arXiv:1501.06883v1 [nucl-ex] 27 Jan 2015

Progress in characterization of the Photomultiplier
Tubes for XENON1T Dark Matter Experiment.
arXiv:1502.01000v1 [astro-ph.IM] 3 Feb 2015
Alexey Lyashenkoa∗ , XENON Collaboration
Abstract—We report on the progress in characterization of
the Hamamatsu model R11410-21 Photomultiplier tubes (PMTs)
for XENON1T dark matter experiment. The absolute quantum
efficiency (QE) of the PMT was measured at low temperatures
down to -110 0 C (a typical the PMT operation temperature in
liquid xenon detectors) in a spectral range from 154.5 nm to
400 nm. At -110 0 C the absolute QE increased by 10-15% at
175 nm compared to that measured at room temperature. A new
low power consumption, low radioactivity voltage divider for the
PMTs is being developed. The measurement results showed that
the PMT with the current version of the divider demonstrated
a linear response (within 5%) down to 5·104 photoelectrons
at a rate of 200 Hz. The radioactive contamination induced
by the PMT and the PMT voltage divider materials satisfies
the requirements for XENON1T detector not to exceed a total
radioactive contamination in the detector of 0.5 evts/year/1tonn.
Most of the PMTs received from the manufacturer showed a
high quantum efficiency exceeding 30%. In the mass production
tests the measurements at room temperature showed clear single
photoelectron peaks for all PMTs been under study. The optimal
operation conditions were found at a gain of 2·106 . The operation
stability for most of the PMTs was also demonstrated at a
temperature of -100 0 C. A dedicated setup was built for testing
the PMTs in liquid xenon using the XENON1T signal readout
components including voltage dividers, cables and feedthroughs.
The PMTs tested in liquid xenon demonstrated a stable operation
for a time period of more than 5 months.
Index Terms—XENON1T, PMTs, Liquid Noble Gas Detectors.
Some of the PMT characteristics including Quantum Efficiency (QE), gain and linearity of the response could be
temperature dependent and, therefore, should be measured at
liquid xenon temperature. The performance of the R11410
PMT in a liquid xenon environment was studied elsewhere
[3] demonstrating the stability of gain for a period of more
then 5 months and a stable PMT operation in the vicinity of
a strong electric field.
Understanding the internal radioactive background in
XENON1T detector is a crucial part of the experiment as it
will define its sensitivity to the dark matter signal. As shown
in [4] and [5], the most of the gamma background and a
significant part of the neutron background comes from the
PMTs as they contain much more radioactive contaminants
than the other elements of the detector. Therefore, it is
also crucial to measure the radioactive background from the
detector components and the PMTs in particular [6].
In the present article we summarize our ongoing campaign
on the characterization of the R11410 PMTs that will be
employed in XENON1T detector. We report on the recent
measurements of the absolute QE at liquid xenon temperature.
The measurement results of the linearity of the PMT response
are also shown. We present the experimental setups for the
mass production tests both at a room temperature and at -100
0
C. We also present the experimental setup and the results of
the PMT measurements in liquid xenon environment.
I. I NTRODUCTION
T
HE new Hamamatsu model R11410-21 Photomultiplier
Tube (PMT) specially designed for XENON1T direct
dark matter search experiment [1] that is being under construction in Gran Sasso National Laboratory (Italy) since June
2013. The detector sensitive volume will accommodate 2.2 ton
of liquid xenon monitored by 248 PMTs. The PMT response to
various light signals produced in a liquid xenon detector should
be well understood as it directly affects the experimental
results. Before instrumenting the detector with the PMTs each
individual PMT has to be carefully characterized by measuring
its single photoelectron peak, gain and afterpulse rate. Some
properties like linearity of the response and photocathode
uniformity are common for most of the PMTs. These properties can be measured for few PMT samples. Gain, dark
count rate, linearity of the response, photocathode uniformity
and afterpulsing for the older version R11410-10 PMTs were
previously measured [2].
∗
Corresponding author. E-mail:[email protected]
University of California Los Angeles, Department of Physics and Astronomy, 475 Portola Plaza, Los Angeles, CA 90095
a
II. PMT
ARRANGEMENT IN
XENON1T
DETECTOR
As mentioned above XENON1T dark matter experiments
will accommodate a total of 248 PMTs as shown in figure 1.
These PMTs will be split between the top and the bottom
PMT supporting structures shown in figure 1b. The top PMT
supporting structure (see figure 1b) will be instrumented with
127 PMTs arranged in concentric circles in order to improve
the resolution of the event position reconstruction. The bottom
PMT array (see figure 1b) will incorporate 121 PMTs arranged
in a hexagonal pattern to optimize the optical coverage. The
gaps between the PMTs in the top and in the bottom PMT
arrays will be covered with a single piece PTFE reflector to
improve the light collection efficiency.
III. M EASUREMENTS OF THE ABSOLUTE QE
R11410-10 PMT S
FOR
The measurements of the absolute QE of Hamamatsu model
R11410-10 PMTs has been recently reported [8]. Although the
measurements were performed for the older PMT version, they
Measurement of the Formation Rate of Muonic Hydrogen Molecules
V.A. Andreev,1 T.I. Banks,2 R.M. Carey,3 T.A. Case,2 S.M. Clayton,4, ∗ K.M. Crowe,2, † J. Deutsch,5, † J. Egger,6
S.J. Freedman,2, † V.A. Ganzha,1 T. Gorringe,7 F.E. Gray,8, 2 D.W. Hertzog,9, 4 M. Hildebrandt,6 P. Kammel,9, 4
B. Kiburg,9, 4, ‡ S. Knaack,4, § P.A. Kravtsov,1 A.G. Krivshich,1 B. Lauss,6 K.R. Lynch,10 E.M. Maev,1
O.E. Maev,1 F. Mulhauser,4, 6 C. Petitjean,6 G.E. Petrov,1 R. Prieels,5 G.N. Schapkin,1 G.G. Semenchuk,1
M.A. Soroka,1 V. Tishchenko,7, ¶ A.A. Vasilyev,1 A.A. Vorobyov,1 M.E. Vznuzdaev,1 and P. Winter9, 4, ∗∗
(MuCap Collaboration)
1
Petersburg Nuclear Physics Institute, Gatchina 188350, Russia
Department of Physics, University of California, Berkeley, and LBNL, Berkeley, CA 94720, USA
3
Department of Physics, Boston University, Boston, MA 02215, USA
4
Department of Physics, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA
5
Institute of Nuclear Physics, Université Catholique de Louvain, B-1348, Louvain-la-Neuve, Belgium
6
Paul Scherrer Institute, CH-5232 Villigen PSI, Switzerland
7
Department of Physics and Astronomy, University of Kentucky, Lexington, KY 40506, USA
8
Department of Physics and Computational Science, Regis University, Denver, CO 80221, USA
9
Department of Physics, University of Washington, Seattle, WA 98195, USA
10
Department of Earth and Physical Sciences, York College, City University of New York, Jamaica, NY 11451, USA
(Dated: February 4, 2015)
arXiv:1502.00913v1 [nucl-ex] 3 Feb 2015
2
Background: The rate λppµ characterizes the formation of ppµ molecules in collisions of muonic pµ atoms with
hydrogen. In measurements of the basic weak muon capture reaction on the proton to determine the pseudoscalar
coupling gP , capture occurs from both atomic and molecular states. Thus knowledge of λppµ is required for a
correct interpretation of these experiments.
Purpose: Recently the MuCap experiment has measured the capture rate ΛS from the singlet pµ atom, employing a low density active target to suppress ppµ formation (PRL 110, 12504 (2013)). Nevertheless, given the
unprecedented precision of this experiment, the existing experimental knowledge in λppµ had to be improved.
Method: The MuCap experiment derived the weak capture rate from the muon disappearance rate in ultra-pure
hydrogen. By doping the hydrogen with 20 ppm of argon, a competing process to ppµ formation was introduced,
which allowed the extraction of λppµ from the observed time distribution of decay electrons.
Results: The ppµ formation rate was measured as λppµ = (2.01 ± 0.06stat ± 0.03sys ) × 106 s−1 . This result
updates the λppµ value used in the above mentioned MuCap publication.
Conclusions: The 2.5× higher precision compared to earlier experiments and the fact that the measurement
was performed at nearly identical conditions to the main data taking, reduces the uncertainty induced by λppµ
to a minor contribution to the overall uncertainty of ΛS and gP , as determined in MuCap. Our final value for
λppµ shifts ΛS and gP by less than one tenth of their respective uncertainties compared to our results published
earlier.
PACS numbers: 23.40.-s, 13.60.-r, 14.20.Dh, 24.80.+y, 29.40.Gx, 36.10.-k
I.
INTRODUCTION
Nuclear muon capture on the proton,
µ− + p → n + νµ ,
(1)
is a basic charged-current weak reaction [1–3]. Several
experiments have measured the rate of ordinary muon
∗
†
‡
§
¶
∗∗
Present address: Los Alamos National Laboratory, Los Alamos,
NM 87545, USA
Deceased
Present address:
Fermi National Accelerator Laboratory,
Batavia, IL 60510, USA
Present address: University of Wisconsin, Madison, WI 53706,
USA
Present address: Brookhaven National Laboratory, Upton, NY
11973, USA
Corresponding author; [email protected]; Present address: Argonne National Laboratory, Lemont, IL 60439, USA
capture (Eq. (1)) or the rarer process of radiative muon
capture, µ+p → n+ν +γ, in order to determine the weak
pseudoscalar coupling of the proton, gP , which can be extracted most straightforwardly from muon capture on the
nucleon. A precision determination of gP has been a longstanding experimental challenge [2, 3] due to the small
rate of capture on the proton and complications arising
from the formation of muonic molecules. The most recent
MuCap result, gP = 8.06 ± 0.55 [4], achieved an unprecedented precision of 7 %, thereby providing a sensitive test
of QCD symmetries and confirming a fundamental prediction of chiral perturbation theory, gP = 8.26±0.23 [5–
7].
Experimentally, process (1) is observed after lowenergy muons are stopped in hydrogen, where they form
pµ atoms and ppµ molecules. The overlap in the wavefunctions of the proton and the bound muon leads to
small but observable capture rates at the 10−3 level rel-
Polarimetry: From the Sun to Stars and Stellar Environments
Proceedings IAU Symposium No. 305, 2015
c 2015 International Astronomical Union
K.N. Nagendra, S. Bagnulo,
R. Centeno, & M. Mart´ınez Gonz´
alez, eds.
DOI: 00.0000/X000000000000000X
Preliminary design of the full-Stokes UV and
visible spectropolarimeter for UVMag/Arago
Martin Pertenais1,2 , Coralie Neiner1 , Laurent Par`
es2,3 , Pascal Petit2,3 ,
4
Frans Snik and Gerard van Harten4
arXiv:1502.00856v1 [astro-ph.IM] 3 Feb 2015
1
LESIA, Observatoire de Paris, CNRS UMR 8109, UPMC, Universit´e de Paris-Diderot,
5 place Jules Janssen, 92100 Meudon, France
email: [email protected]
2
Universit´e de Toulouse; UPS-OMP; IRAP Toulouse, France
3
CNRS; IRAP ; 14 avenue Edouard Belin, 31400 Toulouse, France
4
Leiden Observatory, Leiden University, P.O. Box 9513, 2300 RA Leiden, The Netherlands
Abstract. The UVMag consortium proposed the space mission project Arago to ESA at its
M4 call. It is dedicated to the study of the dynamic 3D environment of stars and planets.
This space mission will be equipped with a high-resolution spectropolarimeter working from
119 to 888 nm. A preliminary optical design of the whole instrument has been prepared and
is presented here. The design consists of the telescope, the instrument itself, and the focusing
optics. Considering not only the scientific requirements, but also the cost and size constraints
to fit a M-size mission, the telescope has a 1.3 m diameter primary mirror and is a classical
Cassegrain-type telescope that allows a polarization-free focus. The polarimeter is placed at this
Cassegrain focus. This is the key element of the mission and the most challenging to be designed.
The main challenge lies in the huge spectral range offered by the instrument; the polarimeter
has to deliver the full Stokes vector with a high precision from the FUV (119 nm) to the NIR
(888 nm). The polarimeter module is then followed by a high-resolution echelle-spectrometer
achieving a resolution of 35000 in the visible range and 25000 in the UV. The two channels are
separated after the echelle grating, allowing a specific cross-dispersion and focusing optics for
the UV and visible ranges. Considering the large field of view and the high numerical aperture,
the focusing optic for both the UV and visible channels is a Three-Mirror-Anastigmat (TMA)
telescope, in order to focus the various wavelengths and many orders onto the detectors.
Keywords. polarization, instrumentation: polarimeters, instrumentation: spectrographs, techniques: polarimetric, stars: magnetic fields, ultraviolet: stars
1. Introduction
The Arago space mission is an ambitious project with a 1.3 m diameter telescope
dedicated to spectropolarimetry in the UV and in the visible. The measurement of spectra
of all types of stars in the UV and visible domains provides important insights into their
formation and evolution. The UV domain is very rich in atomic and molecular lines,
it also contains the signatures of the stellar environment (e.g. of the chromosphere)
and most of the flux of hot stars. At the same time, the visible spectrum allows us
to gain information about the surface of the star itself. For the first time, 3D maps of
stars and their environment will be reconstructed thanks to simultaneous UV and visible
spectroscopy over a full stellar rotation period. The capabilities of extracting information
on stellar magnetospheres, winds, disks, and magnetic fields will be multiplied tenfold
by adding polarimetric power to the spectrograph. Arago will also be able to measure
the changes in the stellar UV radiation and magnetic activity, allowing the study of the
interaction between stars and their planets, in particular magnetospheric interactions
1
Nonlinear X-ray Compton Scattering
Matthias Fuchs1,2, Mariano Trigo2,3, Jian Chen2,3, Shambhu Ghimire2,3, Sharon Shwartz4,
Michael Kozina2,3, Mason Jiang2,3, Thomas Henighan2,3, Crystal Bray2,3, Georges
Ndabashimiye2, P. H. Bucksbaum2,3,7, Yiping Feng5, Sven Herrmann6, Gabriella Carini6,
Jack Pines6, Philip Hart6, Christopher Kenney6, Serge Guillet5, Sébastien Boutet5, Garth
Williams5, Marc Messerschmidt5, Marvin Seibert5, Stefan Möller5, Jerome B. Hastings5,
David A. Reis2,3,7
1
Department of Physics and Astronomy, University of Nebraska - Lincoln, Lincoln, NE
68588, USA
2
Stanford PULSE Institute, SLAC National Accelerator Laboratory, Menlo Park, CA
94025, USA
3
Stanford Institute for Materials and Energy Sciences, SLAC National Accelerator
Laboratory, Menlo Park, CA 94025, USA
4
Physics Department and Institute of Nanotechnology, Bar Ilan University, Ramat Gan,
52900 Israel
5
Linac Coherent Light Source, SLAC National Accelerator Laboratory, 2575 Sand Hill
Road, Menlo Park, California 94025, USA
6
SLAC National Accelerator Laboratory, Menlo Park, CA 94025, USA
7
Department of Photon Science and Applied Physics, Stanford University, Stanford, CA
94305, USA
X-ray scattering is typically used as a weak linear probe of matter. It is primarily
sensitive to the spatial position of electrons and their momentum distribution1,2 .
Elastic X-ray scattering forms the basis of atomic-scale structural determination3
while inelastic Compton scattering4 is often used as a spectroscopic probe of both
single-particle excitations and collective modes5. X-ray free-electron lasers (XFELs)
are unique tools for studying matter on its natural time and length scales due to
their bright and coherent ultrashort X-ray pulses. However, in the focus of an
XFEL the assumption of a weak linear probe breaks down, and nonlinear lightmatter interactions can become ubiquitous, as recently demonstrated6-15. The
electromagnetic field can be sufficiently high that even non-resonant multiphoton
interactions at hard X-rays wavelengths become relevant. Here we report the
observation of one of the most fundamental nonlinear X-ray - matter interactions,
the simultaneous Compton scattering of two identical photons producing a single
higher-energy photon at nearly twice the photon energy. We measure scattered
photons with an energy near 18 keV generated from solid beryllium irradiated by
8.8–9.75 keV XFEL pulses. The intensity in the X-ray focus reaches up to 4x1020
Watt/cm2, which corresponds to a peak electric field two orders of magnitude higher
than the atomic unit of field-strength and within four orders of magnitude of the
quantum electrodynamic critical field16,17. The observed signal is well above
background. It scales quadratically in intensity and is emitted into a non-dipolar
pattern, consistent with the simultaneous two-photon scattering from a collection of
free electrons. However, the energy of the generated photons shows an anomalously
large redshift only present at high intensities. This indicates that the instantaneous
high-intensity scattering effectively interacts with a different electron momentum
distribution than linear Compton scattering, with implications for the study of
atomic-scale structure and dynamics of matter.
In linear Compton scattering a photon transfers energy and momentum to an electron
during the scattering process. The spectrum of scattered photons at a given momentum
transfer is a direct probe of a material’s electron momentum distribution when the energy
loss is large compared to the relevant binding energies (the impulse approximation)18. At
the focus of X-ray free-electron lasers, nonlinear X-ray matter interactions can become
2 Quantum enhanced estimation of optical detector efficiencies
Marco Barbieri,1, 2 Animesh Datta,2 Tim J. Bartley,2, 3 Xian-Min Jin,2, 4 W. Steven Kolthammer,2 and Ian A. Walmsley2
arXiv:1502.00681v1 [quant-ph] 2 Feb 2015
1
Dipartimento di Scienze, Universit`a degli Studi Roma Tre, Via della Vasca Navale 84, 00146, Rome, Italy
2
Clarendon Laboratory, University of Oxford, Parks Road, OX1 3PU, Oxford, UK
3
National Institute of Standards and Technology, 325 Broadway, Boulder, CO 80303, USA
4
Department of Physics and Astronomy, Shanghai Jiao Tong University, Shanghai 200240, China
Quantum mechanics establishes the ultimate limit to the scaling of the precision on any parameter, by identifying optimal probe states and measurements. While this paradigm is, at least in principle, adequate for the
metrology of quantum channels involving the estimation of phase and loss parameters, we show that estimating the loss parameters associated with a quantum channel and a realistic quantum detector are fundamentally
different. While Fock states are provably optimal for the former, we identify a crossover in the nature of the
optimal probe state for estimating detector imperfections as a function of the loss parameter. We provide explicit
results for on-off and homodyne detectors, the most widely used detectors in quantum photonics technologies.
I.
INTRODUCTION
Achieving measurements that approach the ultimate limits to precision is a fundamental, and stimulating, challenge
for experimental sciences, and efforts might be rewarded with
glimpses of novel effects or solid confirmation of existing
paradigms at unprecedented scales. Quantum mechanics dictates the essential limits to measurement, and it is often the
case that ultimate precision is achieved by preparing a measurement probe in a state exhibiting genuine quantum features, such as entanglement [1–5] , and squeezing [6, 7]. Optical metrology has shown particular promise in this regard;
obtaining a quantum advantage in the precision of phase estimation is within reach with some improvements in present
technology [8–10]. Concepts from quantum parameter estimation readily apply to quantum processes. A familiar example is given by the estimation of a lossy bosonic channel,
which could describe, for example, an optical transmission
line [12–14]. The relevant parameter, the transmissivity of the
channel, can be estimated with quantum limited uncertainty
by using Gaussian probe states, such as coherent states, and
photon counting [12, 15].
A key aspect of quantum estimation is the characterization
of a measurement apparatus itself. Increasingly efficient particle detectors are being leveraged in a wide range of fields,
from observational astronomy to quantum information science, and the task of precisely determining the efficiency of
these detectors is necessary for their optimal use [17, 18]. Inefficient detectors can be modelled as a channel loss followed
by a perfect detector; such a description is commonplace in
the analysis of quantum optical experiments [16], and applies
to detectors of electromagnetic radiation in general.
When estimating the inefficiency of a detector, the POVM
is fixed to that implemented by the ideal detector. This is fundamentally different from the estimation of any other parameter, where one is free to choose the measurement. This restriction in the choice of the measurement may in general prevent
the attainment of the quantum Cram´er-Rao bound [19, 20].
This makes the estimation of the efficiency of a detector
conceptually different from other quantum-limited estimation
problems.
In this paper, we analyse the use of Fock states as resource
for the estimation of detector efficiency applied to the two
most common optical detection apparatus: on-off counting
and homodyne detection. While both play a vital role in quantum optics experiments, the former has also appeal in astronomical observations. Our main finding is the existence of regions in which Fock states provide superior performances to
coherent states normally used for calibration. Crucially, this
is in the region of very high efficiencies where the forefront of
detector development lies. We also stress the advantages that
Fock states are able to provide with respect to the case of a
lossy channel estimation. Our analysis of the performance of
quantum light in estimating detector efficiencies with a higher
precision concerns those researchers looking for applications
of these detectors in their studies. These include, for instance,
the use of single-photon correlation measurement of celestial
bodies in astrometry, and the employ of single-photon detection for precise imaging in biology.
II.
ON-OFF DETECTORS
An extensively adopted detection scheme uses on/off
(Geiger-mode) detectors, which are able to tell the presence of
photons in the measured field, but can not determine the number of photons. Limitations to the detection efficiency arise
from the physical process of light absorption and conversion
of the energy into an electrical signal, as well as any loss on
the detection setup. Here we address the task of estimating
detection efficiency, and show an advantage of quantum optical probe states over classical probes with the same energy.
As we show, this advantage is rapidly lost in the presence of
dark counts.
We perform our analysis in terms of the Fisher informa2
tion [19, 20] F (η)= x (∂η p(x; η)) /p(x; η) for the efficiency η, in which the sum is over each possible detector
outcome x occuring with probability p(x; η). The Fisher information bounds the uncertainty, quantified by the variance
∆2 η, attained after M repetitions of the experiment through
the Cram´er-Rao bound ∆2 η ≥ 1/(M F (η)) [20]. Since
the detection scheme is fixed by definition in our problem,
there is no possibility of invoking the quantum Cram´er-Rao
bound, i.e. the Cram´er-Rao optimised over all possible mea-
arXiv:1502.00984v1 [nucl-ex] 3 Feb 2015
International Journal of Modern Physics: Conference Series
c The Authors
Studies of Transverse-Momentum-Dependent distributions with
A Fixed-Target ExpeRiment using the LHC beams (AFTER@LHC)
L. Massacrier1,2,∗, M. Anselmino3 , R. Arnaldi3 , S. J. Brodsky4 , V. Chambert2 ,
W. den Dunnen5 , J.P. Didelez2 , B. Genolini2 , E.G. Ferreiro6 , F. Fleuret7 , Y. Gao8 ,
C. Hadjidakis2 , I. Hˇrivn´
aˇ
cov´
a2 , J.P. Lansberg2 , C. Lorc´
e4,9 , R. Mikkelsen10 , C. Pisano11 ,
A. Rakotozafindrabe12 , P. Rosier2 , I. Schienbein13 , M. Schlegel5 , E. Scomparin4 , B. Trzeciak14 ,
U.I. Uggerhøj10 , R. Ulrich15 , and Z. Yang8 .
1 LAL, Universit´
e Paris-Sud, CNRS/IN2P3, Orsay, France
IPNO, Universit´
e Paris-Sud, CNRS/IN2P3, F-91406, Orsay, France
3 Dip. di Fisica and INFN Sez. Torino, Via P. Giuria 1, I-10125, Torino, Italy
4 SLAC Nat. Accel. Lab., Theoretical Physics, Stanford U. Menlo Park, CA 94025, USA
5 Institute for Theoretical Physics, T¨
ubingen U., D-72076 T¨
ubingen, Germany
6 Dpt. de F´
ısica de Part´ıculas, Universidade de Santiago de C., 15782 Santiago de C., Spain
7 Laboratoire Leprince Ringuet, Ecole
´
Polytechnique, CNRS/IN2P3, 91128 Palaiseau, France
8 CHEP, Department of Engineering Physics, Tsinghua University, Beijing, China
9 IFPA, AGO Dept., Universit´
e de Li`
ege, Sart-Tilman, 4000 Li`
ege, Belgium
10 Dept. of Physics and Astronomy, University of Aarhus, Denmark
11 Nikhef & Dept. of Phys. & Astr., VU Amsterdam, NL-1081 HV Amsterdam, The Netherlands
12 IRFU/SPhN, CEA Saclay, 91191 Gif-sur-Yvette Cedex, France
13 LPSC, Univerist´
e Joseph Fourier, CNRS/IN2P3/INPG, F-38026 Grenoble, France
14 FNSPE, Czech Technical U., Prague, Czech Republic
15 Institut f¨
ur Kernphysik, Karlsruhe Institute of Technology (KIT), 76021 Karlsruhe, Germany
2
We report on the studies of Transverse-Momentum-Dependent distributions (TMDs)
at a future fixed-target experiment –AFTER@LHC– using the p+ or Pb ion LHC
beams, which would be the most energetic fixed-target experiment ever performed.
AFTER@LHC opens new domains of particle and nuclear physics by complementing
collider-mode experiments, in particular those of RHIC and the EIC projects. Both
with an extracted beam by a bent crystal or with an internal gas target, the luminosity
achieved by AFTER@LHC surpasses that of RHIC by up to 3 orders of magnitude. With
an unpolarised target, it allows for measurements of TMDs such as the Boer-Mulders
quark distributions and the distribution of unpolarised and linearly polarised gluons in
unpolarised protons. Using polarised targets, one can access the quark and gluon Sivers
TMDs through single transverse-spin asymmetries in Drell-Yan and quarkonium production. In terms of kinematics, the fixed-target mode combined with a detector covering
ηlab ∈ [1, 5] allows one to measure these asymmetries at large x↑ in the polarised nucleon.
Keywords: TMD; AFTER@LHC; single spin asymmetry; Drell-Yan; quarkonium
PACS numbers:07.90.+c, 21.10.Hw, 13.20.Gd
∗ contact
e-mail: [email protected]
This is an Open Access article published by World Scientific Publishing Company. It is distributed
under the terms of the Creative Commons Attribution 3.0 (CC-BY) License. Further distribution
of this work is permitted, provided the original work is properly cited.
1
arXiv:1502.00932v1 [stat.AP] 3 Feb 2015
Density Estimation Trees in High Energy Physics
Lucio Anderlini
Istituto Nazionale di Fisica Nucleare – Sesto Fiorentino, Firenze
January 30th, 2015
Abstract
Density Estimation Trees can play an important role in exploratory data analysis for multi-dimensional, multi-modal data models of large samples. I briefly
discuss the algorithm, a self-optimization technique based on kernel density estimation, and some applications in High Energy Physics.
1
Introduction
The usage of nonparametric density estimation techniques has seen a quick growth
in the latest years both in High Energy Physics (HEP) and in other fields of Science
dealing with multi-variate data samples. Indeed, the improvement in the computing
resources available for data analysis allows today to process a much larger number of
entries requiring more accurate statistical models. Avoiding parametrization for the
distribution with respect to one or more variables allows to enhance accuracy removing unphysical constraints on the shape of the distribution. The improvement becomes
more evident when considering the joint probability density function with respect to
correlated variables, for whose model a too large number of parameters would be required.
Kernel Density Estimation (KDE) is a nonparametric density estimation technique
based on the estimator
fˆKDE (x) =
1
Ntot
Ntot
k(x − xi ),
(1)
i=1
where x = (x(1) , x(2) , ..., x(d) ) is the vector of coordinates of the d-variate space S
describing the data sample of Ntot entries, k is a normalized function referred to as
kernel. KDE is widely used in HEP Cranmer (2001); Poluektov (2014) including notable applications to the Higgs boson mass measurement by the ATLAS Collaboration
Aad et al. (2014). The variables considered in the construction of the data-model are
the mass of the Higgs boson candidate and the response of a Boosted Decision Tree
(BDT) algorithm used to classify the data entries as Signal or Background candidates
Breiman et al. (1984). This solution allows to synthesize a set of variables, input of the
1
Exploiting the Symmetries of P and S wave for B → K ∗ µ+ µ−
Lars Hofer and Joaquim Matias
arXiv:1502.00920v1 [hep-ph] 3 Feb 2015
Universitat Aut`
onoma de Barcelona, 08193 Bellaterra, Barcelona
After summarizing the current theoretical status of the four-body decay B → K ∗ (→ Kπ)µ+ µ− ,
we apply the formalism of spin-symmetries to the full angular distribution, including the S-wave part
involving a broad scalar resonance K0∗ . While we recover in the P-wave sector the known relation
()
between the angular observables Pi , we find in the S-wave sector two new relations connecting
the coefficients of the S-wave angular distribution and reducing the number of independent S-wave
observables from six to four. Included in the experimental data analysis, these relations can help to
reduce the background from S-wave pollution. We further point out the discriminative power of the
maximum of the angular observable P2 as a charm-loop insensitive probe of right-handed currents.
Moreover, we show that in absence of right-handed currents the angular observables P4 and P5 fulfill
the relation P4 = βP5 at the position where P2 reaches its maximum.
I.
INTRODUCTION
Rare B decays constitute one of the cornerstones in the search for physics beyond the Standard Model (SM). Among
them, the semileptonic mode B → K ∗ (→ Kπ)µ+ µ− represents a particularly interesting channel as the measurement
of the 4-body angular distribution provides a plethora of information which can be used to probe and discriminate
different scenarios of New Physics (NP). In 2013, LHCb presented results of the measurement of an optimized set
()
{Pi } of angular observables [1–5] based on 1 fb−1 data. These observables are constructed in such a way that, to
leading order in the strong coupling constant αs and in the large-recoil expansion, non-perturbative form factors cancel
in the region of low squared invariant mass q 2 of the dilepton pair, a unique and powerful feature in the hadronic
environment.
Experimental data showed several interesting tensions with respect to SM expectations [6]: Most striking is the 4 σ
anomaly1 encountered in the observable P5 [4] in the bin [4.3, 8.68] GeV2 . The observable P2 [2, 3] further displayed
a 2.9 σ deviation in the q 2 -bin [2, 4.3] GeV2 . The position of its zero (q02 = 4.9 ± 0.9 GeV2 ), which is identical to
the zero of the forward-backward asymmetry AFB , is in agreement with the SM prediction q02 4 GeV2 but allows
for higher values. It is remarkable that all these deviations point to the same negative NP contribution C9NP to the
Wilson coefficient of the semileptonic operator O9 , possibly accompanied by a NP contribution C7NP to the Wilson
coefficient of the magnetic operator O7 . New Physics contributions to the Wilson coefficient C10 , and, in particular,
to the coefficients C7,9,10 of the chirality-flipped operators are consistent with zero already at 1 σ. The full pattern,
first pointed out in Ref. [6] and obtained using all available experimental bins in B → K ∗ µ+ µ− together with data
on B → K ∗ γ, B → Xs γ, B → Xs µ+ µ− and Bs → µ+ µ− , is given by the 1 σ ranges
C9NP ∈ [−1.6, −0.9],
C9NP ∈ [−0.2, 0.8],
C7NP ∈ [−0.05, −0.01],
C7NP ∈ [−0.04, 0.02],
NP
C10
∈ [−0.4, 1.0],
NP
C10 ∈ [−0.4, 0.4],
(1)
NP
where the mild preference for a positive C10
is mainly driven by Bs → µ+ µ− data.
The large negative NP contribution to C9 was independently confirmed later on by other groups, using different observables Si [9, 10] (relying on the single large-recoil bin [1,6] GeV2 and low recoil data), different statistical approaches
[11] or form factor input from lattice [12]. Although it had been shown in Refs. [6, 13] that a large C9NP + C9 < 0 was
preferred in order to explain the P5 anomaly, the possibility of a substantial positive C9 enforcing C9NP + C9 ∼ 0 was
discussed in Refs. [10, 14], driven mainly by the 1 fb−1 data [15] on the charged B decay B + → K + µ+ µ− in the region
of low hadronic recoil. The situation has become more coherent recently as the latest 3 fb−1 data on B + → K + µ+ µ−
and B 0 → K 0 µ+ µ− provided by LHCb [16] is also in good agreement with the solution C9N P + C9 < 0 [6], both in
the region of large as well as low hadronic recoil [17, 18]. The three modes thus seem to point to a consistent overall
1
In Ref. [7] this discrepancy is quoted as a 3.7 σ tension between the experimental result and the 68.3% confidence level of the theoretical
prediction, while we have quoted the tension between the experimental result and the theoretical central value. Note also that using the
updated predictions [8] for all observables, including parametric and form factor errors, factorizable power corrections together with an
estimate of non-factorizable ones and charm-loop effects, the tensions with data, albeit slightly reduced, are still clearly present.
Leptophobic Z ′ in Models with Multiple Higgs Doublet Fields
Cheng-Wei Chiang,1, 2, 3, 4, ∗ Takaaki Nomura,5, † and Kei Yagyu1, 6, ‡
1 Department
of Physics and Center for Mathematics and Theoretical Physics,
National Central University, Taoyuan 32001, Taiwan
arXiv:1502.00855v1 [hep-ph] 3 Feb 2015
2 Institute
3 Physics
of Physics, Academia Sinica, Taipei 11529, Taiwan
Division, National Center for Theoretical Sciences, Hsinchu 30013, Taiwan
4 Kobayashi-Maskawa
Institute for the Origin of Particles and the Universe,
Nagoya University, Nagoya 464-8602, Japan
5 Department
of Physics, National Cheng Kung University, Tainan 70101, Taiwan
6 School
of Physics and Astronomy, University of Southampton,
Southampton, SO17 1BJ, United Kingdom
Abstract
We study the collider phenomenology of the leptophobic Z ′ boson from an extra U (1)′ gauge symmetry in
models with N -Higgs doublet fields. We assume that the Z ′ boson at tree level has (i) no Z-Z ′ mixing, (ii)
no interaction with charged leptons, and (iii) no flavour-changing neutral current. Under such a setup, it is
shown that in the N = 1 case, all the U (1)′ charges of left-handed quark doublets and right-handed up- and
down-type quarks are required to be the same, while in the N ≥ 3 case one can take different charges for the
three types of quarks. The N = 2 case is not well-defined under the above three requirements. We study
the pp → Z ′ V → b¯bV processes (V = γ, Z and W ± ) with the leptonic decays of Z and W ± at the LHC.
The most promising discovery channel or the most stringent constraint on the U (1)′ gauge coupling constant
comes from the Z ′ γ process below the tt¯ threshold and from the tt¯ process above the threshold. Assuming
the collision energy of 8 TeV and the integrated luminosity of 19.6 fb−1 , we find that the constraint from
the Z ′ γ search in the lower mass regime can be stronger than that from the UA2 experiment. In the N ≥ 3
case, we consider four benchmark points for the Z ′ couplings with quarks. It is noted that if such a Z ′
is discovered, a careful comparison between the Z ′ γ and Z ′ W signals is crucial to reveal the nature of Z ′
couplings with quarks. We also present the discovery reach of the Z ′ boson at the 14-TeV LHC in both
N = 1 and N ≥ 3 cases.
PACS numbers: 12.60.Cn, 14.70.Hp
∗
Electronic address: [email protected]
Electronic address: [email protected]
‡
Electronic address: [email protected]
†
1
Light axial vector mesons
Kan Chen1,2 ,∗ Cheng-Qun Pang1,2 ,† Xiang Liu1,2‡ ,§ and Takayuki Matsuki3,4,1¶
1
arXiv:1501.07766v1 [hep-ph] 30 Jan 2015
2
School of Physical Science and Technology, Lanzhou University, Lanzhou 730000, China
Research Center for Hadron and CSR Physics, Lanzhou University & Institute of Modern Physics of CAS, Lanzhou 730000, China
3
Tokyo Kasei University, 1-18-1 Kaga, Itabashi, Tokyo 173-8602, Japan
4
Theoretical Research Division, Nishina Center, RIKEN, Saitama 351-0198, Japan
Inspired by the abundant experimental observation of axial vector states, we study whether the observed
axial vector states can be categorized into the conventional axial vector meson family. In this paper we carry
out analysis based on the mass spectra and two-body Okubo-Zweig-Iizuka-allowed decays. Besides testing
the possible axial vector meson assignments, we also predict abundant information for their decays and the
properties of some missing axial vector mesons, which are valuable to further experimental exploration of the
observed and predicted axial vector mesons.
PACS numbers: 14.40.Be, 12.38.Lg, 13.25.Jx
I.
INTRODUCTION
Among the light unflavored mesons listed in Particle Data
Group (PDG) [1], there are abundant light axial vector mesons
with a spin-parity quantum number J P = 1+ , which form a Pwave meson family. Usually, we adopt h1 , b1 , f1 , and a1 to
express the corresponding states with the quantum numbers,
I G (J PC ) = 0− (1+− ), 1+ (1+− ), 0+ (1++ ), and 1− (1++ ), respectively. In Table I, we collect the experimental information of
the observed h1 , b1 , f1 , and a1 states with resonance parameters. In the next section, we will briefly review their experimental and theoretical status.
TABLE I: Resonance parameters of the axial vector states collected
in PDG [1]. The mass and width are average values taken from PDG.
The states omitted from PDG summary table are marked by a superscript ♮, while the states listed as further states in PDG are marked by
a superscript ♭.
I G (J PC )
a1 (1260)
a1 (1640)
−
++
1 (1 ) a1 (1930)
♮
♭
♭
Mass (MeV)
Width (MeV)
1230 ± 40
250 ∼ 600
1647 ± 22
254 ± 27
1930+30
−70
155 ± 45
a1 (2095) 2096 ± 17 ± 121 451 ± 41 ± 81
Facing so many axial vector states in PDG, we need to examine whether all these states can be categorized into the axial vector meson family, which is crucial to reveal their underlying structures. In this work, we carry out a systematic
study of the axial vector states by analyzing mass spectra and
Okubo-Zweig-Iizuka (OZI)-allowed two-body strong decay
behaviors. Comparing our numerical results with the experimental data, we can further test the possible assignments of
the states in the axial vector meson family. In addition, information of the predicted decays of the axial vector states
observed or still missing in experiment is valuable to further
experimental study of axial vector meson.
This paper is organized as follows. After Introduction, we
give a concise review of the research status of the observed
axial vector states in Sec. II. In Sec. III, we present the phenomenological analysis by combining our theoretical results
with the corresponding experimental data, where the Regge
trajectory analysis is adopted to study mass spectra of the axial vector meson family and the quark pair creation (QPC)
model is applied to calculate their OZI-allowed strong decay
behavior. Finally, a short summary is given in Sec. IV.
State
a1 (2270)♭
2270+55
−40
305+70
−40
b1 (1235)
1229.5 ± 3.2
142 ± 9
1960 ± 35
345 ± 75
2240 ± 35
320 ± 85
f1 (1285)
1282.1 ± 0.6
24.2 ± 1.1
f1 (1420)
1426.4 ± 0.9
54.9 ± 2.6
1 (1 ) b1 (1960)
♭
b1 (2240)
♭
+
+
+−
++
♮
1518 ± 5
73 ± 25
f1 (2310)♭
1971 ± 15
240 ± 45
2310 ± 60
255 ± 70
h1 (1170)
1170 ± 20
360 ± 40
1386 ± 19
91 ± 30
0 (1 ) f1 (1510)
f1 (1970)♭
h1 (1380)♮
0− (1+− ) h1 (1595)♮
h1 (1965)♭
h1 (2215)
II.
♭
1594 ± 15+10
−60
384 ± 60+70
100
1965 ± 45
345 ± 75
2215 ± 40
325 ± 55
EEXPERIMENTAL STATUS OF OBSERVED AXIAL
VECTOR STATES
A. a1 states
‡ Corresponding author
∗ Electronic address: chenk˙[email protected]
† Electronic
address: [email protected]
address: [email protected]
¶ Electronic address: [email protected]
§ Electronic
The a1 (1260) was first predicted in Ref. [2]. Later, the
Amsterdam-CERN-Nijmegen-Oxford Collaboration reported
a spin-parity J P = 1+ ρπ enhancement with mass 1040 MeV
FERMILAB-PUB-15-029-E
arXiv:1502.00967v1 [hep-ex] 3 Feb 2015
Tevatron Constraints on Models of the Higgs Boson with Exotic Spin and Parity
Using Decays to Bottom-Antibottom Quark Pairs
T. Aaltonen † ,21 V.M. Abazov ‡ ,13 B. Abbott ‡ ,116 B.S. Acharya ‡ ,80 M. Adams ‡ ,98 T. Adams ‡ ,97 J.P. Agnew ‡ ,94
G.D. Alexeev ‡ ,13 G. Alkhazov ‡ ,88 A. Alton ‡jj ,31 S. Amerio †yy ,39 D. Amidei † ,31 A. Anastassov †w ,15 A. Annovi
† 17
, J. Antos † ,12 G. Apollinari † ,15 J.A. Appel † ,15 T. Arisawa † ,52 A. Artikov † ,13 J. Asaadi † ,47 W. Ashmanskas
† 15
, A. Askew ‡ ,97 S. Atkins ‡ ,106 B. Auerbach † ,2 K. Augsten ‡ ,62 A. Aurisano † ,47 C. Avila ‡ ,60 F. Azfar † ,38
F. Badaud ‡ ,65 W. Badgett † ,15 T. Bae † ,25 L. Bagby ‡ ,15 B. Baldin ‡ ,15 D.V. Bandurin ‡ ,122 S. Banerjee ‡ ,80
A. Barbaro-Galtieri † ,26 E. Barberis ‡ ,107 P. Baringer ‡ ,105 V.E. Barnes † ,43 B.A. Barnett † ,23 P. Barria †aaa ,41
J.F. Bartlett ‡ ,15 P. Bartos † ,12 U. Bassler ‡ ,70 M. Bauce †yy ,39 V. Bazterra ‡ ,98 A. Bean ‡ ,105 F. Bedeschi † ,41
M. Begalli ‡ ,57 S. Behari † ,15 L. Bellantoni ‡ ,15 G. Bellettini †zz ,41 J. Bellinger † ,54 D. Benjamin † ,14 A. Beretvas † ,15
S.B. Beri ‡ ,78 G. Bernardi ‡ ,69 R. Bernhard ‡ ,74 I. Bertram ‡ ,92 M. Besan¸con ‡ ,70 R. Beuselinck ‡ ,93 P.C. Bhat ‡ ,15
S. Bhatia ‡ ,108 V. Bhatnagar ‡ ,78 A. Bhatti † ,45 K.R. Bland † ,5 G. Blazey ‡ ,99 S. Blessing ‡ ,97 K. Bloom ‡ ,109
B. Blumenfeld † ,23 A. Bocci † ,14 A. Bodek † ,44 A. Boehnlein ‡ ,15 D. Boline ‡ ,113 E.E. Boos ‡ ,86 G. Borissov ‡ ,92
D. Bortoletto † ,43 M. Borysova ‡uu ,91 J. Boudreau † ,42 A. Boveia † ,11 A. Brandt ‡ ,119 O. Brandt ‡ ,75 L. Brigliadori
†xx 6
, R. Brock ‡ ,32 C. Bromberg † ,32 A. Bross ‡ ,15 D. Brown ‡ ,69 E. Brucken † ,21 X.B. Bu ‡ ,15 J. Budagov † ,13
H.S. Budd † ,44 M. Buehler ‡ ,15 V. Buescher ‡ ,76 V. Bunichev ‡ ,86 S. Burdin ‡kk ,92 K. Burkett † ,15 G. Busetto †yy ,39
P. Bussey † ,19 C.P. Buszello ‡ ,90 P. Butti †zz ,41 A. Buzatu † ,19 A. Calamba † ,10 E. Camacho-P´erez ‡ ,83 S. Camarda
† 4
, M. Campanelli † ,28 F. Canelli †dd ,11 B. Carls † ,22 D. Carlsmith † ,54 R. Carosi † ,41 S. Carrillo †l ,16 B. Casal †j ,9
M. Casarsa † ,48 B.C.K. Casey ‡ ,15 H. Castilla-Valdez ‡ ,83 A. Castro †xx ,6 P. Catastini † ,20 S. Caughron ‡ ,32 D. Cauz
†f f f ggg 48
, V. Cavaliere † ,22 A. Cerri †e ,26 L. Cerrito †r ,28 S. Chakrabarti ‡ ,113 K.M. Chan ‡ ,103 A. Chandra ‡ ,121
E. Chapon ‡ ,70 G. Chen ‡ ,105 Y.C. Chen † ,1 M. Chertok † ,7 G. Chiarelli † ,41 G. Chlachidze † ,15 K. Cho † ,25
S.W. Cho ‡ ,82 S. Choi ‡ ,82 D. Chokheli † ,13 B. Choudhary ‡ ,79 S. Cihangir ‡ ,15 D. Claes ‡ ,109 A. Clark † ,18 C. Clarke
† 53
, J. Clutter ‡ ,105 M.E. Convery † ,15 J. Conway † ,7 M. Cooke ‡tt ,15 W.E. Cooper ‡ ,15 M. Corbo †z ,15 M. Corcoran
‡ 121
,
M. Cordelli † ,17 F. Couderc ‡ ,70 M.-C. Cousinou ‡ ,67 C.A. Cox † ,7 D.J. Cox † ,7 M. Cremonesi † ,41 D. Cruz † ,47
J. Cuevas †y ,9 R. Culbertson † ,15 D. Cutts ‡ ,118 A. Das ‡ ,120 N. d’Ascenzo †v ,15 M. Datta †gg ,15 G. Davies ‡ ,93
P. de Barbaro † ,44 S.J. de Jong ‡ ,84, 85 E. De La Cruz-Burelo ‡ ,83 F. D´eliot ‡ ,70 R. Demina ‡ ,44 L. Demortier † ,45
M. Deninno † ,6 D. Denisov ‡ ,15 S.P. Denisov ‡ ,87 M. D’Errico †yy ,39 S. Desai ‡ ,15 C. Deterre ‡ll ,94 K. DeVaughan
‡ 109
,
F. Devoto † ,21 A. Di Canto †zz ,41 B. Di Ruzza †p ,15 H.T. Diehl ‡ ,15 M. Diesburg ‡ ,15 P.F. Ding ‡ ,94
J.R. Dittmann † ,5 A. Dominguez ‡ ,109 S. Donati †zz ,41 M. D’Onofrio † ,27 M. Dorigo †hhh ,48 A. Driutti †f f f ggg ,48
A. Dubey ‡ ,79 L.V. Dudko ‡ ,86 A. Duperrin ‡ ,67 S. Dutt ‡ ,78 M. Eads ‡ ,99 K. Ebina † ,52 R. Edgar † ,31 D. Edmunds
‡ 32
, A. Elagin † ,47 J. Ellison ‡ ,96 V.D. Elvira ‡ ,15 Y. Enari ‡ ,69 R. Erbacher † ,7 S. Errede † ,22 B. Esham † ,22 H. Evans
‡ 101
,
V.N. Evdokimov ‡ ,87 S. Farrington † ,38 A. Faur´e ‡ ,70 L. Feng ‡ ,99 T. Ferbel ‡ ,44 J.P. Fern´
andez Ramos † ,29
F. Fiedler ‡ ,76 R. Field † ,16 F. Filthaut ‡ ,84, 85 W. Fisher ‡ ,32 H.E. Fisk ‡ ,15 G. Flanagan †t ,15 R. Forrest † ,7
M. Fortner ‡ ,99 H. Fox ‡ ,92 M. Franklin † ,20 J.C. Freeman † ,15 H. Frisch † ,11 S. Fuess ‡ ,15 Y. Funakoshi † ,52
C. Galloni †zz ,41 P.H. Garbincius ‡ ,15 A. Garcia-Bellido ‡ ,44 J.A. Garc´ıa-Gonz´alez ‡ ,83 A.F. Garfinkel † ,43 P. Garosi
†aaa 41
, V. Gavrilov ‡ ,33 W. Geng ‡ ,67, 32 C.E. Gerber ‡ ,98 H. Gerberich † ,22 E. Gerchtein † ,15 Y. Gershtein ‡ ,110
S. Giagu † ,46 V. Giakoumopoulou † ,3 K. Gibson † ,42 C.M. Ginsburg † ,15 G. Ginther ‡ ,15, 44 N. Giokaris † ,3
P. Giromini † ,17 V. Glagolev † ,13 D. Glenzinski † ,15 O. Gogota ‡ ,91 M. Gold † ,34 D. Goldin † ,47 A. Golossanov † ,15
G. Golovanov ‡ ,13 G. Gomez † ,9 G. Gomez-Ceballos † ,30 M. Goncharov † ,30 O. Gonz´alez L´
opez † ,29 I. Gorelov † ,34
A.T. Goshaw † ,14 K. Goulianos † ,45 E. Gramellini † ,6 P.D. Grannis ‡ ,113 S. Greder ‡ ,71 H. Greenlee ‡ ,15 G. Grenier
‡ 72
‡ 65
‡ 68
‡ll 70
† 11
† 51, 15
, Ph. Gris , J.-F. Grivaz , A. Grohsjean , C. Grosso-Pilcher , R.C. Group ,
S. Gr¨
unendahl ‡ ,15
M.W. Gr¨
unewald ‡ ,81 T. Guillemin ‡ ,68 J. Guimaraes da Costa † ,20 G. Gutierrez ‡ ,15 P. Gutierrez ‡ ,116 S.R. Hahn
† 15
‡ 117
† 44
‡ 59
† 17
† 49
‡ 94
† 50
, J. Haley ,
J.Y. Han , L. Han , F. Happacher , K. Hara , K. Harder , M. Hare , A. Harel ‡ ,44
R.F. Harr † ,53 T. Harrington-Taber †m ,15 K. Hatakeyama † ,5 J.M. Hauptman ‡ ,104 C. Hays † ,38 J. Hays ‡ ,93
T. Head ‡ ,94 T. Hebbeker ‡ ,73 D. Hedin ‡ ,99 H. Hegab ‡ ,117 J. Heinrich † ,40 A.P. Heinson ‡ ,96 U. Heintz ‡ ,118
C. Hensel ‡ ,56 I. Heredia-De La Cruz ‡mm ,83 M. Herndon † ,54 K. Herner ‡ ,15 G. Hesketh ‡oo ,94 M.D. Hildreth ‡ ,103
5
91
Taras Shevchenko National University of Kyiv, Kiev, Ukraine
92
Lancaster University, Lancaster LA1 4YB, United Kingdom
93
Imperial College London, London SW7 2AZ, United Kingdom
94
The University of Manchester, Manchester M13 9PL, United Kingdom
95
University of Arizona, Tucson, Arizona 85721, USA
96
University of California Riverside, Riverside, California 92521, USA
97
Florida State University, Tallahassee, Florida 32306, USA
98
University of Illinois at Chicago, Chicago, Illinois 60607, USA
99
Northern Illinois University, DeKalb, Illinois 60115, USA
100
Northwestern University, Evanston, Illinois 60208, USA
101
Indiana University, Bloomington, Indiana 47405, USA
102
Purdue University Calumet, Hammond, Indiana 46323, USA
103
University of Notre Dame, Notre Dame, Indiana 46556, USA
104
Iowa State University, Ames, Iowa 50011, USA
105
University of Kansas, Lawrence, Kansas 66045, USA
106
Louisiana Tech University, Ruston, Louisiana 71272, USA
107
Northeastern University, Boston, Massachusetts 02115, USA
108
University of Mississippi, University, Mississippi 38677, USA
109
University of Nebraska, Lincoln, Nebraska 68588, USA
110
Rutgers University, Piscataway, New Jersey 08855, USA
111
Princeton University, Princeton, New Jersey 08544, USA
112
State University of New York, Buffalo, New York 14260, USA
113
State University of New York, Stony Brook, New York 11794, USA
114
Brookhaven National Laboratory, Upton, New York 11973, USA
115
Langston University, Langston, Oklahoma 73050, USA
116
University of Oklahoma, Norman, Oklahoma 73019, USA
117
Oklahoma State University, Stillwater, Oklahoma 74078, USA
118
Brown University, Providence, Rhode Island 02912, USA
119
University of Texas, Arlington, Texas 76019, USA
120
Southern Methodist University, Dallas, Texas 75275, USA
121
Rice University, Houston, Texas 77005, USA
122
University of Virginia, Charlottesville, Virginia 22904, USA
123
University of Washington, Seattle, Washington 98195, USA
(Dated: February 3, 2015)
Combined constraints from the CDF and D0 Collaborations on models of the Higgs boson with
exotic spin J and parity P are presented and compared with results obtained assuming the standard
model value J P = 0+ . Both collaborations analyzed approximately 10 fb−1 of proton-antiproton
collisions with a center-of-mass energy of 1.96 TeV collected at the Fermilab Tevatron. Two models
predicting exotic Higgs bosons with J P = 0− and J P = 2+ are tested. The kinematic properties of
exotic Higgs boson production in association with a vector boson differ from those predicted for the
standard model Higgs boson. Upper limits at the 95% credibility level on the production rates of
the exotic Higgs bosons, expressed as fractions of the standard model Higgs boson production rate,
are set at 0.36 for both the J P = 0− hypothesis and the J P = 2+ hypothesis. If the production
rate times the branching ratio to a bottom-antibottom pair is the same as that predicted for the
standard model Higgs boson, then the exotic bosons are excluded with significances of 5.0 standard
deviations and 4.9 standard deviations for the J P = 0− and J P = 2+ hypotheses, respectively.
PACS numbers: 13.85.Rm, 14.80.Bn, 14.80.Ec
The Higgs boson discovered by the ATLAS [1] and
CMS [2] Collaborations in 2012 using data produced in
proton-proton collisions at the Large Hadron Collider
(LHC) at CERN allows many stringent tests of the electroweak symmetry breaking in the standard model (SM)
and extensions to the SM to be performed. To date,
measurements of the Higgs boson’s mass and width [3–
6], its couplings to other particles [3, 7–11], and its spin
and parity quantum numbers J and P [10–15] are consistent with the expectations for the SM Higgs boson. The
CDF and D0 Collaborations at the Fermilab Tevatron
observed a 3.0 standard deviation (s.d.) excess of events
consistent with a Higgs boson signal, largely driven by
those channels sensitive to the decay of the Higgs boson
to bottom quarks (H → b¯b) [16, 17]. The Tevatron data
are also consistent with the predictions for the properties
of the SM Higgs boson [17–21].
Ref. [22] proposed to use the Tevatron data to test
models for the Higgs boson with exotic spin and parity,
using events in which the exotic Higgs boson X is produced in association with a W or a Z boson and decays to
a bottom-antibottom quark pair, X → b¯b. This proposal
used two of the spin and parity models in Ref. [23], one
with a pseudoscalar J P = 0− state and the other with
arXiv:1502.00985v1 [hep-th] 3 Feb 2015
MITP/15-006
IFT-UAM/CSIC-15-008
Discrete Abelian gauge symmetries and axions
Gabriele Honecker1 and Wieland Staessens2
1
Cluster of Excellence PRISMA & Institute of Physics (WA THEP), Johannes Gutenberg
University, 55099 Mainz, Germany
2
Instituto de F´ısica Te´
orica UAM-CSIC, Cantoblanco, 28049 Madrid, Spain
E-mail:
1
[email protected];
2
[email protected]
Abstract. We combine two popular extensions of beyond the Standard Model physics within
the framework of intersecting D6-brane models: discrete Zn symmetries and Peccei-Quinn
axions. The underlying natural connection between both extensions is formed by the presence
of massive U (1) gauge symmetries in D-brane model building. Global intersecting D6-brane
models on toroidal orbifolds of the type T 6 /Z2N and T 6 /Z2 × Z2M with discrete torsion offer
excellent playgrounds for realizing these extensions. A generation-dependent Z2 symmetry is
identified in a global Pati-Salam model, while global left-right symmetric models give rise to
supersymmetric realizations of the DFSZ axion model. In one class of the latter models, the
axion as well as Standard Model particles carry a non-trivial Z3 charge.
1. Introduction
Extensions of the Standard Model are characterised by the inclusion of new particles to resolve
open questions arising in particle physics and cosmology. These new particles are subject to
novel (gauge) symmetries which constrain the interactions between the observed particles and
the yet to be discovered ones. Obviously the most simple new gauge symmetries to appear in
beyond the Standard Model scenarios are local Abelian symmetries. As it turns out, Abelian
gauge symmetries are ubiquitous in four-dimensional string theory vacua, and within Type II
string theory compactifications they appear in three species:
• closed string vectors associated to isometries of the compact six-dimensional space,
• massless open string vectors with matter charged under them,
• massive open string vectors with masses of the order of Mstring .
It has only been realized recently that the massive vectors do not decouple completely from the
low-energy effective field theory, but that discrete subgroups can survive [1, 2, 3, 4, 5, 6] and act
as ultimate selection rules for non-perturbative instantonic m-point couplings. This observation
complements the discussion of discrete gauge symmetries in field theory invoked to guarantee
e.g. the stability of the proton [7, 8, 9, 10].
On the other hand, the massive Abelian vectors lead to global U (1) symmetries in
perturbation theory, which might be broken explicitly by vacuum expectation values of light
charged open string states. This is exactly the stringy generalization of a field theoretical
Peccei-Quinn symmetry. The relevant charged pseudo-Nambu Goldstone bosons or axions stem
arXiv:1502.00983v1 [physics.gen-ph] 29 Jan 2015
Stability of an electron embedded in Higgs
condensate
ˇ anek ∗
Eugen Sim´
Department of Physics, University of California, Riverside, Calif. 92521
Abstract
We study stability of an electron distributed on the surface of a spherical cavity in Higgs condensate. The surface tension of the cavity prevents
the electron from flying apart due to Coulomb repulsion. A similar model
was introduced by Dirac in 1962, though without reference to Higgs condensate. In his model, the equilibrium radius of the electron equals the
classical electron radius, Rec ≃ 2.8 × 10−13 cm, that is about 105 times the
radius consistent with experimental data. To address this problem, we
replace the Coulomb term in the total energy of the electron by fermion
self-energy involving screening by electrons occupying the negative energies of the vacuum. The tension of the cavity is obtained using the approximation ξ0 ≪ R0 where ξ0 is the coherence length. For ξ0 = 10−3 R0 ,
the equilibrium radius in this model is R0 ≃ 9.2 × 10−32 cm.
For such a small radius, we find the gravitational energy of the electron
to be large enough to cancel the energy h
¯ c/R, coming from the vibrational
zero point energy and the kinetic energy of the embedded electron.
PACS number(s): 12.15.-y, 12.20.-m, 12.39.Ba, 14.60.Cd, 14.80.Bn
1
Introduction
The problem of stability of an electron has an old history. It began with the introduction of the concept of an electron of finite size first proposed by Abraham
[1] and Lorentz [2]. This concept is, however, faced with difficulties. The main
problem comes from the assumption that the mass of the electron is entirely of
electromagnetic origin and that no non-electromagnetic mass exists [1,2]. First,
it is clear that a finite charge distribution cannot be stable under pure electromagnetic forces. There is also a problem with the relativistic transformation
properties of the energy and momentum of the electromagnetic field of the electron since these quantities do not form a 4-vector. These difficulties can only
be overcome by introducing non-electromagnetic stresses [3, 4].
Dirac [5] introduced in 1962 a simple model of an electron which consists
of a charged conducting surface of a cavity in the electromagnetic field. The
∗ electronic
address: [email protected]
1
arXiv:1502.00930v1 [astro-ph.CO] 3 Feb 2015
Studies of inflation and dark energy
with coupled scalar fields
Susan Vu
Submitted for the degree of Doctor of
Philosophy
School of Mathematics and Statistics
September 2014
Supervisor: Prof. Carsten van de Bruck
University of Sheffield
Abstract
Currently there is no definitive description for the accelerated expansion of the Universe at both early and late times; we know these
two periods as the epochs of inflation and dark energy. Contained
within this Thesis are two studies of inflation and one in the context of dark energy. The first study involves two noncanonical kinetic terms each in a two-field scenario, and their effects on the
generation of isocurvature modes. As a result, these terms affect
the isocurvature perturbations produced, and consequently the Cosmic Microwave Background. In the following study, the impact of a
sharp transition upon the effective Planck mass is considered in both
a single-field and two-field model. A feature in the primordial power
spectrum arising from these transitions is found in single-field models, but not for two-field models. The final model discussed is on the
subject of dark energy. A type of nonconformal coupling is examined namely the “disformal” coupling; in this scenario a scalar field
is disformally coupled to matter species. Two consistency checks are
undertaken, the first to provide a fluid description and the second,
a kinetic theory. From this, observables are constructed and used
to create constraints on the individual coupling strengths.
Global surpluses of spin-base invariant fermions
Holger Gies and Stefan Lippoldt
Theoretisch-Physikalisches Institut, Friedrich-Schiller-Universität Jena, Max-Wien-Platz 1, D-07743 Jena, Germany
arXiv:1502.00918v1 [hep-th] 3 Feb 2015
The spin-base invariant formalism of Dirac fermions in curved space maintains the essential symmetries of general covariance as well as similarity transformations of the Clifford algebra. We
emphasize the advantages of the spin-base invariant formalism both from a conceptual as well as
from a practical viewpoint. This suggests that local spin-base invariance should be added to the list
of (effective) properties of (quantum) gravity theories. We find support for this viewpoint by the
explicit construction of a global realization of the Clifford algebra on a 2-sphere which is impossible
in the spin-base non-invariant vielbein formalism.
I.
INTRODUCTION
The mutual interrelation of matter and spacetime
(“matter curves spacetime - spacetime determines the
paths of matter”) is particularly apparent for fermions.
For instance for Dirac fermions, information about both
spin as well as spacetime meets in the Clifford algebra,
{γµ , γν } = 2gµν I,
(1)
where the Dirac matrices γµ as well as the metric gµν generally are spacetime dependent. While many tests of classical gravity rely on vacuum solutions to Einstein’s equation, also many attempts at quantizing gravity primarily
concentrate on the dynamics of spacetime without matter, cf. [1]. This is similar in spirit to “quenched” QCD
which allows to understand already many features of the
strong interactions at the quantum level even quantitatively. Only recently, some evidence has been collected
that the existence of matter degrees of freedom can constrain the existence of certain quantum gravity theories
[2–6]. This is again analogous to QCD where the presence of too many dynamical fermions can destroy the
high-energy completeness of the theory.
The interrelation of gravity and fermions provided by
(1) has also been interpreted in various partly conflicting directions: read from right to left, one is tempted to
conclude that one first needs a spacetime metric gµν in
order to give a meaning to spinorial degrees of freedom
and corresponding physical observables such as currents
¯ µ ψ. On the other hand, representation theory of the
∼ ψγ
Lorentz group in flat space suggests that all nontrivial
representations can be composed out of the fundamental
spinorial representation, culminating into (1) for Dirac
spinors. If so, then also the metric might be a composite
degree of freedom, potentially arising as an expectation
value of composite spinorial operators, see, e.g., [7–9].
As a starting point to disentangle this hen-or-egg problem – spinors or metric first? – we consider the Clifford
algebra (1) as fundamental in this work. We emphasize that this is different from a conventional approach
[10], where one starts from the analogous Clifford algebra in flat (tangential) space, {γ(f) a , γ(f) b } = 2ηab I,
with fixed γ(f) a and then uplifts the Clifford algebra to
curved space with the aid of a vielbein eµa (x), such that
γ(e) µ = eµ a γ(f) a satisfies (1). In addition to diffeomor-
phism invariance, the vielbein approach supports a local
SO(3,1) symmetry of Lorentz transformations in tangential space, i.e. with respect to the roman bein index.
By contrast, the Clifford algebra (1) actually supports a
bigger symmetry of local similarity (spin-base) transformations in addition to general covariance.
Developing a formalism that features this full spin-base
invariance has first been initiated by Schrödinger [11] and
amended with the required spin metric by Bargmann [12]
in 1932. Surprisingly, it has been rarely used in the literature, see, e.g., [13–19], or even reinvented [20]. A full
account of the formalism also including spin torsion has
recently been given in [21]. Particular advantages are
not only the inclusion and generalization of the vielbein
formalism. In a quantized setting, it even justifies the
widespread use of the vielbein as an auxiliary quantity
and not as a fundamental entity. Common quantization
schemes relying on the metric as fundamental degree of
freedom remain applicable also with fermionic matter.
Hence, a Jacobian from the variable transformation to
the vielbein does not have to be accounted for [21].
In this work, we present further advantages of the spinbase invariant formalism and discuss some general aspects in order to elucidate the interplay between diffeomorphisms and spin-base transformations. We point out
various options of defining the spin-base group, differing
by the possible field content of further interactions and
also naturally permitting a Spinc structure. Since the
conventional vielbein formalism can always be recovered
within the spin-base invariant formalism, it is tempting
to think that the latter is merely a technical perhaps overabundant generalization of the former. We demonstrate
that this is not the case by an explicit construction of a
global spin-base on the 2-sphere – a structure which is not
possible in the conventional formalism because of global
obstructions from the Poincaré-Brouwer (hairy-ball) theorem. We believe that this example is paradigmatic for
the surpluses of the spin-base invariant formalism.
II.
GENERAL COVARIANCE AND SPIN-BASE
INVARIANCE
Local symmetries are expected to be fundamental,
since symmetry-breaking perturbations typically contain
On the near-threshold pp
¯ invariant mass spectrum measured in J/ψ and ψ ′ decays
Xian-Wei Kang1 ,∗ Johann Haidenbauer1 ,† and Ulf-G. Meißner1,2‡
arXiv:1502.00880v1 [nucl-th] 3 Feb 2015
1
Institute for Advanced Simulation, J¨
ulich Center for Hadron Physics,
and Institut f¨
ur Kernphysik, Forschungszentrum J¨
ulich, D-52425 J¨
ulich, Germany
2
Helmholtz-Institut f¨
ur Strahlen- und Kernphysik and Bethe Center
for Theoretical Physics, Universit¨
at Bonn, D-53115 Bonn, Germany
A systematic analysis of the near-threshold enhancement in the p¯p invariant mass spectrum seen
in the decay reactions J/ψ → x¯
pp and ψ ′ (3686) → x¯
pp (x = γ, ω, ρ, π, η) is presented. The
¯ N final-state interaction (FSI) and the pertinent FSI
enhancement is assumed to be due to the N
effects are evaluated in an approach that is based on the distorted-wave Born approximation. For the
¯ N interaction a recent potential derived within chiral effective field theory and fitted to results of a
N
partial-wave analysis of p¯p scattering data is considered and, in addition, an older phenomenological
model constructed by the J¨
ulich group. It is shown that the near-threshold spectrum observed in
various decay reactions can be reproduced simultaneously and consistently by our treatment of the
p¯p FSI. It turns out that the interaction in the isospin-1 1 S0 channel required for the description of
¯ N bound state.
the J/ψ → γ p¯p decay predicts a N
I.
INTRODUCTION
The origin of the enhancement in the antiprotonproton (¯
pp) mass spectrum at low invariant masses observed in heavy meson decays like J/ψ → γ p¯p, B → K p¯p
¯ → Dp¯p, but also in the reaction e+ e− ↔ p¯p,
and B
is an interesting and still controversially discussed issue.
In particular, the spectacular near-threshold enhancement in the p¯p invariant mass spectrum for the reaction J/ψ → γ p¯p, first observed in a high-statistics and
high-mass-resolution experiment by the BES Collaboration [1], has led to numerous publications with speculations about the discovery of a new resonance [1] or of a p¯p
bound state (baryonium) [2–4], and was even associated
with exotic glueball states [5–7]. However, in the above
processes the hadronic final-state interaction (FSI) in the
p¯p system should play a role too. Indeed, the group in
J¨
ulich-Bonn [8, 9] but also others [10–17] demonstrated
that the near-threshold enhancement in the p¯p invariant mass spectrum of the reaction J/ψ → γ p¯p could be
simply due to the FSI between the outgoing proton and
antiproton. Specifically, the calculation [8, 9] based on
¯ N ) model [18–
the realistic J¨
ulich antinucleon–nucleon (N
20], the one by the Paris group [15], utilizing the Paris
¯ N model [21], and that of Entem and Fern´
N
andez [14],
¯ N interaction derived from a constituent quark
using a N
model [22], explicitly confirmed the significance of FSI
effects estimated in the initial studies [10–12] within the
effective range approximation.
In the present work we perform a systematic analysis of the near threshold enhancements in the reactions
J/ψ → x¯
pp and ψ ′ (3686) → x¯
pp (x = γ, ω, ρ, π, η) with
emphasis on the role played by the p¯p interaction. The
aim is to achieve a simultaneous and consistent descrip-
∗
†
‡
[email protected]
[email protected]
[email protected]
tion of all p¯p invariant mass spectra measured in the various reactions. FSI effects for different decay channels
cannot be expected to be quantitatively the same. In
particular, with regard to p¯p, the two baryons have to be
in different states if the quantum numbers of the third
particle in the decay channel differ, in accordance with
the general conservation laws. Furthermore, it is possible that dynamical selection rules, reflecting the details
of the reaction mechanism, could suppress the decay into
p¯p S-waves for some decays near threshold. Thus, in different decay modes the final p¯p system can and must be
in different partial waves and, accordingly the FSI effects
will differ too.
As mentioned, initial studies of FSI effects in the decay J/ψ → γ p¯p were done in the rather simplistic effective range approximation. Later investigations, like the
ones performed by us [8, 9], employed directly scatter¯ N potential models. Still
ing amplitudes from realistic N
also here the treatment of the FSI is done within the
so-called Migdal-Watson approach [23, 24] where the elementary decay (or production) amplitude is simply multiplied with the p¯p T -matrix. It is known that this approach works reasonably well for reactions with a final
N N system [25]. In this case the scattering length a is
fairly large, for example, a ≈ −24 fm for a final np system (in the 1 S0 state). Measurements of the level shifts
in antiprotonic hydrogen atoms suggest that the scattering lengths for p¯p scattering are presumably only in the
order of 1 to 2 fm [26]. Moreover, those scattering lengths
are complex due to the presence of annihilation channels.
Therefore, in the present paper we consider an alternative
and more refined approach for taking into account the
FSI. Specifically, we use the Jost function which is calcu¯ N potentials. FSI effects
lated directly from realistic N
are then taken into account by multiplying the reaction
amplitude with the inverse of this Jost function. This is
practically equivalent to a treatment of such decay reactions within a distorted-wave Born approximation. Note
that this is different from the popular Jost-function approach based on the effective range approximation [27]
arXiv:1502.00811v1 [gr-qc] 3 Feb 2015
COSMOLOGY IN TERMS OF THE
DECELERATION PARAMETER. PART I
Yu.L. Bolotin, V.A. Cherkaskiy, O.A. Lemets
D.A. Yerokhin and L.G. Zazunov
”All of observational cosmology is the search
for two numbers: H0 and q0 .”
Allan Sandage, 1970
Abstract
In the early seventies, Alan Sandage defined cosmology as the search for two
numbers: Hubble parameter H0 and deceleration parameter q0 . The first
of the two basic cosmological parameters (the Hubble parameter) describes
the linear part of the time dependence of the scale factor. Treating the
Universe as a dynamical system it is natural to assume that it is non-linear:
indeed, linearity is nothing more than approximation, while non-linearity
represents the generic case. It is evident that future models of the Universe
must take into account different aspects of its evolution. As soon as the
scale factor is the only dynamical variable, the quantities which determine
its time dependence must be essentially present in all aspects of the Universe’
evolution. Basic characteristics of the cosmological evolution, both static and
dynamical, can be expressed in terms of the parameters H0 and q0 . The very
parameters (and higher time derivatives of the scale factor) enable us to
construct model-independent kinematics of the cosmological expansion.
Time dependence of the scale factor reflects main events in history of
the Universe. Moreover it is the deceleration parameter who dictates the
expansion rate of the Hubble sphere and determines the dynamics of the observable galaxy number variation: depending on the sign of the deceleration
parameter this number either grows (in the case of decelerated expansion),
or we are going to stay absolutely alone in the cosmos (if the expansion is
accelerated).
The intended purpose of the report is reflected in its title — ”Cosmology in terms of the deceleration parameter”. We would like to show that
practically any aspect of the cosmological evolution is tightly bound to the
deceleration parameter.
arXiv:1502.00751v1 [hep-th] 3 Feb 2015
Quantization of systems with OSp(2|2) symmetry
Yoshiharu K AWAMURA∗
Department of Physics, Shinshu University,
Matsumoto 390-8621, Japan
February 4, 2015
Abstract
We study the quantization of systems with OSp(2|2) symmetry. Systems contain
ordinary fields and their counterparts with different statistics. The unitarity of systems holds by imposing subsidiary conditions on states.
1 Introduction
The spin-statistics theorem explains that observed particles of integer spin obey BoseEinstein statistics and are quantized by the commutation relations, and those of half odd
integer spin obey Fermi-Dirac statistics and are quantized by the anti-commutation relations in the framework of relativistic quantum field theory [1, 2, 3, 4, 5, 6, 7, 8, 9, 10,
11, 12, 13]. The study on abnormal fields has been little carried out [14, 15, 16, 17], except for Faddeev-Popov ghosts, i.e., ghost fields appearing on the quantization of systems with local symmetries [18]. Here, abnormal fields mean particles obeying different
statistics from ordinary ones. We refer to a scalar field following anti-commutation relations as a ‘fermionic scalar field’ and to a spinor field following commutation relations
as a ‘bosonic spinor field’.
The reasons for the indifference of abnormal fields would be as follows. First, they
seem unrealistic because the standard model does not contain abnormal ones irrelevant to gauge symmetries. Second, in the introduction of abnormal fields, states with
negative norm appear and the unitarity of systems can be violated. Third, even if such
unfavorable states are projected out by imposing subsidiary conditions on states, abnormal fields become unphysical and cannot give any effects on physical processes. Hence,
we suppose that the existence of abnormal fields cannot be verified directly or this is the
same as the non-existence.
Nevertheless, it would be meaningful to examine systems with abnormal fields from
following reasons. There is a possibility that unphysical objects exist in nature if they
are not prohibited from the consistency of theories. This is a similar idea to that Dirac
predicted the existence of magnetic monopole based on quantum theory. Unphysical
∗
E-mail: [email protected]
1
Massless Neutrino Oscillations via Quantum Tunneling
Hai-Long Zhao
Jiuquan satellite launch center, 732750, China
Abstract: In order for different kinds of neutrino to transform into each other, the eigenvalues of energy
of neutrino must be different. In the present theory of neutrino oscillations, this is guaranteed by the mass
differences between the different eigenstates of neutrino. Thus neutrino cannot oscillate if it is massless.
We suggest an explanation for neutrino oscillations by analogy with the oscillation of quantum two-state
system, where the flipping of one state into the other may be regarded as a process of quantum tunneling
and the required energy difference between the two eigenstates comes from the barrier potential energy. So
neutrino with vanishing mass can also oscillate. One of the advantages of the explanation is that neutrino
can still be described with Weyl equation within the framework of standard model.
Keywords: Neutrino oscillations; Hamiltonian matrix; Quantum tunneling
1 Introduction
The experiments with solar, atmospheric and reactor neutrinos have provided compelling
evidences for the existence of neutrino oscillations [1–11]. At present, neutrino oscillations are
considered to be caused by nonzero neutrino masses and neutrino flavour mixing. Neutrino
oscillations include oscillations in matter and in vacuum, and the Seesaw mechanism has been
proposed to explain why the neutrino masses are so small [12, 13]. The formula of oscillation length
can be derived by the present theory [14], but some of the assumptions, such as equal-energy or
equal-momentum of the mass state in the production process of neutrino, are controversial. This has
led to the wave packet description of neutrino by some authors [15–25]. Although there exists dispute
about the description of neutrino as a plane wave approximation, it is still the foundation of the
theoretical analysis of neutrino oscillations. So we mainly discuss the method of plane wave
approximation. It should be noted that there exists an issue whether for the plane wave or the wave
packet description, which will be discussed in the subsequent section.
In order for neutrino to be described within the framework of standard model, we suggest that the
flavour changing of neutrino can be realized by barrier tunneling by analogy with the quantum
two-state system. The eigenvalues of the energy of the neutrino will be different taking into account
the potential energy of the barrier. In this case, the neutrino can oscillate even with a vanishing mass.
2 The present theory of neutrino oscillations
When approximated as a plane wave, neutrino can oscillate via four assumptions: equal–energy,
equal–momentum, energy–momentum conservation and equal–velocity. For a review of these
assumptions one may see [24, 25]. We mainly discuss the former two assumptions, which can be
found in [14]. For the latter two assumptions, one may see [24, 25] and the references therein.
For convenience we work in the natural units, where h = c = 1 . For the sake of simplicity, we leave
the tauon neutrino vτ out of the following and assume that only the electron neutrino v e and muon
neutrino v μ mix with each other. As both v e and v μ are not the eigenstates of the neutrino energy,
we denote the eigenstates of the Hamiltonian with v1 and v 2 , respectively, and for which we make
the following ansatz
1
MSUHEP-150202
Sign Structure of Susceptibilities of Conserved Charges in the (2 + 1) Polyakov Quark
Meson Model
Sandeep Chatterjee∗
Theoretical Physics Division, Variable Energy Cyclotron Centre, Kolkata 700064, India
Kirtimaan A. Mohan†
Department of Physics and Astronomy, Michigan State University, East Lansing, Michigan 48824, USA
arXiv:1502.00648v1 [nucl-th] 2 Feb 2015
Abstract
The sign structure of correlations of conserved charges are investigated in a QCD like model: the (2 + 1) flavor Polyakov
Quark Meson model. We compute all susceptibilities of the conserved charges on the (µB − T ) plane up to fourth order and a
few at higher order as well. By varying the mass of the sigma meson, we are able to study and compare scenarios with as well
as without a critical point. In the hadron-quark transition regime we identify certain correlations that turn negative unlike
expectation from ideal hadron resonance gas calculations. The striking feature being that these remain negative deep into the
hadronic side and thus could be measured in experiments. Measurement of such quantities in heavy ion collision experiments
can elucidate the location of the QCD transition curve and possibly the critical point.
PACS numbers: 12.38Mh, 25.75.Nq, 12.38.Gc
∗
†
[email protected]
[email protected]
1
Reminiscences of my work with Richard Lewis Arnowitt
Pran Nath∗
Department of Physics, Northeastern University, Boston, MA 02115, USA
arXiv:1502.00639v1 [physics.hist-ph] 2 Feb 2015
This article contains reminiscences of the collaborative work that Richard Arnowitt and I did
together which stretched over many years and encompasses several areas of particle theory. The
article is an extended version of my talk at the Memorial Symposium in honor of Richard Arnowitt
at Texas A&M, College Station, Texas, September 19-20, 2014.
My collaboration with Richard Arnowitt (1928-2014)1
started soon after I arrived at Northeastern University in
1966 and continued for many years. In this brief article
on the reminiscences of my work with Dick I review some
of the more salient parts of our collaborative work which
includes effective Lagrangians, current algebra, scale invariance and its breakdown, the U (1) problem, formulation of the first local supersymmetry and the development of supergravity grand unification in collaboration
of Ali Chamseddine, and its applications to the search
for supersymmetry.
I.
EFFECTIVE LAGRANGIANS AND
CURRENT ALGEBRA
In 1964 Gell-Mann [1] proposed that “quark-type”
equal-time commutation relations for the vector and the
axial vector currents of weak interaction theory serve
as a basis for calculations involving strongly interacting
particles. Combined with the conserved vector current
(CVC), partially conserved vector current (PCAC) and
the soft pion approximation many successful results were
obtained (see, for example, [2]). However, around 1967
an important issue arose which concerned the breakdown
of the soft pion approximation in the analysis of ρ → ππ
and A1 → ρπ where the soft pion approximation gave
very poor results [3, 4]. This problem was overcome
in work with Dick and Marvin Friedman by giving up
the soft pion approximation and using the effective Lagrangian method to compute the mesonic vertices [5–7].
A number of other techniques were being pursued at that
time such as Ward identities by Schnitzer and Weinberg
[8], phenomenological Lagrangians by Schwinger[9], Wess
and Zumino [10], and by Ben Lee and Nieh [11] and other
techniques [12–14].
Here I describe briefly the approach that Dick
Arnowitt, Marvin Friedman and I followed which first
of all involved developing an effective Lagrangian for the
πρA1 system but then using current algebra conditions
to constrain the parameters of the effective Lagrangian.
The effective Lagrangian was a deduction from the following set of conditions: (i) single particle saturation in
computation of T-products of currents, (ii) Lorentz invariance, (iii) “spectator” approximation, and (iv) locality which implies a smoothness assumption on the vertices. The above assumptions lead us to the conclusion
that the simplest way to achieve these constraints is via
an effective Lagrangian which for T-products of three
currents requires writing cubic interactions involving π, ρ
and A1 fields and allowing for no derivatives in the first
-order formalism and up to one-derivative in the second
order formalism. The effective Lagrangian is to be used
to first order in the coupling constants for three point
functions [5–7] and to second order in the coupling constants for 4-point functions and to N − 2 order in the
coupling constant for N point functions. The second step
consisted of the imposition of the constraints of current
algebra, CVC and PCAC to determine the parameters
appearing in the effective Lagrangian. Thus in addition
to [5–7] several other applications of the effective Lagrangian method were made [15–21]. Below we give some
further details of the effective Lagrangian construction.
We begin by considering a T-product of three currents,
i.e.
β
µ
F αµβ ≡< 0|T (Aα
a (x)Vc (z)Ab (y))|0 > ,
For the time ordering x0 > z 0 > y 0 Eq. (1) can be expanded so that
F αµβ =
n,m
n,m
1
Email: [email protected]
Obituary: Richard Lewis Arnowitt, Physics Today 67(12), 68
(2014). Submitted by Stanley Deser (Brandeis and Caltech),
Charles Misner (University of Maryland), Pran Nath (Northeastern University), and Marlan Scully (Texas A&M and Princeton
University).
β
µ
< 0|Aα
a |n >< n|Vc |m >< m|Ab |0 > . (2)
Here the states n and m can only be either π or A1
mesons. It is clear that the matrix elements that are
µ
involved are < 0|Aα
a |πq1 a1 >, < πq1 a1 |Vc |πq2 a2 > and
β
< πq2 a2 |Ab |0 > and additional terms where π is replaced
by A1 . For another time ordering , i.e., y 0 > x0 > z 0 one
has
F αµβ =
∗
(1)
µ
< 0|Aβb |n >< n|Aα
a |m >< m|Vc |0 > , (3)
where the state n must be a π or A1 state. However, the
state m must be a ρ state to have non-vanishing matrix
elements. In addition to the above we must also include
two particle intermediate states so that for the time ordering x0 > y 0 > z 0 we have
February 3, 2015
arXiv:1502.00556v1 [astro-ph.CO] 2 Feb 2015
Exit from Inflation with a First-Order Phase Transition
and a Gravitational Wave Blast
Amjad Ashoorioon1,a
a
Institutionen f¨or fysik och astronomi Uppsala Universitet, Box 803, SE-751 08 Uppsala,
Sweden
Abstract
In double-field inflation, which exploits two scalar fields, one of the fields rolls
slowly during inflation whereas the other field is trapped in a meta-stable vacuum.
The nucleation rate from the false vacuum to the true one becomes substantial enough
that triggers a first order phase transition and ends inflation. We revisit the question
of first order phase transition in an “extended” model of hybrid inflation, realizing the
double-field inflationary scenario, and correctly identify the parameter space that leads
to a first order phase transition at the end of inflation. We compute the gravitational
wave profile which is generated during this first order phase transition. Assuming
instant reheating, the peak frequency falls in the 1 GHz to 10 GHz frequency band
and the amplitude varies in the range 10−8 ΩGW h2 10−11 , depending on the value
of the cosmological constant in the false vacuum. The signature could be observed
by the planned Chongqing high frequency gravitational probe. For a narrow band of
vacuum energies, the first order phase transition can happen after the end of inflation
via the violation of slow-roll, with a peak frequency that varies from 1 THz to 100
THz. For smaller values of cosmological constant, even though inflation can end via
slow-roll violation, the universe gets trapped in a false vacuum whose energy drives a
second phase of eternal inflation. This range of vacuum energies do not lead to viable
inflationary models, unless the value of the cosmological constant is compatible with
the observed value, M ∼ 10−3 eV.
1
e-mail: [email protected]
K → ππ ∆I = 3/2 decay amplitude in the continuum limit.
T.Blum,1, 2 P.A.Boyle,3 N.H.Christ,4 J.Frison,3 N.Garron,5, 6
T.Janowski,7 C.Jung,8 C.Kelly,2 C.Lehner,8 A.Lytle,9
R.D. Mawhinney,4 C.T.Sachrajda,7 A.Soni,8 H.Yin,4 and D.Zhang4
arXiv:1502.00263v1 [hep-lat] 1 Feb 2015
(RBC and UKQCD Collaborations)
1 Physics
2 RIKEN-BNL
3 SUPA,
Department, University of Connecticut, Storrs, CT 06269-3046, USA
Research Center, Brookhaven National Laboratory, Upton, NY 11973, USA
School of Physics, The University of Edinburgh, Edinburgh EH9 3JZ, UK
4 Physics
5 DAMTP,
6 School
7 School
Department, Columbia University, New York, NY 10027, USA
University of Cambridge, Wilberforce Road, Cambridge CB3 0WA, UK
of Computing and Mathematics, Plymouth University Plymouth, PL4 8AA, UK
of Physics and Astronomy, University of Southampton, Southampton SO17 1BJ, UK
8 Brookhaven
9 SUPA,
National Laboratory, Upton, NY 11973, USA
School of Physics and Astronomy, University of Glasgow, Glasgow, G12 8QQ, UK
We present new results for the amplitude A2 for a kaon to decay into two
pions with isospin I = 2: ReA2 = 1.50(4)stat (14)syst × 10−8 GeV; ImA2 =
−6.99(20)stat (84)syst × 10−13 GeV. These results were obtained from two ensembles
generated at physical quark masses (in the isospin limit) with inverse lattice spacings
a−1 = 1.728(4) GeV and 2.358(7) GeV. We are therefore able to perform a continuum
extrapolation and hence largely to remove the dominant systematic uncertainty from
our earlier results [1, 2], that due to lattice artefacts. The only previous lattice computation of K → ππ decays at physical kinematics was performed using an ensemble
at a single, rather coarse, value of the lattice spacing (a−1 ≃ 1.37(1) GeV). We confirm the observation reported in [3] that there is a significant cancellation between
the two dominant contributions to ReA2 which we suggest is an important ingredient in understanding the ∆I = 1/2 rule, ReA0 /ReA2 ≃ 22.5, where the subscript
denotes the total isospin of the two-pion final state. Our result for A2 implies that
the electroweak penguin contribution to ǫ′ /ǫ is Re(ǫ′ /ǫ)EWP = −(6.6 ± 1.0) × 10−4 .
PACS numbers: 11.15.Ha, 11.30.Rd, 12.15.Ff, 12.38.Gc
MPP-2015-13
arXiv:1502.01011v1 [hep-ph] 3 Feb 2015
keV Sterile Neutrino Dark Matter from Singlet
Scalar Decays: Basic Concepts and Subtle Features
Alexander Merle∗ and Maximilian Totzauer†
Max-Planck-Institut f¨
ur Physik (Werner-Heisenberg-Institut),
F¨ohringer Ring 6, 80805 M¨
unchen, Germany
February 4, 2015
Abstract
We perform a detailed and illustrative study of the production of keV sterile neutrino Dark Matter (DM) by decays of singlet scalars in the early Universe. In the
current study we focus on providing a clear and general overview of this production mechanism. For the first time we study all regimes possible on the level of
momentum distribution functions, which we obtain by solving a system of Boltzmann equations. These quantities contain the full information about the production
process, which allows us to not only track the evolution of the DM generation but
to also take into account all bounds related to the spectrum, such as constraints
from structure formation or from avoiding too much dark radiation. In particular we show that this simple production mechanism can, depending on the regime,
lead to strongly non-thermal DM spectra which may even feature more than one
peak in the momentum distribution. These cases could have particularly interesting
consequences for cosmological structure formation, as their analysis requires more
refined tools than the simplistic estimate using the free-streaming horizon. Here
we present the mechanism including all concepts and subtleties involved, for now
using the assumption that the effective number of relativistic degrees of freedom is
constant during DM production, which is applicable in a significant fraction of the
parameter space. This allows us to derive analytical results to back up our detailed
numerical computations, thus leading to the most comprehensive picture of keV
sterile neutrino DM production by singlet scalar decays that exists up to now.
∗
†
email: [email protected]
email: [email protected]
DESY 14–186
MITP/14–073
LPSC-14–260
January 2015
ISSN 0418–9833
arXiv:1502.01001v1 [hep-ph] 3 Feb 2015
Inclusive B-meson production at small pT in
the general-mass variable-flavor-number
scheme
B. A. Kniehl1 , G. Kramer1 , I. Schienbein2 and H. Spiesberger3
1
II. Institut f¨
ur Theoretische Physik, Universit¨at Hamburg,
Luruper Chaussee 149, 22761 Hamburg, Germany
2
LPSC, Universit´e Grenoble-Alpes, CNRS/IN2P3,
53 avenue des Martyrs, 38026 Grenoble, France
3
PRISMA Cluster of Excellence, Institut f¨
ur Physik,
Johannes Gutenberg-Universit¨at, 55099 Mainz, Germany,
and Centre for Theoretical and Mathematical Physics and Department of Physics,
University of Cape Town, Rondebosch 7700, South Africa
Abstract
We calculate the cross section for the inclusive production of B mesons in pp and p¯
p
collisions at next-to-leading order in the general-mass variable-flavor-number scheme
and show that a suitable choice of factorization scales leads to a smooth transition
to the fixed-flavor-number scheme. Our numerical results are in good agreement
with data from the Tevatron and LHC experiments at small and at large transverse
momenta.
PACS: 12.38.Bx, 12.39.St, 13.85.Ni, 14.40.Nd
1
Prepared for submission to JHEP
arXiv:1502.00925v1 [hep-ph] 3 Feb 2015
Matching the Nagy-Soper parton shower at
next-to-leading order
M. Czakon, H. B. Hartanto, M. Kraus, M. Worek
Institut für Theoretische Teilchenphysik und Kosmologie, RWTH Aachen University, D-52056
E-mail: [email protected], [email protected],
[email protected], [email protected]
Abstract: We present an Mc@Nlo-like matching of next-to-leading order QCD calculations with the Nagy-Soper parton shower. An implementation of the algorithm within the
Helac-Dipoles Monte Carlo generator is used to address the uncertainties and ambiguities of the matching scheme. First results obtained using the Nagy-Soper parton shower
implementation in Deductor in conjunction with the Helac-Nlo framework are given
√
for the pp → tt¯j + X process at the LHC with s = 8 TeV. Effects of resummation are
discussed for various observables.
Keywords: Hadronic Colliders, Monte Carlo Simulations, QCD Phenomenology
TTK-15-07
Low-energy phenomenology of trinification: an effective left-right-symmetric model
Jamil Hetzel∗ and Berthold Stech†
arXiv:1502.00919v1 [hep-ph] 3 Feb 2015
Institut f¨
ur Theoretische Physik, Universit¨
at Heidelberg, D-69120 Heidelberg, Germany
(Dated: February 4, 2015)
The trinification model is an interesting extension of the Standard Model (SM) based on the
gauge group SU (3)C × SU (3)L × SU (3)R . We study its low-energy phenomenology by constructing
a low-energy effective field theory, thereby reducing the number of particles and free parameters that
need to be studied. The resulting model predicts that several new scalar particles have masses in
the O (100 GeV) range. We study a few of the interesting phenomenological scenarios, such as the
presence of a light fermiophobic scalar in addition to a SM-like Higgs, or a degenerate (twin) Higgs
state at 126 GeV. We point out regions of the parameter space that lead to measurable deviations
from SM predictions of the Higgs couplings. Hence the trinification model awaits crucial tests at
the Large Hadron Collider in the coming years.
PACS numbers: 12.10.Dm
I.
INTRODUCTION
The discovery of the Higgs boson [1, 2] marks the establishment of the Standard Model (SM) of particle physics
as the model that correctly describes physics at experimentally available energies to date. All SM particles have
been discovered, and the experimental data gathered at
particle colliders match the predictions of the SM to good
precision [3]. Yet, the SM is regarded to be an incomplete
theory of nature: it lacks a dark matter candidate and is
incompatible with the observation of non-zero neutrino
masses. Also, the fermion masses and mixings are free
parameters that display hierarchical patterns, and parity
violation is introduced by hand. Therefore our quest towards a better theory of nature requires us to extend the
SM.
Grand Unified Theories (GUTs) [4–8] are interesting
extensions of the SM in which the SM gauge group
SU (3)C ×SU (2)L ×U (1)Y is embedded in a larger simple
gauge group. The exceptional group E6 is an attractive
example of a GUT group [6–8]. It is anomaly-free and
left-right-symmetric (LR-symmetric), and as such it provides an explanation for parity violation in the SM by
spontaneous symmetry breaking. It appears in the compactification of string theories, which leads to either fourdimensional E6 gauge symmetry or one of E6 ’s maximal
subgroups [9, 10]. One of these maximal subgroups is the
‘trinification group’ G333 ≡ SU (3)C × SU (3)L × SU (3)R .
Models based on the trinification group have been studied in several contexts [11–18].
In this work, ‘trinification model’ will refer to the setup
described in refs. [7, 19–23]. The setup described there is
interesting for several reasons: fermion masses and mixings of the SM can be reproduced using only a few parameters, with a satisfactory fit for the solar neutrino
mass difference and the neutrino mixing pattern. Also,
a Standard-Model-like Higgs with a mass close to 126
∗
†
[email protected]
[email protected]
GeV appears in a large region of parameter space of the
model. Furthermore, it gives predictions for the matrix
element of neutrinoless double-beta decay and the neutrino masses, which allow the model to be tested with
low-energy experiments. It also allows for various interesting phenomenological scenarios, such as the presence
of a light fermiophobic Higgs in addition to the StandardModel-like Higgs, or even a degenerate Higgs state at 126
GeV.
In order to compare the trinification model with experiment, a study of the low-energy phenomenology is
necessary. Due to the large number of scalars, a study
of the full scalar mass matrix is challenging. However,
several of the scalar fields will obtain very large masses
when the trinification symmetry is broken, and thus can
be integrated out from the theory. The result is an effective field theory with the LR-symmetric gauge group
SU (3)C × SU (2)L × SU (2)R × U (1)B−L , and fewer scalar
fields than in the trinification model. This model has
the same low-energy properties as the trinification model,
but is easier to study. We will refer to this model as the
low-energy trinification (LET) model.
Neither the LET model nor the trinification model resolves the hierarchy problem. This problem is hidden in
the vacuum expectation values (vevs) used in the model,
which are presently not understood. In our treatment,
all dimensionful parameters and masses are fully determined by these vevs multiplied by dimensionless coupling
constants. Since these vevs are momentum- and scaleindependent (except for wave-function renormalization),
their use as fixed parameters is justified.
Left-right symmetric models based on the gauge group
SU (3)C × SU (2)L × SU (2)R × U (1)B−L [24–26] have
been studied extensively in the literature. Moreover,
these models have many features in common with the
two-Higgs-doublet model (2HDM) [27]. However, the
LET model has properties that distinguish it from more
general LR-symmetric models and the 2HDM, due to
the trinification origin at high energy scales. A LRsymmetric model in the context of the trinification model
has not been studied before to the best of our knowledge.
arXiv:1502.00899v1 [hep-ph] 3 Feb 2015
International Journal of Modern Physics: Conference Series
c The Authors
Overview of TMD evolution
Dani¨
el Boer
Van Swinderen Institute, University of Groningen
Nijenborgh 4, NL-9747 AG Groningen, The Netherlands
[email protected]
Received Day Month Year
Revised Day Month Year
Published Day Month Year
Transverse momentum dependent parton distributions (TMDs) appear in many scattering processes at high energy, from the semi-inclusive DIS experiments at a few GeV to the
Higgs transverse momentum distribution at the LHC. Predictions for TMD observables
crucially depend on TMD factorization, which in turn determines the TMD evolution
of the observables with energy. In this contribution to SPIN2014 TMD factorization is
outlined, including a discussion of the treatment of the nonperturbative region, followed
by a summary of results on TMD evolution, mostly applied to azimuthal asymmetries.
Keywords: QCD; Spin and Polarization Effects; Transverse Momentum Dependence
PACS numbers: 12.38.-t; 13.88.+e
1. TMD factorization
Many angular asymmetries in the transverse momentum distribution of produced
hadrons in semi-inclusive deep inelastic scattering (SIDIS), e p → e′ h X, have been
measured by the HERMES, COMPASS, and JLab experiments. Evolution is needed
to compare those results obtained at different energies. The evolution is dictated by
the appropriate factorization. As SIDIS is sensitive to the transverse momentum of
quarks through a measurement of Ph⊥ , the observed transverse momentum of the
produced hadron, it is naturally described within the framework of TMD factorization. Several forms of TMD factorization have been put forward in the literature
for a number of processes [1-6], which besides SIDIS includes the Drell-Yan (DY)
process (lepton pair production in hadron-hadron collisions), back-to-back hadron
production in electron-positron annihilation (e+ e− → h1 h2 X), and Higgs production. The main differences among the various approaches concern the treatment of
spurious rapidity or lightcone divergences, in order to make each factor well-defined,
and the redistribution of contributions to avoid the appearance of large logarithmic
This is an Open Access article published by World Scientific Publishing Company. It is distributed
under the terms of the Creative Commons Attribution 3.0 (CC-BY) License. Further distribution
of this work is permitted, provided the original work is properly cited.
1
Thermodynamics and fluctuations of conserved charges in Hadron Resonance Gas
model in finite volume
Abhijit Bhattacharyya∗
Department of Physics, University of Calcutta, 92, A. P. C. Road, Kolkata - 700009, INDIA
arXiv:1502.00889v1 [hep-ph] 3 Feb 2015
Rajarshi Ray† and Subhasis Samanta‡
Center for Astroparticle Physics & Space Science,
Bose Institute, Block-EN, Sector-V,
Salt Lake, Kolkata-700091, INDIA
&
Department of Physics, Bose Institute,
93/1, A. P. C Road, Kolkata - 700009, INDIA
Subrata Sur§
Department of Physics, Panihati Mahavidyalaya,
Barasat Road, Sodepur, Kolkata - 700110, INDIA
The thermodynamics of hot and dense matter created in heavy-ion collision experiments are
usually studied as a system of infinite volume. Here we report on possible effects for considering
a finite system size for such matter in the framework of the Hadron Resonance Gas model. The
bulk thermodynamic variables as well as the fluctuations of conserved charges are considered. We
find that the finite size effects are insignificant once the observables are scaled with the respective
volumes. The only substantial effect is found in the fluctuations of electric charge which may
therefore be used to extract information about the volume of fireball created in heavy-ion collision
experiments.
PACS numbers: 12.38.Mh, 21.60.-n,21.65.Mn,25.75.-q
Our present day universe contains a significant fraction of matter in hadronic form. In the very early universe − a
few microseconds after the Big Bang [1] when the temperature was extremely high, the strongly interacting matter is
expected to have existed in the partonic form. Similar exotic state of matter may exist inside compact stars due to
extremely high matter density attained by gravitational compression [2]. For the last few decades various experimental
efforts are being made to recreate such exotic matter through the collisions of heavy-ions at ultra-relativistic energies.
Experimental facilities at CERN (France/Switzerland), BNL (USA) and the upcoming facility at GSI (Germany) are
at the forefront of efforts taken to create these exotic states of matter. One of the major goals in the experiments
is to study thermodynamic properties of strongly interacting matter at high temperatures and densities. Present
experimental data as well as lattice QCD simulations seem to indicate a smooth cross over from hadronic to quark
gluon matter at low density and high temperature [3, 4]. At high density and low temperature a first order transition
is expected [5–10].
Usually any thermodynamic study assumes the system volume to be infinite. However the fireball created in the
relativistic heavy ion collision experiments has a finite spatial volume. The size of such√spatial volume critically
depends on three parameters : the size of the colliding nuclei, the center of mass energy ( s) and the centrality of
collisions. Analysis of experimental data could reveal the freeze out volume of the system. One way to carry out such
an analysis is the study of HBT radii which√has been done in Ref. [11]. The major finding of this study indicates
that the freeze out volume increases as the s increases and the estimated freeze out volume varies from 2000 f m3
to 3000 f m3 . Another way of estimating the system size is through the comparison of simulation results with the
experimental data as done in Ref. [12] where the UrQMD model [13] is used for this purpose. A study of the P b − P b
collisions at different energies and centralities resulted in a freeze-out volume in the range 50 f m3 to 250 f m3 . We
note that these volumes as quoted above, are the freeze out volumes and as one looks back to the early times in the
evolution of the system, a smaller volume is expected. So both in the quark phase and in the hadron phase (after
the putative phase transition) finite volume could play an important role in the quantitative measurement of the
∗ Electronic
address:
address:
‡ Electronic address:
§ Electronic address:
† Electronic
[email protected]
[email protected]
[email protected]
[email protected]
The hidden symmetries in the PMNS matrix and the light sterile neutrino(s)
Hong-Wei Ke1∗ , Jia-Hui Zhou1 , Shuai Chen1 , Tan Liu1 and Xue-Qian Li2†
1
arXiv:1502.00875v1 [hep-ph] 3 Feb 2015
2
School of Science, Tianjin University, Tianjin 300072, China
School of Physics, Nankai University, Tianjin 300071, China
The approximately symmetric form of the PMNS matrix suggests that there could exist a hidden symmetry
which makes the PMNS matrix different from the CKM matrix for quarks. In literature, all the proposed fully
symmetric textures exhibit an explicit µ − τ symmetry in addition to other symmetries which may be different
for various textures. Observing obvious deviations of the practical PMNS matrix elements from those in the
symmetric textures, there must be a mechanism to distort the symmetry. It might be due to existence of light
sterile neutrinos. In this work, we study the case of the Tribimaximal texture and propose that its apparent
symmetry disappears due to existence of a sterile neutrino. We observe that introducing just one sterile neutrino
is still not sufficient to recover the data, thus a slight µ − τ symmetry breaking is also needed. By considering
those factors, we obtain the PMNS matrix elements which are consistent with data within the experimental
tolerance.
PACS numbers: 14.60.Pq, 14.60.Lm, 14.60.St
Numerous experiments which are carried out in past several decades make the behaviors of neutrinos understandable.
As is commonly accepted, mixing among different flavors of
leptons are due to the mismatch between the mass eigenstates
and flavor eigenstates, it is the same as the quark case, but
different in structures. To bring the weak interaction eigenstates (flavor) to the physical ones (mass), the PontecorvoMaki-Nakawaga-Sakata (PMNS) matrix[1, 2] should be introduced. If there are only three active neutrinos the mixing
matrix is written as


 U11 U12 U13 

V =  U21 U22 U23  .
(1)
U31 U32 U33
Generally there are four independent parameters, namely
three mixing angles and one CP-phase. There are various
schemes to parameterize the matrix in literature. For example, the Chau-Keung(CK) parametrization [3] is


c12 c13
s12 c13
s13 


V =  −c12 s23 s13 eiδ − s12 c23 −s12 s23 s13 eiδ + c12 c23 s23 c13  ,
−c12 s23 s13 eiδ + s12 s23 −s12 s23 s13 eiδ − c12 s23 c23 c13
(2)
where s jk and c jk denote sin θ jk and cos θ jk with j, k = 1, 2, 3.
The measured values of the PMNS matrix exhibit an approximately symmetric form which may hint that the practical matrix originates from a high symmetry, but is distorted
by some mechanisms. Indeed, one of the physics achievements of the 20th century convinces us that symmetry and
symmetry breaking compose the main picture of the nature,
so one may reasonably expect that an underlying symmetry
determines the mixing matrix of leptons which later is distorted somehow. Lam has shown this possibility in terms of
the group theory[4] where the CKM and PMNS matrices are
separately resulted via different routes to break the large symmetry. In Lam’s scheme, the resultant PMNS still possesses
∗
[email protected][email protected]
an obvious symmetry with θ13 strictly being zero, therefore
to reach the practical PMNS a further symmetry breaking is
needed. It is natural to ask if one can provide a reasonable
mechanism to explain the distortion.
Meanwhile some phenomenological symmetries are observed, such as the quark-lepton complementarity and selfcomplementarity[5–11], µ − τ symmetry[12]. But all those
symmetries are only approximate, so it also implies that there
should be some mechanisms to result in their deviations from
exact symmetric forms.
The high-precision measurements [13–16] determines
θ12 ≈ 34◦ and θ23 ≈ 45◦ and a small, but non-zero θ13 .
The values could be traced to a mixing pattern with high
symmetries i.e. for example the tribimaximal (TB) mixing
pattern[12] which is one of the possible symmetric textures


2


√1
0


3
3
 .
VT B =  − √1
(3)
√1
√1 
3

6
2 

1
1
1
√
√
√
−
6
3
2
which means that θ12 = 35.26◦, θ23 = 45◦ and θ13 = 0◦ in the
adopted parametrization. In this scenario the µ − τ symmetry
holds and in the mass eigenstate ν2 , νe , νµ and ντ have the
same probability.
However the measurements of the accelerator and reactor neutrino oscillation experiments[17–21] determine θ12 =
◦
+1.40 ◦
+0.46 ◦
(33.65+1.11
−1.00) , θ23 = (38.41−1.21) and θ13 = (8.93−0.48 ) . It is
noted that the values are set based on the scenario for three
generations of neutrinos. Apparently the TB mixing patterns
decline from the data. One may ask whether the symmetries
in the TB mixing patterns should be abandoned? Even though
it is too early to make a definite conclusion yet, there exists a
possibility that those symmetries still hold, whereas the matrix
might be distorted from the symmetric form by new physics.
Recently, the anomalies of short-baseline neutrino
experiments[22–24] hint that there may exist light sterile
neutrinos which mix with the active ones. If this picture
indeed works, the existence of sterile neutrinos would play a
role to make the mixing matrix being in the superficial form
where the original symmetries are just hidden somehow or
arXiv:1502.00789v1 [hep-ph] 3 Feb 2015
CYCU-HEP-15-01
EPHOU-15-002
Gauge extension of non-Abelian discrete flavor
symmetry
Florian Beye1∗ , Tatsuo Kobayashi2† and Shogo Kuwakino3‡
1
Department of Physics, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8602, Japan
2
Department of Physics, Hokkaido University, Sapporo 060-0810, Japan
3
Department of Physics, Chung-Yuan Christian University, 200, Chung-Pei Rd.
Chung-Li,320, Taiwan
Abstract
We investigate a gauge theory realization of non-Abelian discrete flavor symmetries
and apply the gauge enhancement mechanism in heterotic orbifold models to fieldtheoretical model building. Several phenomenologically interesting non-Abelian discrete
symmetries are realized effectively from a U (1) gauge theory with a permutation symmetry. We also construct a concrete model for the lepton sector based on a U (1)2 ⋊ S3
symmetry.
∗
Electronic address: [email protected]
Electronic address: [email protected]
‡
Electronic address: [email protected]
†
1
IPMU15-0013, ICRR-Report 697-2014-23
IceCube potential for detecting the Q-ball dark matter in gauge mediation
arXiv:1502.00715v1 [hep-ph] 3 Feb 2015
a
Shinta Kasuyaa, Masahiro Kawasakib,c and Tsutomu T. Yanagidac
Department of Mathematics and Physics, Kanagawa University, Kanagawa 259-1293, Japan
b
Institute for Cosmic Ray Research, the University of Tokyo, Chiba 277-8582, Japan
c
Kavli Institute for the Physics and Mathematics of the Universe (WPI),
Todai Institutes for Advanced Study, the University of Tokyo, Chiba 277-8582, Japan
(Dated: February 3, 2015)
We study the Q-ball dark matter in the gauge-mediated supersymmetry breaking, and seek for
the detection possibility in the IceCube experiment. We find that the Q balls would be the dark
matter in the parameter region different from that for the gravitino dark matter. In particular, the
Q ball is a good dark matter candidate for low reheating temperature, which may be suitable for the
Affleck-Dine baryogenesis and/or nonthermal leptogenesis. The dark matter Q balls are detectable
by the IceCube-like experiments in the future, which is the peculiar feature compared to the case
of the gravitino dark matter.
I.
INTRODUCTION
Gauge mediation [1–7] is very attractive for the supersymmetry (SUSY) breaking mechanism, since it is free from the
serious flavor-changing neutral current (FCNC) problem in the supersymmetric standard model. It has been recently
pointed out that the minimal gauge mediation model is still consistent with all experimental data and cosmological
constraints [8]. Thus it is worthwhile to discuss possible dark matter candidates in the gauge mediation model.
A well-known candidate is the gravitino. If the gravitino mass m3/2 is lighter than 1 keV, the gravitino cannot
<
saturate the dark matter abundance [9]. Even for 1 keV <
∼ m3/2 ∼ 100 keV, the gravitino dark matter would be too
warm, and the small-scale fluctuations will be erased [10, 11]. In addition, the reheating temperature after inflation
3
should be TRH >
∼ 10 GeV for the thermally produced gravitinos to be dark matter [12].
To test the gravitino dark matter we need to produce first SUSY particles in collider experiments and see their
decays into the almost massless gravitino. However, even if the SUSY particles are produced it is very challenging to
observe the decay if the gravitino is heavier than 1 keV.
In this paper we discuss the other possibility that the Q ball is the dark matter in the gauge mediation, and seek
for the discovery potential for the Q-ball dark matter at the IceCube. There are a lot of flat directions in SUSY. They
consist of some combinations of squarks (and sleptons) in the minimal supersymmetric standard model (MSSM), so
that they carry baryonic charge. The Q ball is a non topological soliton, the energy minimum configuration of the flat
direction for a finite baryon number [13]. Since the Q ball with large enough charge (the baryon number) is stable
against the decay into nucleons in the gauge mediation, it is reasonable to consider the Q ball as the dark matter
candidate [14].
We show that the Q ball would be the dark matter in the parameter region different from that for the gravitino
<
dark matter: the gravitino mass could range for keV <
∼ m3/2 ∼ GeV, or even smaller, and the reheating temperature
4
must be TRH <
∼ 10 GeV. Moreover, it has completely different detection procedure from that for other dark matter
candidate. Very large volume neutrino detectors such as the IceCube can detect directly the dark matter Q balls in
the (near) future.
The sketch of the paper is as follows. In the next section, we show the set-up of the gauge-mediated SUSY breaking.
In sec. 3, we review the Q ball in the gauge mediation, where there are two types of Q ball. We estimate the abundance
of the Q ball and show the parameter region where the Q balls could be the dark matter in Sec. 4. In the same section,
we seek for the possibility of the Q-ball detection in the IceCube. Sec. 5 is devoted to our conclusion.
II.
GAUGE-MEDIATED SUSY BREAKING
We adopt the so-called minimal direct gauge mediation described only by a few parameters, and free from both the
SUSY FCNC problem and the SUSY CP problem [8]. The SUSY is spontaneously broken by the vacuum expectation
value of the SUSY breaking field Z as
Z = 0,
FZ = F.
(1)
¯ a pair of some
The SUSY breaking sector is connected to the observable sector by the messenger fields Ψ and Ψ,
representations and anti-representations of the minimal GUT group SU (5), through the Yukawa interactions as
¯
¯ + Mmess ΨΨ,
W = kZ ΨΨ
(2)
Revisit of the neutron/proton ratio puzzle in intermediate-energy heavy-ion collisions
Hai-Yun Kong,1, 2 Yin Xia,1, 2 Jun Xu∗ ,1 Lie-Wen Chen,3, 4 Bao-An Li,5 and Yu-Gang Ma1, 6
1
arXiv:1502.00778v1 [nucl-th] 3 Feb 2015
Shanghai Institute of Applied Physics, Chinese Academy of Sciences, Shanghai 201800, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
3
Department of Physics and Astronomy and Shanghai Key Laboratory for Particle Physics and Cosmology,
Shanghai Jiao Tong University, Shanghai 200240, China
4
Center of Theoretical Nuclear Physics, National Laboratory of Heavy Ion Accelerator, Lanzhou 730000, China
5
Department of Physics and Astronomy, Texas A&M University-Commerce, Commerce, TX 75429-3011, USA
6
Shanghai Tech University, Shanghai 200031, China
(Dated: February 4, 2015)
Incorporating a newly improved isospin- and momentum-dependent interaction in the isospindependent Boltzmann-Uehling-Uhlenbeck transport model IBUU11, we have investigated relative
effects of the density dependence of nuclear symmetry energy Esym (ρ) and the neutron-proton effective mass splitting m∗n − m∗p on the neutron/proton ratio of free nucleons and those in light clusters.
It is found that the m∗n − m∗p has a relatively stronger effect than the Esym (ρ) and the assumption of m∗n ≤ m∗p leads to a higher neutron/proton ratio. Moreover, this finding is independent of
the in-medium nucleon-nucleon cross sections used. However, results of our calculations using the
Esym (ρ) and m∗n − m∗p both within their current uncertainty ranges are all too low compared to the
recent NSCL/MSU double neutron/proton ratio data from central 124 Sn+124 Sn and 112 Sn+112 Sn
collisions at 50 and 120 MeV/u, thus calling for new mechanisms to explain the puzzlingly high
neutron/proton ratio observed in the experiments.
PACS numbers: 25.70.-z, 24.10.Lx, 21.30.Fe
To pin down the density dependence of nuclear symmetry energy Esym (ρ) has long been a major challenge
for both nuclear physics and astrophysics. While much
progress has been made in the past decade, many interesting issues remain to be resolved [1–4]. A larger
symmetry energy generally corresponds to a more repulsive (attractive) underlying symmetry potential Usym
for neutrons (protons) in neutron-rich nuclear matter as
they are linearly proportional to each other according
to the Hugenholtz-Van Hove theorem [5] or the Bruckner theory [6, 7], see, e.g., Refs. [8–10] for the explicit
relationship between Esym (ρ) and Usym . On the other
hand, the in-medium nucleon effective mass describes to
the first order effects due to the non-locality of the underlying nuclear interactions and the Pauli exchange effects in many-fermion systems [11]. It can be calculated
from the momentum dependence of the single-particle
potential in non-relativistic models or the Schr¨odingerequivalent potential in relativistic models. The nucleon
effective mass is related to many interesting problems in
both nuclear physics and astrophysics [12–15]. It has further been found that the neutrons and protons may have
different effective masses in neutron-rich matter due to
the momentum dependence of the symmetry (isovector)
potential. However, calculations within different models
using various interactions, e.g., the Brueckner-HartreeFock approach [16], the relativistic mean-field model [17],
and the Skyrme-Hartree-Fock calculation [18, 19], predict rather different values for the neutron-proton effec-
∗ corresponding
author: [email protected]
tive mass splitting m∗n−p ≡ m∗n − m∗p . Thus, currently
there is no consensus as to whether the m∗n−p is negative, zero, or positive. However, the value of m∗n−p affects significantly isospin-sensitive observables in heavyion collisions [20–30] as well as thermal and transport
properties of neutron-rich matter [31, 32]. It also has important ramifications in astrophysics [33]. For instance,
the equilibrium neutron/proton ratio in the primordial
∗
nucleosynthesis is determined by (n/p)eq = e−mn−p/T in
the early (≥ 1ms) universe when the temperature T was
high (≥ 3 MeV) [34].
Recently, analyses of the free neutron/proton double
ratio from central 124 Sn+124 Sn and 112 Sn+112 Sn collisions at 50 and 120 MeV/u at the NSCL/MSU seem
to indicate that protons have a slightly larger effective
mass than neutrons based on comparisons with calculations within an improved quantum molecular dynamics model using Skyrme interactions [35]. However, the
preferred Skyrme interaction leading to a negative m∗n−p
also predicts a nucleon symmetry potential that is increasing with nucleon energy/momentum at saturation
density. This kind of energy dependence of the symmetry potential is opposite to that extracted from optical
model analyses of huge sets of nucleon-nucleus scattering
data accumulated since the earlier 60’s [10, 36–42]. This
situation clearly calls for more theoretical studies with
different transport models and examinations of various
model ingredients. In fact, it was known that the isospindependent Boltzmann-Uehing-Uhlenbeck (IBUU) transport model using a momentum-dependent symmetry potential corresponding to m∗n > m∗p [43] under-predicts
the old NSCL double neutron/proton data [44] no matter
how the density-dependence of symmetry energy and the
arXiv:1502.00848v1 [nucl-ex] 3 Feb 2015
Femtoscopic pΛ Correlations in Pb-Pb Collisions
√
at sNN = 2.76 TeV with ALICE
Hans Beck∗
Institut f¨
ur Kernphysik, Goethe-Universit¨at, Frankfurt, Germany
for the ALICE Collaboration
February 4, 2015
Abstract
Two-particle correlations at small relative momenta give insight into
the size of the emitting source. Of particular interest is the test of hydrodynamic models which predict a universal, apparent decrease of the extent
of the system with increasing transverse mass mT as a consequence of the
strong radial flow in heavy-ion collisions at LHC energies. This contribution presents a study of correlations for protons and Λ particles in Pb-Pb
√
collisions at sNN = 2.76 TeV measured with ALICE. The investigated
particle species expand the experimental reach in mT due to their high
rest mass. Residual impurities in the samples from misidentification and
contributions from feed-down are corrected using data-driven techniques.
Correlation functions are obtained in several centrality classes and mT
intervals at large mT and show the expected decrease in volume of the
strongly interacting fireball for more peripheral collisions. We observe
the decrease of the source size with mT , predicted by hydrodynamical
calculations, out to mT = 2.18 GeV/c2 .
∗ email:
[email protected]
1
arXiv:1502.00703v1 [nucl-ex] 3 Feb 2015
Study of in-medium mass modification at J-PARC
K. Aoki1 for the J-PARC E16 Collaboration
1
KEK, High Energy Accelerator Research Organization,
Tsukuba, Ibaraki 305-0801, Japan
February 4, 2015
Abstract
Study of in-medium mass modification has attracted interest in terms of
the restoration of the spontaneously broken chiral symmetry, which is responsible for the generation of hadron mass. Many experiments were performed
to measure in-medium property of hadrons but there is no consensus yet. JPARC E16 has been proposed to study in-medium property of vector mesons
via dilepton decay channel. The status of spectrometer R & D is explained.
Other related experiments planned at J-PARC are also introduced.
1
Introduction
Spontaneous breaking of the chiral symmetry is considered to be the origin of hadron
mass. The chiral symmetry is expected to be (partially) restored in finite density
and the hadron mass is predicted to decrease, even at the normal nuclear density.
Our purpose is to investigate the origin of hadoron mass through mass modification
of hadrons.
The dilepton decay channel of vector mesons produced in nuclear reactions is
a good probe of in-medium mass modification since it is free from the final state
interactions. We take p + A → φ + X reaction as an example to explain the
1