minimum cross section of 3.5 × 10−40 cm2 (90% C.L.) for a 45 GeV/c2 WIMP. The same
dataset is used to search for axions and axion-like-particles. The best limits to date are
set on the axion-electron coupling constant for solar axions, gAe < 7.7 × 10−12 (90%
C.L.), and for axion-like-particles, gAe < 1 × 10−12 (90% C.L.) for masses between 5 and
10 keV/c2 .
1 Introduction
Non-Barionic dark matter constitutes the 26.8 % of the total energy and the 84.5% of the total matter
of the known universe. The nature of the dark matter is among the fundamental open questions in
modern physics. One of the most favorite dark matter candidates are the Weakly Interacting Massive
Particle (WIMPs), cold and not-charged exotic relics of the Big Bang that interact with the ordinary
matter only through gravity and the weak interaction. Axions and Axion-Like Particles (ALPs) are
other well-motivated candidates for cold dark matter, introduced by many extensions of the Standard
Model of particle physics.
The XENON program aims at the direct detection of the dark matter in the universe using dualphase Time Projection Chambers (TPCs) of increasing sensitivity [1–3] filled with ultra pure liquid
xenon (LXe). All the XENON experiments are located underground at the Laboratori Nazionali del
Gran Sasso (LNGS) in Italy. The XENON100 experiment [2], having a total mass of 161 kg of LXe,
is the second phase of the program and has set the most stringent limits on the WIMP-nucleon spinindependent [4] and spin-dependent [5] cross sections and on the axion-electron coupling [6]. The
successor, XENON1T [3, 7], is a ton-scale TPC that will be commissioned in 2015 and it is expected
to improve the sensitivity by two orders of magnitude.
a Corresponding author e-mail: [email protected]
b Present address: IFIC, CSIC-Universidad de Valencia, E-46071 Valencia, Spain
Relativistic description of nuclear matrix elements in neutrinoless double-β decay
L. S. Song,1 J. M. Yao,2, 3 P. Ring,4, 1 and J. Meng1, 5, 6
arXiv:1407.1368v2 [nucl-th] 15 Jan 2015
1
State Key Laboratory of Nuclear Physics and Technology,
School of Physics, Peking University, Beijing 100871, China
2
Department of Physics, Tohoku University, Sendai 980-8578, Japan
3
School of Physical Science and Technology, Southwest University, Chongqing 400715, China
4
Physik Department, Technische Universit¨
at M¨
unchen, D-85748 Garching, Germany
5
School of Physics and Nuclear Energy Engineering, Beihang University, Beijing 100191, China
6
Department of Physics, University of Stellenbosch, Stellenbosch 7602, South Africa
Background: Neutrinoless double-β (0νββ) decay is related to many fundamental concepts in nuclear and
particle physics beyond the standard model. Currently there are many experiments searching for this weak
process. An accurate knowledge of the nuclear matrix element for the 0νββ decay is essential for determining the
effective neutrino mass once this process is eventually measured.
Purpose: We report the first full relativistic description of the 0νββ decay matrix element based on a state-ofthe-art nuclear structure model.
Methods: We adopt the full relativistic transition operators which are derived with the charge-changing nucleonic
currents composed of the vector coupling, axial-vector coupling, pseudoscalar coupling, and weak-magnetism
coupling terms. The wave functions for the initial and final nuclei are determined by the multireference covariant
density functional theory (MR-CDFT) based on the point-coupling functional PC-PK1. Correlations beyond the
mean field are introduced by configuration mixing of both angular momentum and particle number projected
quadrupole deformed mean-field wave functions.
Results: The low-energy spectra and electric quadrupole transitions in 150 Nd and its daughter nucleus 150 Sm
+
are well reproduced by the MR-CDFT calculations. The 0νββ decay matrix elements for both the 0+
1 → 01
+
+
150
and 01 → 02 decays of
Nd are evaluated. The effects of particle number projection, static and dynamic
deformations, and the full relativistic structure of the transition operators on the matrix elements are studied in
detail.
+
Conclusions: The resulting 0νββ decay matrix element for the 0+
1 → 01 transition is 5.60, which gives the most
optimistic prediction for the next generation of experiments searching for the 0νββ decay in 150 Nd.
PACS numbers: 21.60.Jz, 24.10.Jv, 23.40.Bw, 23.40.Hc
I.
INTRODUCTION
Double-β (ββ) decay is a second-order weak process
in which a nucleus decays to the neighboring nucleus
by emitting two electrons and, usually, other light particles [1],
(A, Z) → (A, Z + 2) + 2e− + light particles.
(1)
Owing to the huge β decay background, events of this
process could, so far, only be recorded in some eveneven nuclei, where the β decay is energetically forbidden. There are several ββ decay modes including the
two-neutrino double-β (2νββ) decay mode,
(A, Z) → (A, Z + 2) + 2e− + 2¯
νe ,
(2)
and the neutrinoless (0νββ) decay mode,
(A, Z) → (A, Z + 2) + 2e− .
(3)
The 2νββ mode is allowed in the standard model (SM),
while the existence of the 0νββ decay would require to
go beyond the SM. Evidence for the 0νββ decay would
be a proof that neutrinos with definite masses are Majorana particles and that neutrino masses have an origin
beyond the SM [2]. This conclusion is independent of the
underlying mechanism governing the weak process [3].
So far, half-lives of the 2νββ decay have been measured
in 11 isotopes, which are of the order of 1018−24 y [4, 5].
However, the 0νββ event has never been seen. Only limits of the half-lives can be drawn from current experi0ν
> 1021−25 y. Searches for the
ments, which are T1/2
0νββ signals in the ββ candidates are ongoing or proposed in a number of laboratories around the world (see
Refs. [1, 6, 7] for comprehensive reviews).
0ν
Limits of the half-lives T1/2
drawn from experiments
provide stringent limits on the parameters associated
with the assumed underlying mechanism. Assuming a
long-range interaction based on the exchange of a light
Majorana neutrino between two weak interaction vertices
and restricting the currents to the standard (V −A) form,
the part that is proportional to the neutrino mass will be
picked out from the neutrino propagator by the same helicity of the coupled leptonic currents [1, 8]. Therefore, in
this case the associated parameter is the effective Majorana neutrino mass. This is called the mass mechanism.
Being regarded as the minimal extension of the SM, the
mass mechanism is the most popular assumption in current existing theoretical calculations.
Using the mass mechanism, one expects that the 0νββ
observation, combined with the results of neutrino oscillation experiments, will allow to obtain important infor-
arXiv:1501.04093v1 [hep-ph] 16 Jan 2015
Analytical Theory of Neutrino Oscillations in Matter with CP violation
Mikkel B. Johnson
Los Alamos National Laboratory, Los Alamos, NM 87545
Ernest M. Henley
Department of Physics, University of Washington, Seattle, WA 98195
Leonard S. Kisslinger
Department of Physics, Carnegie-Mellon University, Pittsburgh, PA 15213
We develop an exact analytical formulation of neutrino oscillations in matter within the framework
of the Standard Neutrino Model assuming 3 Dirac Neutrinos. Our Hamiltonian formulation, which
includes CP violation, leads to expressions for the partial oscillation probabilities that are linear
combinations of spherical Bessel functions in the eigenvalue differences. The coefficients of these
Bessel functions are polynomials in the neutrino CKM matrix elements, the neutrino mass differences
squared, the strength of the neutrino interaction with matter, and the neutrino mass eigenvalues
in matter. We give exact closed-form expressions for all partial oscillation probabilities in terms
of these basic quantities. Adopting the Standard Neutrino Model, we then examine how the exact
expressions for the partial oscillation probabilities might simplify by expanding in one of the small
parameters α and sin θ13 of this model. We show explicitly that for small α and sin θ13 there are
branch points in the analytic structure of the eigenvalues that lead to singular behavior of expansions
near the solar and atmospheric resonances. We present numerical calculations that indicate how to
use the small-parameter expansions in practice.
PACS Indices
Keywords:
I.
INTRODUCTION
In this paper, we develop an exact analytical
representation of neutrino oscillations [1] in matter within the framework of the Standard Neutrino
Model (SNM) [2] with 3 Dirac Neutrinos. Our
formulation expresses the time-evolution operator
S(t, t) in terms of the neutrino Hamiltonian Hν using
the Lagrange interpolation formula given in Ref. [3].
One distinctive feature of this formulation is its formal structure, which closely parallels the familiar
theory of neutrino oscillations in the vacuum limit.
Our exact analytical results are given by closed-form
expressions, thus making approximations unnecessary to achieve transparent predictions.
The paper is divided into two main parts. In the
first part, we summarize the main results of our theory. Details underlying the derivation are given in
Appendices. In the second, we address other analytical formulations found in the literature.
The neutrino oscillation probability based on the
same Hν and expanded in one of the small parameters of the SNM, α and sin2 θ13 , is of particular interest. The seminal work along these lines is found in
Refs. [4–6]. This work underlies many of the analyses and exploratory studies of experiments at present
and future neutrino facilities, including our earlier
work [7–10].
The present paper was undertaken, and used, for
the purpose of independently confirming the results
of Refs. [7–10]. We find that the accuracy of the
expanded oscillation probability is restricted by the
presence of branch points in the analytic structure of
the eigenvalues of neutrinos propagating in matter.
We also show that the regions where the expanded
results are reliable is different for expansions in α [4]
and sin2 θ13 [5, 6]. We then map out regions where
the expanded results are reliable by comparing numerical results to the exact results of our Hamiltonian formulation.
Another recent study [11] takes a complementary
approach and finds that the predictions of Refs. [7–
10] can be improved in certain regions using an exact
evaluation of the integral Iα∗ rather than the approximate one found there. It concludes that within
these regions, predictions of (µ, e) oscillations improve for certain values of the experimental parameters.
II.
NEUTRINO DYNAMICS
We will be interested in the dynamics of the
three known neutrinos and their corresponding antineutrinos in matter. This dynamics is determined
Final state interactions at the threshold of Higgs boson pair production
Zhentao Zhang∗
School of Physics, Peking University, Beijing 100871, China
arXiv:1501.04014v1 [hep-ph] 16 Jan 2015
We study the effect of final state interactions at the threshold of Higgs boson pair production in
the Glashow-Weinberg-Salam model. We consider three major processes of the pair production in
the model: lepton pair annihilation, ZZ fusion, and WW fusion. We find that the corrections caused
by the effect for these processes are markedly different. According to our results, the effect can cause
non-negligible corrections to the cross sections for lepton pair annihilation and small corrections for
ZZ fusion, and this effect is negligible for WW fusion.
Half a century after the construction of the GlashowWeinberg-Salam (GWS) model of electroweak interactions, the observation of Higgs boson at the Large Hadron
Collider (LHC) is its latest triumph [1, 2]. In the future,
further determination of the properties of Higgs boson
will be extremely important for us to understand the fundamental laws of nature. Any deviation from the predictions in the Standard Model can give us a valuable clue
to the long-hunted new physics. In order to ascertain the
role of Higgs field in the Standard Model, we need to measure its couplings to fermions and gauge bosons. Furthermore, to reconstruct the details of the Higgs potential,
we have to precisely determine the Higgs self-couplings.
At present, an important task of us is to explore potential effects for the Higgs self-interactions, and the BroutEnglert-Higgs mechanism must be precisely tested from
low to high energies. In this paper, we shall investigate
the effect of final state interactions at the threshold of
Higgs boson pair production in the GWS model. This
effect is tied to the low-energy properties of the Higgs
self-interactions.
After the electroweak symmetry breaking, the Higgs
self-interactions can be written in the form
V (H) =
1 2 2
λ
m H + λvH 3 + H 4 .
2 H
4
(1)
For the mass of Higgs boson mH ≃ 125.5 GeV and the
vacuum expectation value v ≃ 246 GeV, λ = m2H /2v 2 ≈
0.130.
To establish the non-relativistic potential for the Higgs
self-interactions, we need to consider the scattering process HH → HH to leading order in λ. There are four
Feynman diagrams that contribute, see Fig. 1.
Note that we can separately discuss the contributions
from each diagram, since there is Bose statistics involved
in the process. In the beginning, let us consider the contributions from the quadrilinear self-coupling of Higgs
boson. The diagram is shown in Fig. 1a, and its amplitude is
iM = −i6λ.
∗
[email protected]
(2)
(a)
(b)
(c)
(d)
FIG. 1. The tree-level diagrams for the scattering of two Higgs
bosons in the Standard Model. Solid line denotes Higgs boson.
According to the definition of Born scattering amplitude in non-relativistic quantum mechanics, the nonrelativistic potential in the momentum spaces is
3λ
.
Vequad (q) =
2m2H
(3)
Using the Fourier transform to Ve (q), we can get the nonrelativistic potential for this term
Vquad (r) =
3λ (3)
δ (r).
2m2H
(4)
The presence of the delta function comes from the specially local structure in the diagram.
The amplitude for the “annihilation” diagram in
Fig. 1b is
iM = −18λm2H
i
,
(p1 + p2 )2 − m2H
(5)
where p1 and p2 are the 4-momenta of the incoming Higgs
bosons. The amplitude in the non-relativistic domain can
be approximated as Eq. (2), and then we can get a well
defined non-relativistic potential. It is not necessary to
repeat the same calculations. The non-relativistic potential for the “annihilation” diagram is defined as
Vann (r) =
3λ (3)
δ (r).
2m2H
(6)
LU TP 15-04
January 2015
Leading chiral logarithms for the nucleon mass
arXiv:1501.03979v1 [hep-ph] 16 Jan 2015
Alexey. A. Vladimirov and Johan Bijnens
Department of Astronomy and Theoretical Physics, Lund University,
Sölvegatan 14A, SE 223 62 Lund, Sweden
Abstract. We give a short introduction to the calculation of the leading chiral logarithms, and present the results of the recent
evaluation of the LLog series for the nucleon mass within the heavy baryon theory. The presented results are the first example
of LLog calculation in the nucleon ChPT. We also discuss some regularities observed in the leading logarithmical series for
nucleon mass. The talk has been presented at Quark Confinement and Hadron Spectrum XI.
Keywords: Chiral logarithms, nucleon mass, Chiral perturbation theory
PACS: 12.39.Fe, 14.20.Dh
INTRODUCTION
Chiral perturbation theory (ChPT) is an efficient tool for the evaluation of hadron observables at low energies (for a
comprehensive introduction to meson and nucleon ChPT, see [1]). The predictions of ChPT are used in many branches
of modern physics from matching of lattice calculations (for a review see [2]), till the investigations of the nuclei
properties (see e.g. [3]).
Nowadays, all practically interesting quantities has been calculated at two-loop order (for the recent status of meson
ChPT see [4], also see the talk given by J.Bijnens [5]). However, the straightforward expansion to higher-orders of
ChPT is meaningless. The main point is the rapidly growing number of low-energy constants (LECs). Indeed, in the
meson ChPT at order p6 the number of LECs is of order of hundred (depending on the number of active mesons)[6]. In
the nucleon ChPT the number of fields and invariant structures grows even faster, and this amount of LECs is reached
already at order p4 [7]. Such an enormous amount of LECs cannot be fixed in any reasonable way by a recent data.
Since the straightforward way to the higher precision theoretical predictions is closed, one should investigate
other possibilities to improve the chiral perturbation series. One of the promising approach is the evaluation of the
leading logarithm (LLog) part of the chiral expansion. Besides the possibility to improve the theoretical estimations
for observables, the investigation of LLogs grants us a chance to understand the mathematical structure of the
theory at high orders of perturbative expansion. In contrast to renormalizable quantum field theories, where, roughly
speaking, LLog approximation consists in the powering of one-loop diagrams, LLogs in non-renormalizable theories
are highly non-trivial. The structure of LLogs in non-renormalizable theories resembles the structure of the whole
perturbative expansion. Therefore, observation of any regularities within LLog approximation is the reflection of
general regularities.
In ChPT (as in any non-renormalizable theory) the evaluation of LLog coefficient of any given order implies the
evaluation of diagrams at this order. However, this can be done in a relatively efficient way. In this paper we give a
short introduction to the calculation of the leading chiral logarithms, and present the results of the recent evaluation of
the LLog series for the nucleon mass: the first example of LLog calculation in the nucleon ChPT.
LEADING LOGARITHMS IN CHPT
In ChPT the LLog coefficients can be calculated using solely one-loop calculations. This statement was proven in [8],
and relies on the fact that the LLog coefficient is proportional to the main ultraviolet (UV) divergency of a diagram.
In its own turn, the leading UV divergency of the diagram is composed from the individual divergencies of one-loop
subgraphs. In the renormalizable theories, the individual UV divergences of one-loop subgraphs are just multiplied on
Preprint typeset in JHEP style - HYPER VERSION
SLAC-PUB-16176
TTK-15-01
arXiv:1501.03942v1 [hep-ph] 16 Jan 2015
Simplified models for same-spin new physics scenarios
Lisa Edelh¨
ausera , Michael Kr¨
amera,b and Jory Sonnevelda
a
b
Institute for Theoretical Particle Physics and Cosmology, RWTH Aachen University,
52056 Aachen, Germany
SLAC National Accelerator Laboratory, Stanford University, Stanford, CA 94025, USA
Abstract: Simplified models are an important tool for the interpretation of searches for
new physics at the LHC. They are defined by a small number of new particles together with
a specific production and decay pattern. The simplified models adopted in the experimental
analyses thus far have been derived from supersymmetric theories, and they have been
used to set limits on supersymmetric particle masses. We investigate the applicability
of such simplified supersymmetric models to a wider class of new physics scenarios, in
particular those with same-spin Standard Model partners. We focus on the pair production
of quark partners and analyze searches for jets and missing energy within a simplified
supersymmetric model with scalar quarks and a simplified model with spin-1/2 quark
partners. Despite sizable differences in the detection efficiencies due to the spin of the new
particles, the limits on particle masses are found to be rather similar. We conclude that the
supersymmetric simplified models employed in current experimental analyses also provide
a reliable tool to constrain same-spin BSM scenarios.
Keywords: Beyond the Standard Model, phenomenological models.
January 2015
BI-TP 2015/04
Quarkonium Binding and Entropic Force
arXiv:1501.03940v1 [hep-ph] 16 Jan 2015
Helmut Satz
Fakult¨at f¨
ur Physik, Universit¨at Bielefeld
D-33501 Bielefeld, Germany
Abstract:
¯ bound state represents a balance between repulsive kinetic and attractive potential
A QQ
energy. In a hot quark-gluon plasma, the interaction potential experiences medium effects.
Color screening modifies the attractive binding force between the quarks, while the increase
¯ separation gives rise to a growing repulsion. We study the role of these
of entropy with QQ
¯ binding and dissociation. It is found that the relevant potential
phenomena for in-medium QQ
¯ binding is the free energy F ; with increasing QQ
¯ separation, further binding through
for QQ
the internal energy U is compensated by repulsive entropic effects.
1
Introduction
The concept of entropic forces, emerging as a result of collective many-body phenomena, has
in recent times attracted increasing interest; see e. g. [1–4]. The effect of such forces arises
from the thermodynamic drive of a many-body system to increase its entropy, rather than from
a specific underlying microscopic force. This can also provide a way of studying the role of
entropy maximization for a specific dynamic system immersed in a thermal medium. We here
want to use this approach to address quarkonium binding and dissociation in a hot deconfined
quark-gluon plasma [5].
¯ binding in a medium of temperature T is in terms of the
The simplest approach to study QQ
Schr¨odinger equation
"
#
1 2
2mQ −
∇ + V (r, T ) Φi (r, T ) = Mi (T )Φi (r, T ).
mQ
(1)
Its solution gives the resulting quarkonium masses Mi (T ), with i = 0 for the ground state
and i = 1, 2, ... for the subsequent excited states. Here mQ denotes the c or b quark mass,
while V (r, T ) describes the in-medium binding potential. To obtain a feeling for the resulting
behavior, it is helpful to consider the semi-classical limit of eq. (1) [6],
"
2mQ +
c
mQ r 2
#
+ V (r, T ) = E(r, T ),
1
(2)
Charge asymmetry in the differential cross section of high-energy
bremsstrahlung in the field of a heavy atom
P.A. Krachkov1, 2, ∗ and A. I. Milstein1, †
1
Budker Institute of Nuclear Physics, 630090 Novosibirsk, Russia
arXiv:1501.03897v1 [hep-ph] 16 Jan 2015
2
Novosibirsk State University, 630090 Novosibirsk, Russia
(Dated: January 19, 2015)
Abstract
The distinction between the charged particle and antiparticle differential cross sections of highenergy bremsstrahlung in the electric field of a heavy atom is investigated. The consideration is
based on the quasiclassical approximation to the wave functions in the external field. The charge
asymmetry (the ratio of the antisymmetric and symmetric parts of the differential cross section)
arises due to the account for the first quasiclassical correction to the differential cross section. All
evaluations are performed with the exact account of the atomic field. We consider in detail the
charge asymmetry for electrons and muons. For electrons, the nuclear size effect is not important
while for muons this effect should be taken into account. For the longitudinal polarization of the
initial charged particle, the account for the first quasiclassical correction to the differential cross
section leads to the asymmetry in the cross section with respect to the replacement ϕ → −ϕ, where
ϕ is the azimuth angle between the photon momentum and the momentum of the final charged
particle.
PACS numbers: 12.20.Ds, 32.80.-t
Keywords:
∗
Electronic address: peter˙[email protected]
†
Electronic address: [email protected]
1
arXiv:1501.03888v1 [hep-ph] 16 Jan 2015
KEK-TH-1791
A symmetry breaking mechanism by parity assignment
in the noncommutative Higgs model
Masaki J.S. Yang
Institute of Particle and Nuclear Studies,
High Energy Accelerator Research Organization (KEK)
Tsukuba 305-0801, Japan
Abstract
We apply the orbifold GUT mechanism to the noncommutative Higgs model.
An assignment of Z2 parity to the “pre-gauge fields” induces both of the parity
assignments of the gauge and Higgs bosons, because these bosons are treated as
some kind of composite fields in this formalism. As a result, a part of the gauge
bosons and colored triplet Higgs boson receive heavy mass comparable to GUT
scale, and the gauge symmetry is broken. No particle appear other than the SM
ones in the massless states.
Zc (4200)+ decay width as a charmonium-like tetraquark state
Wei Chen and T. G. Steele
Department of Physics and Engineering Physics,
University of Saskatchewan, Saskatoon, SK, S7N 5E2, Canada
Hua-Xing Chen∗
School of Physics and Nuclear Energy Engineering and Intern ational Research Center
for Nuclei and Particles in the Cosmos, Beihang University, Beijing 100191, China
arXiv:1501.03863v1 [hep-ph] 16 Jan 2015
Shi-Lin Zhu†
School of Physics and State Key Laboratory of Nuclear Physics and Technology, Peking University, Beijing 100871, China
Collaborative Innovation Center of Quantum Matter, Beijing 100871, China
Center of High Energy Physics, Peking University, Beijing 100871, China
To identify the nature of the newly observed charged resonance Zc (4200)+ , we study its hadronic
¯ ∗0 as a charmonium-like
decays Zc (4200)+ → J/ψπ + , Zc (4200)+ → ηc ρ+ and Zc (4200)+ → D+ D
tetraquark state. In the framework of the QCD sum rules, we calculate the three-point functions
and extract the coupling constants and decay widths for these interaction vertices. Including all
these channels, the full decay width of the Zc (4200)+ state is consistent with the experimental value
reported by the Belle Collaboration, supporting the tetraquark interpretation of this state.
PACS numbers: 12.38.Lg, 11.40.-q, 12.39.Mk
Keywords: Three-point function, Tetraquark state, QCD sum rules
I.
INTRODUCTION
Recently, a new charged charmoniumlike resonance Zc (4200)+ was observed by the Belle Collaboration [1]. It
was observed in the Zc (4200)+ → J/ψπ + process with the mass and decay width M = 4196+31+17
−29−13 MeV and
Γ = 370+70+70
MeV,
with
a
significance
of
6.2σ.
Its
preferred
assignment
of
the
quantum
numbers
is J P = 1+ . The
−70−132
+
G PC
+ +−
G-parity of Zc (4200) is positive. Thus, the quantum numbers of its neutral partner is I J
=1 1 .
The family of the charged charmoniumlike states have become more abundant after the discovery of Zc (4200)+ [1]
and Zc (4050) [2]. Before this, the first member Z(4430)+ was observed in the ψ(2S)π + invariant mass spectrum in
¯ 0 → ψ(2S)π + K − by the Belle Collaboration [3] and confirmed recently by the LHCb Collaboration [4].
the process B
Later, Belle also reported a broad doubly peaked structure in the π + χc1 invariant mass distribution, which are called
Z(4050)+ and Z(4250)+ [5]. Several other similar charged states were observed in last two years. In 2013, the BESIII
Collaboration reported Zc (3900)+ in J/ψπ + final states in the process Y (4260) → J/ψπ + π − [6]. Zc (3900)+ was also
observed by Belle [7] and confirmed in CLEO data [8]. The BESIII Collaboration also observed Zc (4025)± in the
¯ ∗ )± π ∓ process [9] and Zc (4020)± in the hc π ± mass spectrum in the
π ∓ recoil mass spectrum in the e+ e− → (D∗ D
+ −
+ −
process e e → hc π π [10]. Moreover, the Belle Collaboration also observed two charged bottomoniumlike states
Zb (10610) and Zb (10650) in the π ± Υ(nS) and hb π ± mass spectra in the Υ(5S) decay [11].
These newly observed charged states have the exotic flavor contents c¯
cud¯ for Zc states and b¯bud¯ for Zb states. It is
natural to understand them as different manifestations of four-quark states: hadron molecules, hadrocharmonium or
¯ 1 molecular state in Refs. [12–16] and a tetraquark
tetraquark states. For example, Z(4430)+ was described as a D∗ D
state in Refs. [17–19], the Zc (3900)+ was speculated to be a molecular state in Refs. [20–22], the Zc (4025)+ was
¯ tetraquark with the quantum numbers
¯ ∗ molecular state in Refs. [23–26] and as a [cu][¯
cd]
interpreted as a D∗ D
P
+
¯ ∗ and B
¯ ∗ B ∗ molecular states in Refs. [28–30].
J = 2 in Ref. [27]. Zb (10610) and Zb (10650) were studied as BB
Being composed of a diquark and antidiquark pair, a hidden-charm tetraquark state can decay very easily into a
pair of open-charm D mesons or one charmonium state plus a light meson through quark rearrangement, implying
that tetraquark states should be very broad resonances while the experimental XY Z states are usually quite narrow,
such as Zc (3900)+ [6–8] and Zc (4025)+ [9, 10]. However, the experimental width value of the Zc (4200)+ [1] is broad
enough to be a good tetraquark candidate. In Ref. [31], Zc (4200)+ was studied as a tetraquark state by considering
∗ Electronic
† Electronic
address: [email protected]
address: [email protected]
Extracting the Odderon from pp and p¯
p scattering data
Andr´as Ster∗
Wigner Research Centre for Physics, Hungarian Academy of Sciences,
arXiv:1501.03860v1 [hep-ph] 16 Jan 2015
H-1525 Budapest 114, P.O. Box 49, Hungary
L´aszl´o Jenkovszky†
BITP, National Academy of Sciences of Ukraine, Kiev, 03680 Ukraine and
Wigner Research Centre for Physics, Hungarian Academy of Sciences,
H-1525 Budapest 114, P.O. Box 49, Hungary
Tam´as Cs¨org˝o‡
Wigner Research Centre for Physics, Hungarian Academy of Sciences,
H-1525 Budapest 114, P.O. Box 49, Hungary and
KRF, H-3200 Gy¨ongy¨os, M´atrai u
´t 36, Hungary
(Dated: January 19, 2015)
Abstract
Starting from a simple empirical parametrization of the scattering amplitude, successfully describing the dip-bump structure of elastic pp scattering in t at fixed values of s, we construct a toy
model interpolating between missing energy intervals to extract the Odderon contribution from
the difference between p¯p and pp elastic and total cross sections. The model is fitted to data from
√
s = 23.5 GeV to 7 TeV and used to extract the Odderon and its ratio to the Pomeron.
PACS numbers: 13.75.Cs, 13.85.-t
Keywords: Elastic and total cross sections
∗
†
‡
[email protected]
[email protected]
[email protected]
1
LYCEN 2015-01
arXiv:1501.03818v1 [hep-ph] 15 Jan 2015
Anarchic Yukawas and top partial compositeness:
the flavour of a successful marriage
Giacomo Cacciapagliaa,b , Haiying Caia,b , Thomas Flackec , Seung J. Leec,d , Alberto Parolinic,e ,
Hugo Serˆodioc
a
b
Universit´e de Lyon, F-69622 Lyon, France; Universit´e Lyon 1, Villeurbanne, France
CNRS/IN2P3, UMR5822, Institut de Physique Nucl´eaire de Lyon F-69622 Villeurbanne Cedex,
France
c
Department of Physics, Korea Advanced Institute of Science and Technology, 335 Gwahak-ro,
Yuseong-gu, Daejeon 305-701, Korea
d
e
School of Physics, Korea Institute for Advanced Study, Seoul 130-722, Korea
Center for Axion and Precision Physics, IBS, 291 Daehak-ro, Yuseong-gu, Daejeon 305-701, Korea
Abstract
The top quark can be naturally singled out from other fermions in the Standard Model due to its
large mass, of the order of the electroweak scale. We follow this reasoning in models of pseudo Nambu
Goldstone Boson composite Higgs, which may derive from an underlying confining dynamics. We
consider a new class of flavour models, where the top quark obtains its mass via partial compositeness,
while the lighter fermions acquire their masses by a deformation of the dynamics generated at a high
flavour scale. One interesting feature of such scenario is that it can avoid all the flavour constraints
without the need of flavour symmetries, since the flavour scale can be pushed high enough. We show
that both flavour conserving and violating constraints can be satisfied with top partial compositeness
without invoking any flavour symmetry for the up-type sector, in the case of the minimal SO(5)/SO(4)
coset with top partners in the four-plet and singlet of SO(4). In the down-type sector, some degree
of alignment is required if all down-type quarks are elementary. We show that taking the bottom
quark partially composite provides a dynamical explanation for the hierarchy causing this alignment.
We present explicit realisations of this mechanism which do not require to include additional bottom
partner fields. Finally, these conclusions are generalised to scenarios with non-minimal cosets and top
partners in larger representations.
Constraining the Eq. of State of Super-Hadronic Matter from Heavy-Ion Collisions
Scott Pratt,1 Evan Sangaline,1 Paul Sorensen,2 and Hui Wang2
1
Department of Physics and Astronomy and National Superconducting Cyclotron Laboratory
Michigan State University, East Lansing, MI 48824, USA
2
Brookhaven National Laboratory, Upton, New York 11973, USA
(Dated: January 19, 2015)
arXiv:1501.04042v1 [nucl-th] 16 Jan 2015
The equation of state of QCD matter for temperatures near and above the quark-hadron transition
(∼ 165 MeV) is inferred within a Bayesian framework through the comparison of data from the
Relativistic Heavy Ion Collider and from the Large Hadron Collider to theoretical models. State-ofthe-art statistical techniques are applied to simultaneously analyze multiple classes of observables
while varying 14 independent model parameters. The resulting posterior distribution over possible
equations of state is consistent with results from lattice gauge theory.
I.
INTRODUCTION
Relativistic heavy ion collisions have been proposed as
a means for investigating the equation of state of hot
matter. For fixed target energies of E/A <
∼ 10 GeV,
analyses of heavy ion collisions have significantly constrained the compressibility of dense hadronic matter [1]
for temperatures <
∼ 100 MeV. Higher energy collisions
probe conditions near and above the transition temperature, where lattice calculations have shown that in a
narrow temperature band, 150 < T < 200 MeV, the
scalar quark condensate melts [2], the degrees of freedom
change [3], and the speed of sound has a minimum [4].
In fact, for some time the transition was postulated to
contain a first-order phase transition accompanied by a
sizable latent heat.
In contrast to the progress of lattice calculations, experimental determination of the equation of state at
high temperature has remained semi-quantitative. The
stunted progress has not been due to a shortage of experimental observables that are known to be sensitive
to the equation of state. Van Hove associated the dependence of the mean transverse momentum, hpt i, as a
function of multiplicity as tool for determining the equation of state [5]. Two-particle femtoscopic correlations
were proposed as a signal for a first-order phase transition [6]. Measurements of azimuthal elliptic flow, which
are now mainly associated with determining the viscosity,
were also shown to be sensitive to the equation of state
[7, 8]. Multiplicities, which are related to entropy, have
also been used to constrain the equation of state [9]. Although femtoscopic analyses have shown that a first order
equation of state with a large latent heat is highly unlikely [10], and that an extremely stiff equation of state,
such as that of a pion gas, is also inconsistent with data
[11], a more quantitative statement of how well the equation of state is constrained has proven elusive. Even if
analysis of experimental data cannot compete with lattice calculations in determining the equation of state for
perfectly equilibrated matter, constraining the equation
of state by experiment can help validate the statement
that the matter created in heavy-ion collisions behaves
like an equilibrated quark gluon plasma.
The road block to turning these sensitivities into a
more robust and rigorous determination of the equation
of state has been the intertwined dependencies between
the many unknown features and parameters of the model,
and the numerous classes of measurement. Two developments now make this next step possible. First, the models used to describe the bulk behavior have converged to
a standard framework based on relativistic viscous hydrodynamics for the evolution of the high temperature
region, >
∼ 165 MeV, [12] coupled to a microscopic simulation of the lower temperature hadronic stage based on
binary collisions. The initial evolution, which feeds into
the hydrodynamic description, remains rather undefined,
but one can represent those uncertainties parametrically.
The second development is in the statistical methodologies and tools required to compare heterogenous data to
models where a large number of parameters are required
to encapsulate the many model uncertainties. Here we
use the statistical tools described in [13] to constrain 14
parameters via a Markov-chain Monte Carlo. The statistical tools are based on a Gaussian-process model emulator, which allows one to estimate observables for a given
point in parameter space by interpolating from a fixed
number of full-model runs.
II.
METHODOLOGY
Here we report on comparisons of model calculations to
data from Au+Au collisions from the highest RHIC (Relativistic Heavy Ion Collider) energy, 100A GeV + 100A
GeV, and from Pb+Pb collisions at the Large Hadron
Collider (LHC), 1.38A TeV + 1.38A TeV. The hydrodynamic and hadronic simulations were the same as those
used in [13] to analyze RHIC data. The analysis involves 14 parameters, two of which vary the equation of
state. The statistical method returns a sampling of the
14-dimensional space that is weighted by the likelihood,
2 Y
(mod)
(exp)
L(~x) ∼
exp − (zi
(~x) − zi
/2. . (1)
i
Here, ~x is the 14-dimensional vector describing a point
in parameter space and zi are principal components of
Comments on “Interference phenomena in the J p = 1/2− - wave in η photoproduction”
by A.V. Anisovich, E. Klempt, B. Krusche, V.A. Nikonov, A.V. Sarantsev, U. Thoma,
D. Werthmuller, arXiv:1501.02093v1 [nucl-ex].
Viacheslav Kuznetsov
arXiv:1501.03961v1 [nucl-ex] 16 Jan 2015
1
Petersburg Nuclear Physics Institute, Gatchina, 188300, St. Petersburg, Russia
The authors of Ref. [1] claimed that “ ... narrow structure observed in the excitation function
of γn → ηn can be reproduced fully with a particular interference pattern in the J p = 1/2−
partial wave...” while a narrow structure in Compton scattering off the neutron is “...a stand-alone
observation unrelated to the structure observed in γn → ηn...”. The source for the second statement
may be a simple numerical error. If so, the interpretation of the narrow structure in γn → ηn as
interference effects in the J p = 1/2− -wave and some conclusions from Ref. [1] are questionable.
The observation of a narrow enhancement at W ∼ 1.68
GeV in η photoproduction [2–5] and Compton scattering off the neutron [6] (the so-called “neutron anomaly”)
raised intensive debates about its nature. One possible
explanation is a signal of a nucleon resonance with unusual properties: the mass near M ∼ 1.68 GeV, the narrow (Γ ≤ 25 MeV) width, the strong photoexcitation on
the neutron, and the suppressed decay to πN final state.
A new one-star N ∗ (1685) resonance was included into
the listing of the Particle Data Group [7].
On the other hand, several groups tried to explain the
bump in the γn → ηn cross section in terms of the interference of well-known wide resonances. The recent
attempt was done in Ref. [1]. The authors concluded
that it can be full explained by the interference of wellknown resonances while the inclusion of N ∗ (1685) only
deteriorates the data fit.
One major challenge for this interpretation is the observation of a narrow peak at the same energy in Compton scattering on the neutron at GRAAL [6]. The
authors of Ref. [1] estimated the total cross section of
N ∗ (1685) in γn → γn assuming the mass M = 1670
MeV, the width Γtot = 30 MeV, and (a priori) the quantum numbers P11 . The result σres = 10.8 pb led to a
conclusion that “...this value is far below the sensitivity
of the GRAAL experiment. If it is not a statistical fluctuation, ... it is a stand-alone observation unrelated to
the structure observed in γn → ηn...”.
Unfortunately, there might be a simple numerical error: if to check Eq.(1) from Ref. [1], the correct number
is σres = 10.8 nb (i.e. 1000 times larger).
Even this number may be pessimistic. The results from
GRAAL and CBELSA/TAPS suggest Γtot ≤ 25 MeV. If
to set Γtot = 20 MeV then σres = 24.3 nb. If in addition
to assume that N ∗ (1685) is a higher-spin resonance, σres
may be significantly larger.
The peak in γn → γn at GRAAL was observed at
157◦. The measured differential cross section of Compton
scattering on the proton at 160◦ and Eγ = 1.025 GeV
is 27.1 ± 5.4 nb/str [8]. In accordance with dispersionrelation calculations [9] Compton cross section on the
neutron (without narrow resonance) at 160◦ and Eγ = 1
GeV may be significantly (∼ 5 times) smaller than that
on the proton. The peak of N ∗ (1685) on the top of the
flat γn → γn cross section could and, if N ∗ (1685) does
exist, should be seen at GRAAL.
The observation of the peak in Compton scattering at
the same energy as in γn → ηn challenges the explanation of the neutron anomaly in terms of interference
effects. The specific interference of wide resonances cannot generate a narrow peak in η photoproduction which is
governed by only isospin-1/2 resonances, simultaneously
generate a peak at the same energy in Compton scattering which governed by both isospin-1/2 and isospin-3/2
resonances, and generate neither of peak in γn → π 0 n
which is governed by the same resonances as Compton
scattering.
At present, the only available explanation is the existence of N ∗ (1685).
[1] by A.V. Anisovich, E. Klempt, B. Krusche, V.A. Nikonov,
A.V. Sarantsev,
U. Thoma,
D. Werthmuller,
arXiv:1501.02093v1 [nucl-ex].
[2] V. Kuznetsov et al., Phys. Lett. B 647, 23 (2007).
[3] I. Jaegle et al., Phys. Rev. Lett. 100, 252002 (2008); I. Jaegle et al., Eur.Phys.J. A47, 89 (2011).
[4] F. Miyahara et al., Prog. Theor. Phys. Suppl. 168, 90
(2007).
[5] D. Werthmulleret al., Phys.Rev.Lett. 111 (2013) 23,
232001; Phys.Rev. C90, 015205 (2014).
[6] V. Kuznetsov et al., Phys. Rev. C83, 022201 (2011).
[7] K.A.Olive et al., [the Particle Data Group Collboration],
Chin. Phys. C38,090001 (2014).
[8] Y. Wada et al., Nucl. Phys. b247, 313 (1984).
[9] A. Lvov, V. Petrun’kin, and M. Shumacher, Phys. Rev.
C55, 355, (1997), and A. L’vov, Private communication.
arXiv:1501.03894v1 [nucl-ex] 16 Jan 2015
Subthreshold Ξ− Production in Collisions of p (3.5 GeV) + Nb
G. Agakishiev7 , O. Arnold9 , A. Balanda3 , D. Belver18 , A. V. Belyaev7 , J. C. Berger-Chen9 , A. Blanco2 ,
M. B¨ohmer10 , J. L. Boyard16 , P. Cabanelas18,a , S. Chernenko7 , A. Dybczak3 , E. Epple9 , L. Fabbietti9 , O. V. Fateev7 ,
P. Finocchiaro1 , P. Fonte2,b , J. Friese10 , I. Fr¨ohlich8 , T. Galatyuk5,c , J. A. Garz´on18 , R. Gernh¨auser10 , K. G¨obel8 ,
M. Golubeva13 , D. Gonz´alez-D´ıaz5 , F. Guber13 , M. Gumberidze5,16 , T. Heinz4 , T. Hennino16 , R. Holzmann4 ,
A. Ierusalimov7 , I. Iori12,d , A. Ivashkin13 , M. Jurkovic10 , B. K¨ampfer6,e , T. Karavicheva13 , I. Koenig4 , W. Koenig4 ,
B. W. Kolb4 , G. Kornakov18 , R. Kotte6 , A. Kr´asa17 , F. Krizek17 , R. Kr¨ucken10 , H. Kuc3,16 , W. K¨uhn11 , A. Kugler17 ,
A. Kurepin13 , V. Ladygin7 , R. Lalik9 , S. Lang4 , K. Lapidus9 , A. Lebedev14 , T. Liu16 , L. Lopes2 , M. Lorenz8,c , L. Maier10 ,
A. Mangiarotti2 , J. Markert8 , V. Metag11 , B. Michalska3 , J. Michel8 , C. M¨untz7 , L. Naumann6 , Y. C. Pachmayer8 ,
M. Palka3 , Y. Parpottas15,f , V. Pechenov4 , O. Pechenova8 , J. Pietraszko4 , W. Przygoda3 , B. Ramstein16 , A. Reshetin13 ,
A. Rustamov8 , A. Sadovsky13 , P. Salabura3 , A. Schmah9,g , E. Schwab4 , J. Siebenson9 , Yu. G. Sobolev17 , S. Spataro11,h ,
B. Spruck11 , H. Str¨obele8 , J. Stroth8,4 , C. Sturm4 , A. Tarantola8 , K. Teilab8 , P. Tlusty17 , M. Traxler4 , R. Trebacz3 ,
H. Tsertos15 , T. Vasiliev7 , V. Wagner17 , M. Weber10 , C. Wendisch4 , J. W¨ustenfeld6 , S. Yurevich4 , Y. V. Zanevsky7
(HADES collaboration)
1
Instituto Nazionale di Fisica Nucleare - Laboratori Nazionali del Sud, 95125 Catania, Italy
LIP-Laborat´orio de Instrumentac¸a˜ o e F´ısica Experimental de Part´ıculas , 3004-516 Coimbra, Portugal
3
Smoluchowski Institute of Physics, Jagiellonian University of Cracow, 30-059 Krak´ow, Poland
4
GSI Helmholtzzentrum f¨ur Schwerionenforschung GmbH, 64291 Darmstadt, Germany
5
Technische Universit¨at Darmstadt, 64289 Darmstadt, Germany
6
Institut f¨ur Strahlenphysik, Helmholtz-Zentrum Dresden-Rossendorf, 01328 Dresden, Germany
7
Joint Institute of Nuclear Research, 141980 Dubna, Russia
8
Institut f¨ur Kernphysik, Johann Wolfgang Goethe-Universit¨at, 60438 Frankfurt, Germany
9
Excellence Cluster ’Origin and Structure of the Universe’, 85748 Garching, Germany
10
Physik Department E12, Technische Universit¨at M¨unchen, 85748 Garching, Germany
11
II.Physikalisches Institut, Justus Liebig Universit¨at Giessen, 35392 Giessen, Germany
12
Istituto Nazionale di Fisica Nucleare, Sezione di Milano, 20133 Milano, Italy
13
Institute for Nuclear Research, Russian Academy of Science, 117312 Moscow, Russia
14
Institute of Theoretical and Experimental Physics, 117218 Moscow, Russia
15
Department of Physics, University of Cyprus, 1678 Nicosia, Cyprus
Institut de Physique Nucl´eaire (UMR 8608), CNRS/IN2P3 - Universit´e Paris Sud, F-91406 Orsay Cedex, France
17
Nuclear Physics Institute, Academy of Sciences of Czech Republic, 25068 Rez, Czech Republic
18
LabCAF F. F´ısica, Univ. de Santiago de Compostela, 15706 Santiago de Compostela, Spain
2
16
a
h
also at Nuclear Physics Center of University of Lisbon, 1649-013 Lisboa, Portugal
b
also at ISEC Coimbra, 3030-199 Coimbra, Portugal
c
also at ExtreMe Matter Institute EMMI, 64291 Darmstadt, Germany
d
also at Dipartimento di Fisica, Universit`a di Milano, 20133 Milano, Italy
e
also at Technische Universit¨at Dresden, 01062 Dresden, Germany
f
also at Frederick University, 1036 Nikosia, Cyprus
g
now at Lawrence Berkeley National Laboratory, Berkeley, USA
now at Dipartimento di Fisica Generale and INFN, Universit`a di Torino, 10125 Torino, Italy
(Dated: January 19, 2015)
Results on the production of the double-strange cascade hyperon Ξ− are reported for collisions of
p (3.5 GeV) + Nb, studied with the High Acceptance Di-Electron Spectrometer (HADES) at SIS18 at GSI
Helmholtzzentrum for Heavy-Ion Research, Darmstadt. For the first time, subthreshold Ξ− production is observed in proton-nucleus interactions. Assuming a Ξ− phase-space distribution similar to that of Λ hyperons,
the production probability amounts to PΞ− = (2.0 ±0.4 (stat) ±0.3 (norm) ±0.6 (syst))×10−4 resulting in
a Ξ− /(Λ + Σ0 ) ratio of PΞ− / PΛ+Σ0 = (1.2 ± 0.3 (stat) ± 0.4 (syst)) × 10−2 . Available model predictions
are significantly lower than the estimated Ξ− yield.
PACS numbers: 25.75.Dw, 25.75.Gz
The double-strange Ξ− baryon (also known as cascade particle) when produced in elementary nucleon-nucleon (NN)
collisions must be co-produced with two kaons ensuring
strangeness conservation, NN → NΞKK. In fixed-target experiments, this requires a minimum beam energy of Ethr =
√
3.74 GeV ( sthr = 3.25 GeV). In heavy-ion and even in