arXiv:1501.01596v1 [cond-mat.mtrl-sci] 3 Jan 2015

Hybridizing matter-wave and classical accelerometers
Hybridizing matter-wave and classical accelerometers
J. Lautier,1 L. Volodimer,1 T. Hardin,1 S.Merlet,1 M. Lours,1 F. Pereira Dos Santos,1 and A. Landragin1, a)
LNE-SYRTE, Observatoire de Paris, CNRS, UPMC, 61 avenue de l’Observatoire, 75014 Paris,
France
arXiv:1410.0050v2 [physics.atom-ph] 6 Oct 2014
(Dated: 8 October 2014)
We demonstrate a hybrid accelerometer that benefits from the advantages of both conventional and atomic
sensors in terms of bandwidth (DC to 430 Hz) and long term stability. First, the use of a real time correction
of the atom interferometer phase by the signal from the classical accelerometer enables to run it at best
performances without any isolation platform. Second, a servo-lock of the DC component of the conventional
sensor output signal by the atomic one realizes a hybrid sensor. This method paves the way for applications
in geophysics and in inertial navigation as it overcomes the main limitation of atomic accelerometers, namely
the dead times between consecutive measurements.
PACS numbers: Valid PACS appear here
Keywords: Atom interferometry, Inertial navigation, Gravimetry
Atom interferometers have demonstrated to be both
very sensitive and accurate inertial sensors, which opens
wide area of applications in geophysics and inertial guidance. These possibilities are already exploited in absolute
cold atom gravimetry1, where record sensitivities have
been demonstrated2,3 . In the mean time, developments
have been conducted in order to simplify and reduce the
size of such devices to be compliant with these applications4–6 . One of the main remaining difficulties comes
from the intrinsic low frequency sampling, in the order
of 1 Hz, of the vibration noise in cold atom devices. This
is associated to their sequential operation leading to dead
times between consecutive measurements and aliasing effect of the high frequency vibration noise7 . Different
methods have been developed to overcome these difficulties. One consists in using an active or a passive isolation
platform1,2,8,9 , in which the cut-off frequency is below the
cycling frequency. However, these set-ups are bulky, and
thus ill-suited for operation in noisy or mobile environment because of their low frequency resonances. A second method consists in increasing the cycling frequency
(330 Hz has been demonstrated10 ) but to the price of a
drastic reduction in the sensitivity, which scales quadratically with the interrogation time. A last method is based
on the post-correlations between simultaneous measurements from classical and atom accelerometers8. It enables reaching the state of the art performance in absolute gravimetry11, and operating an atom interferometer
in a noisy environment such as a plane12 . However, in
the latter, this method leads to a reduction of the effective bandwidth, and does not solve the dead time issue.
Hybridizing classical and matter-wave inertial sensors enables to overcome these limitations.
In this paper, we demonstrate how such a combination
between a force balanced accelerometer and an atomic
gravimeter enables operating both instruments at maximum performances. The combined signal associates the
a) Electronic
mail: [email protected]
large bandwidth of the mechanical sensor and the long
term stability and the accuracy of the atomic sensor. We
exploit the correlation between the acceleration measurements in order to real-time compensate the atomic interferometer phase fluctuations due to vibrations, and to
operate this instrument at its maximum sensitivity. Furthermore, we take the full advantage of these correlations
to correct for the drift of the mechanical accelerometer.
In a reciprocal fashion, the specific advantages of one sensor are used to overtake the characteristic limitations of
the other. This results in a hybrid sensor, which combines
large bandwidth (DC to 430 Hz), and the bias stability
and the accuracy of the atomic interferometer.
We work with a Mach-Zehnder-like π/2−π−π/2 gravimeter relying on stimulated Raman transitions13 . The sensor head of our instrument is described in8 . The laser
system and a typical measurement sequence are presented in14 . A vertical laser beam and its retro-reflection
on a mirror form the two counter-propagating Raman
beams. These are used to coherently split, deflect, and
recombine a cloud of cold 87 Rb atoms in free-fall in the
gravity field. With this geometry, the atomic phaseshift at the output of the interferometer is given by15 :
△Φ=φ1 −2φ2 +φ3 =−~keff~g T 2 where φi is the phase difference between the two Raman lasers, at the position
of the center of mass of the wavepacket, at the time of
the i-th Raman pulse. ~keff is the effective wave-vector, ~g
the acceleration of the Earth gravity, and T the freeevolution time between two consecutive pulses. Such
atomic accelerometers are thus sensitive to the relative
acceleration between the free-falling atoms and the retroreflecting mirror, that sets the phase reference for the Raman lasers. Residual vibrations of the latter thus induce
a phase noise φvib . Besides, the free-fall of the atoms
induces a Doppler shift which requires to sweep phasecontinuously the frequency difference between the two
Raman lasers to drive the two-photon transitions over
the interrogation time, according to △ω(t)=△ω(0)+αt.
This adds an additional contribution αT 2 to the atomic
phase shift △Φ.
A compact micro-wave synthesizer for transportable cold-atom
interferometers
J. Lautier, M. Lours, and A. Landragin1, a)
arXiv:1406.2911v1 [physics.atom-ph] 11 Jun 2014
LNE-SYRTE, Observatoire de Paris, CNRS, UPMC, 61 avenue de l’Observatoire, 75014 Paris,
France
We present the realization of a compact micro-wave frequency synthesizer for an atom interferometer based
on stimulated Raman transitions, applied to transportable inertial sensing. Our set-up is intended to address
the hyperfine transitions of 87 Rb at 6.8 GHz. The prototype is evaluated both in the time and the frequency
domain by comparison with state-of-the-art frequency references developed at LNE-SYRTE. In free-running
mode, it features a residual phase noise level of -65 dBrad2 .Hz−1 at 10-Hz offset frequency and a white phase
noise level in the order of -120 dBrad2 .Hz−1 for Fourier frequencies above 10 kHz. The phase noise effect on
the sensitivity of the atomic interferometer is evaluated for diverse values of cycling time, interrogation time
and Raman pulse duration. To our knowledge, the resulting contribution is well below the sensitivity of any
demonstrated cold atom inertial sensors based on stimulated Raman transitions. The drastic improvement
in terms of size, simplicity and power consumption paves the way towards field and mobile operations.
PACS numbers: Valid PACS appear here
Keywords: Suggested keywords
I.
INTRODUCTION
Over the past two decades atom interferometry has
kept supporting new concepts and experimental set-ups
to perform high precision measurements1 . In particular light pulse atom interferometry using stimulated Raman transitions has been turned into a very powerful tool
for both fundamental and applied sciences, especially to
measure inertial forces such as the acceleration due to
gravity2. Since the first realization of such inertial sensors3 , their accuracy and their stability have been continuously improved4,5 . Instruments of this kind now directly
compete with “classical” devices6 . According to their
principle of operation, the sensitivity and the accuracy
of the measurement rely on the control of the absolute
frequencies and the phase difference of the Raman lasers.
Indeed, it has been shown7 that the atomic phase difference at the interferometer output only depends on the
optical phase difference between the two Raman lasers,
at the position of the center-of-mass of the matter-wave
packets, at the time of the light pulses. In particular,
the measurement sensitivity relies on the performance of
the microwave reference signal used to control the phase
difference between the Raman laser.
As such experiments have been transformed into reliable
instruments, their field deployment becomes more than
relevant8 . For example, accelerometers relying on atom
interferometry have already demonstrated to be transportable9 and robust enough for airborne operations10 .
However some utilizations for fundamental science11 or
more applied studies12 would require a higher level of integration. Progress has been made to simplify the sensor
head13 and the laser system14 , but the same level of improvement is still to be made for the micro-wave reference
a) Electronic
mail: [email protected]
unit. Before this work, the latter was directly adapted
from the developments carried out for atomic fountain
micro-wave clocks15 . The modules currently used for
atom interferometers are bulky and feature a high power
consumption. They thus remain one of the last barriers
to full mobility.
Our synthesizer results from a trade-off between mobility
and high sensitivity measurement. We took into account
the specific frequency response of atom interferometers16 ,
as well as their main limitation due to ground vibrations17 . Requirements on the residual phase noise and
the frequency stability have been derived in the case of a
Mach-Zender-like π/2 − π − π/2 interferometer with 100
ms of interrogation time, as a guideline for the system
specifications.
The paper is organized as follows: first we describe the
main characteristics of our method and the system architecture we implemented. We then evaluate the performances of our set-up compared to the frequency standards operating at LNE-SYRTE in order to validate our
design. We finally assess what sensitivity to acceleration
an atomic interferometer can achieve using our synthesizer, and compare this to the state-of-the-art.
II.
DESIGN AND INTEGRATION
A. Requirements for the operation of a Raman transition
based atom interferometer
Atom interferometers relying on stimulated Raman
transitions employ a micro-wave signal to set the frequency difference between the two lasers at the hyperfine transition frequency of the atoms. In order to further simplify the design, we apply a method previously
demonstrated in one of our gravimeters4. It consists in
using the same micro-wave reference signal during the
Submitted to ‘ Chinese Physics C’
Influence of the Working Gas Properties on the Anode Wire
Modulation Effect of MWPC
WANG Xiao-hu
(Fundamental Science on Nuclear Waste and Environmental Security Laboratory, Southwest University of Science and
Technology,Mianyang,621010)
Abstract
For MWPC used for X-ray position detection, simulation study of the anode wire modulation effect of the
detector was carried out with Garfield program. Different gas mixtures were used as the working gas in the simulation, so as to
obtain the influence of the X-ray cross section and electron diffusion coefficient of the working gases on the anode wire
modulation effect of MWPC which has a anode wire spacing of 2mm. Results show that, though working gas with higher X-ray
cross section implies a larger average drift distance of ionized electrons, using of such gas mixtures is of little use to improve the
anode wire modulation effect of MWPC. And the transverse electron diffusion coefficient is the determining factor that affects
the extent of the anode wire modulation effect of the detector.
Keywords
PACS
Garfield simulation, anode wire modulation, working gas, MWPC
29.40.Cs, 29.40.Gx
thickness of the working gas volume, but a thick
1 Introduction
Position
sensitive
working gas volume implies much more parallax
detectors
based
on
error for planar type gas detectors. For X-ray
Multi-wire Proportional Chambers (MWPCs)
detection, since the attenuation of the X-ray
are widely used in neutron & X-ray detection.
beam in matter obeys the I=I0e-ud rule, most
MWPC own many advantages for particle
X-ray will be absorbed near the entrance
detection, such as stability, cheap, large area can
window if working gas with high X-ray cross
be made, etc. However, because the anode plane
section was used, and this means most ionized
of MWPC is made of a set of parallel wires, the
electron will experience a longer drift distances
anode wire modulation effect is a main factor
than these in working gas with low X-ray cross
that limits the spatial resolution in the direction
sections. Therefore, applying working gas with
across the anode wires of the detector.
high X-ray cross section might be a way to
There are many factors that affect the
reduce the anode wire modulation effect of
magnitude of anode wire modulation effect of
MWPC. Meanwhile, since the diffusion of the
MWPC, including anode wire spacing, drift
electrons can affect the sharing of ionized
distance of the ionized electrons, and electric
electrons on the anode wires, working gas with
field in the drift region, etc. As mentioned in
different diffusion coefficient corresponds to
Ref[1,2], by applying a cathode wire plane with
different anode wire modulation effect of
wires parallel and in registration with the anode
MWPC. Till now, few works have focused on
wires, the anode wire modulation effect and be
the influence of these gas properties on the
reduced significantly. Results in Ref[2, 3] also
anode wire modulation effect of MWPC.
show that, the deeper the drift region is, the
In this paper, the anode wire modulation
smaller anode wire modulation effect the
effect of a MWPC with anode wire spacing of
detector performs. The simplest way to obtain
2mm was studied. The influence of the electron
larger drift region depth is to increase the
diffusion coefficient and X-ray cross section on
arXiv:1501.01626v1 [physics.ins-det] 7 Jan 2015
Characterization of a Spherical Proportional
Counter in argon-based mixtures
F.J. Iguaz∗, A. Rodríguez, J.F. Castel and I.G. Irastorza
Laboratorio de Física Nuclear y Astropartículas, Universidad de Zaragoza, Spain.
E-mail: [email protected]
The Spherical Proportional Counter is a novel type of radiation detector, with a low energy threshold (typically below 100 eV) and good energy resolution. This detector is being developed by the
network NEWS, which includes several applications. We can name between many others Dark
Matter searches, low level radon and neutron counting or low energy neutrino detection from
supernovas or nuclear reactors via neutrino-nucleus elastic scattering. In this context, this works
will present the characterization of a spherical detector of 1 meter diameter using two argon-based
mixtures (with methane and isobutane) and for gas pressures between 50 and 1250 mbar. In each
case, the energy resolution shows its best value in a wide range of gains, limited by the ballistic
effect at low gains and by ion-backflow at high gains. Moreover, the best energy resolution shows
a degradation with pressure. These effects will be discussed in terms of gas avalanche properties.
Finally, the effect of an electrical field corrector in the homogenity of the gain and the energy
threshold measured in our setup will be also discussed.
Technology and Instrumentation in Particle Physics 2014
2-6 June, 2014
Amsterdam, the Netherlands
∗ Speaker.
c Copyright owned by the author(s) under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike Licence.
http://pos.sissa.it/
F.J. Iguaz
The Spherical Proportional Counter
1. Introduction
An active field in instrumentation for Particle Physics is the development of massive detectors
with low background levels and low energy threshold. The range of applications is diverse: the
search of dark matter in the form of Weakly Interacting Massive Particles (WIMPs) [1] or axions
[2], the detection of low-energy neutrinos, radon and neutron counting. In this context, the Spherical Proportional Counter (SPC) was proposed in [3] as a candidate fulfilling all the requirements: a
single readout channel reading a big gas volume (and mass); a potentially low intrinsic background
level with a good selection of radiopure materials; and a high signal-to-noise ratio due to its low
capacitance, which is proportional to the radius of the central electrode and does not depend on the
vessel size. Indeed, previous works have shown that an energy threshold as low as 100 eV [4, 5] is
feasible, keeping a good energy resolution (10% FWHM at 22.1 keV).
Even if there is only one channel, the pulse risetime can be used to discriminate complicate
topologies like muons from point-like events like x-rays; or to set a fiducial volume, as the risetime
depends on the radial position where the energy was deposited by gas diffusion. As an example of
this feature, we present in figure 1 (left) the dependence of the risetime with the pulse amplitude
for our SPC irradiated by a 241 Am source covered by an aluminum foil. As the mean free path of
photons is longer than the sphere’s radius at low pressures, an x-ray may deposit its energy at any
distance. As a result, it will create a distribution in risetime for a fixed amplitude in the 2D plot.
This fact happens for the x-rays of the source (at 11.9, 13.9, 17.7 and 20.8 keV). In contrast, the
fluorescence lines emitted from the stainless steel vessel (chromium at 5.5 keV and iron at 6.4 keV)
and the foil (aluminum at 1.5 keV) form three spots at long risetimes. If we select the events whose
risetime ranges in 4.5-8.7 µs (red lines), we obtain the energy spectrum of figure 1 (right).
The advantages of SPCs have motivated several feasibility studies [5 – 9] and the groups interested in this technology have created the network NEWS (New Experiments With Spheres), which
is now formed by institutes of France, Greece, China and Spain. In this context, the scalability of
high gains and good energy resolutions to higher masses (pressures) is an open question, which is
a key-point to increase the sensitivity of the detector to any particular application. This work tries
to give a first answer, presenting the characterization of a SPC in two argon-based mixtures (with
methane and isobutane) and for gas pressures up to 1250 mbar. The setup, the data-taking procedure and the analysis are described in detail in section 2. We also present the two configurations of
the cental rod (without and with a field corrector), whose main results will be respectively detailed
in sections 3 and 4. Finally, we finish in section 5 with some conclusions and an outlook.
2. Detector description
The detector consists of a large spherical stainless steel vessel 1.0 meter in diameter and a
small metallic ball 5 mm in diameter kept at the center by a stainless steel rod. The radiation from
the calibration source or other sources (like muons and gammas) ionize the gas. As the central
ball is set to a positive voltage and the vessel to ground, electrons drift to the ball where an intense
electric field amplifies the charge. The charge movement induces a signal at the electrode, which
is extracted by a teflon-based high voltage feed-through from the vessel. The signal is decoupled
2
JLAB-THY-15-1999
Triangle Singularities and XYZ Quarkonium Peaks
Adam P. Szczepaniak1, 2, 3
1
arXiv:1501.01691v1 [hep-ph] 8 Jan 2015
3
Department of Physics, Indiana University, Bloomington, IN 47405, USA
2
Theory Center, Thomas Jefferson National Accelerator Facility,
12000 Jefferson Avenue, Newport News, Virginia 23606, USA
Center for Exploration of Energy and Matter, Indiana University, Bloomington, IN 47403, USA
We discuss analytical properties of partial waves derived from projection of a 4-legged amplitude
with crossed-channel exchanges in the kinematic region of the direct channel that corresponds to the
XYZ peaks in charmonium and bottomonium. We show that in general partial waves can develop
anomalous branch points in the vicinity of the direct channel physical region. In a specific case,
when these branch points lie on the opposite side of the unitary cut they pinch the integration
contour in a dispersion relation and if the pinch happens close to threshold, the normal threshold
cusp is enhanced. We show that this effect only occurs if masses of resonances in the crossed channel
are in a specific, narrow range. We estimate the size of threshold enhancements originating from
these anomalous singularities in reactions where the Zc (3900) and the Zb (10610) peaks have been
observed.
I.
INTRODUCTION
There is significant interest in the physics of heavy quarkonia stimulated by discoveries of narrow peaks in the
spectrum. Such peaks may indicate existence of new hadrons. A recent review of the experimental situation and
of the various theoretical models can be found, for example, in [1]. While the quark model provides a remarkably
accurate description of the heavy quarkonium spectrum, these new peaks, also referred to as the XYZ states, appear
at masses that do not, in a natural way, derive from the quark model [2]. Several of these peaks have been observed
in invariant mass distributions of meson pairs that contain one heavy quarkonium, e.g. the J/ψ or the Υ, and one
light meson. The possibility that these new hadrons are therefore multi-quark bound states has been explored, for
¯ ∗ or
example in [3, 4]. The peaks appear near thresholds for production of meson pairs with open flavor, e.g. DD
∗
¯ , and for this reason it has also been suggested that binding between the two flavored mesons may be responsible
BB
for generating some of the XYZ’s [5–8]. It should be recognized, however, that distinction between multi-quark and
hadron bound states is complicated [9] and in any case requires sophisticated amplitude analysis and precise data [10].
It has also been suggested [11] and studied in [12–14] that coupling to nearby open channels may produce peaks even
without presence of new hadrons. In the mathematical language of amplitude analysis the analogous statement would
be that threshold cusp may be enhanced not only by a nearby pole but by a carefully arranged set of branch points.
Amplitude poles correspond to physical particles (stable or unstable) while cuts represent the effect of ”forces”, i.e.
exchanges of known particles in the crossed-channels. If the XYZ peaks were due to the latter it could potentially
weaken the case for new hadrons. To address this issue it is therefore important to perform a systematic analysis of
the ”forces” and the associated singularities.
II.
THE MODEL
We consider a decay of a heavy object of mass M , hereafter referred to as the M particle, to a quasi stable heavy
state of mass M 0 and a pair of degenerate, light mesons of mass µ. We are interested in the energy dependence of the
low spin partial waves, in particular the S-waves, in the M 0 µ channel. In this section we discuss a generic case and
in the section that follows we consider threshold behavior in J/ππ and Υ(1S)π production from Y (4260) → J/ψππ
[15–17] and Υ(5S) → Υ(1S)ππ [18, 19] decays, respectively.
The reaction of interest is M (p1 ) → M 0 (p3 ) + µ(p2 ) + µ(p4 ) with pi referring to the 4-vectors. Ignoring spin, the
reaction amplitude, A(s, t) is a scalar function of two independent Lorentz invariant Mandelstam variables, which we
choose as, s = (p1 − p2 )2 and t = (p1 − p3 )2 . They correspond to the invariant mass squared of the M 0 (p3 )µ(p4 ) pair
and momentum transfer squared between the two heavy mesons, respectively. The amplitude in the kinematics of the
decay region, s > (M 0 + µ)2 , t > 4µ2 can be obtained by analytical continuation in t of the amplitude describing the
s-channel scattering process, M (p1 ) + µ(−p2 ) → M 0 (p3 ) + µ(p4 ). It is the t-channel singularities of the latter that
are responsible for anomalous singularities of the s-channel partial waves. Similar analysis applies to the u-channel
where u = (p1 − p4 )2 . Bose symmetry requires the amplitude to be s ↔ u symmetric.
The s-channel partial waves, Al (s), describe production of the M 0 µ system in a state of fixed angular momentum, l.
Ignoring spin of the external particles, partial waves are the coefficients in expansion of A(s, t) in a series of Legendre
Supersymmetry after the Higgs
Pran Nath∗
arXiv:1501.01679v1 [hep-ph] 7 Jan 2015
Department of Physics, Northeastern University, Boston, MA 02115, USA
A brief review is given of the implications of a 126 GeV Higgs boson for the discovery of supersymmetry. Thus a 126 GeV Higgs boson is problematic within the Standard Model because of
vacuum instability pointing to new physics beyond the Standard Model. The problem of vacuum
stability is overcome in the SUGRA GUT model but the 126 GeV Higgs mass implies that the average SUSY scale lies in the several TeV region. The largeness of the SUSY scale relieves the tension
on SUGRA models since it helps suppress flavor changing neutral currents and CP violating effects
and also helps in extending the proton life time arising from baryon and lepton number violating
dimension five operators. The geometry of radiative breaking of the electroweak symmetry and fine
tuning in view of the large SUSY scale are analyzed.Consistency with the Brookhaven gµ − 2 result
is discussed. It is also shown that a large SUSY scale implied by the 126 GeV Higgs boson mass
allows for light gauginos (gluino, charginos, neutralinos) and sleptons. These along with the lighter
third generation squarks are the prime candidates for discovery at RUN II of the LHC. Implication
of the 126 GeV Higgs boson for the direct search for dark matter is discussed. Also discussed are
the sparticles mass hierarchies and their relationship with the simplified models under the Higgs
boson mass constraint.
In 2012 the Large Hadron Collider (LHC) made a
landmark discovery of a new boson. Thus the CMS and
ATLAS collaborations discovered a boson with a mass
of ∼ 126 GeV [1–4]. It is now confirmed that this newly
discovered particle is the long sought after Higgs boson
[5–8] which plays a central role in the breaking of the
electroweak symmetry. While the observed particle is
the last missing piece of the Standard Model there are
strong indications that at the same time its discovery
portends discovery of a new realm of physics specifically
supersymmetry. Below we elaborate on this theme in
further detail.
the large SUSY scale implied by the Higgs boson mass.
The sparticle landscape is an important indicator of the
underlying fundamental theory and in Section VI we discuss the sparticle landscape after the Higgs boson discovery. The connection of this landscape to the so called
simplified models is also discussed. The implications of
the Higgs boson mass at 126 on the search for dark matter in direct detection is discussed in Section VII. Future
prospects are discussed in Section VIII.
The outline of the rest of the paper is as follows: In
Section I we discuss the status of the Higgs boson in
the Standard Model, the issue of vacuum instability and
the need for new physics beyond the Standard Model.
In Section II we consider the implications of a 126 GeV
Higgs boson within the framework of supersymmetry and
specifically in the framework of supergravity unified models. As is well known a 126 GeV Higgs boson within supersymmetry leads to a high SUSY scale Ms with Ms lying in the TeV region. On the other hand the Brookhaven
gµ −2 experiment [9] shows a 3σ deviation from the Standard Model prediction [10, 11]. An effect of this size requires that the average scale of sparticle masses entering
the loops in the supersymmetric electroweak correction
to gµ −2 be low, i..e, O(100) GeV. Assuming the 3σ effect
is robust we discuss in Section III how to reconcile the
high SUSY scale that is indicated by the 126 GeV Higgs
boson mass with the low SUSY scale indicated by the
Brookhaven experiment. In Section IV we discuss the
implications of the Higgs boson mass and the geometry
of radiative electroweak symmetry breaking (REWSB).
In Section V we discuss the issue of fine tuning in view of
Within the Standard Model a Higgs boson mass of
∼ 126 GeV is problematic. Thus renormalization group
analyses show that vacuum stability up to the Planck
scale can be excluded at the 2σ level for the Standard
Model for Mh < 126 GeV [12] as illustrated in Fig. 1.
Here one finds that the quartic Higgs boson coupling
within the Standard Model turns negative at a scale
around 1011 GeV making the vacuum unstable [12].
However, as exhibited in Fig. 1 the result is rather sensitive to the mass of the top quark (see also [13]). Thus
lower values of the top mass tend to stabilize the vacuum
while the higher values tend to destabilize it. The current value of the top. i.e., Mt = 173.21 ± 0.51 ± 0.71 GeV
suggests vacuum instability. Thus the discovery of ∼ 126
GeV Higgs boson suggests the need for new physics beyond the Standard Model. Such new physics could be
supersymmetry and from here on we focus on this possibility.
∗
I.
II.
HIGGS BOSON AND NEW PHYSICS
126 GEV HIGGS BOSON WITHIN SUSY
In contrast to the case of the Standard Model, in supergravity unified models (SUGRA GUT) with the minimal supersymmetric standard model (MSSM) particle
Email: [email protected]
8
FIG. 10. An illustration of the probe of high SUSY scales
using the electron EDM. The plot exhibits the electron
EDM as a function of m0 for different values of the phase
αµ of the Higgs mixing parameter µ. The curves are
for the cases αµ = −3 (small-dashed, red), αµ = −0.5
(solid), αµ = 1 (medium-dashed, orange), and αµ = 2.5
(long-dashed, green). The other parameters are |µ| = 4.1 ×
102 , |M1 | = 2.8 × 102 , |M2 | = 3.4 × 102 , |Ae | = 3 ×
106 , mν0˜ = 4 × 106 , |Aν0˜ | = 5 × 106 , tanβ = 30 . All masses
are in GeV, phases in rad and EDM in ecm.The analysis
shows that improvements in the electron EDM constraint
can probe scalar masses in the 100 TeV- 1 PeV region and
beyond. The top horizontal line is the current experimental
limit from the ACME Collaboration [124]. From [125].
Further precision experiments can allow one to probe
even higher SUSY mass scales. [125–129]. One example
is to use EDMs as a probe of high SUSY scales. Thus
the electron EDM is most stringently constrained by the
ACME Collaboration [124] which gives
|de | < 8.7 × 10−29 ecm .
(13)
Fig. 10 exhibits the dependence of the electron EDM
on m0 for a number of CP phases of the Higgs mixing
parameter. The electron EDM limit is likely to improve
by an order of magnitude in the coming years and thus
the future EDM measurements will allow one to extend
the probe of new physics up to a PeV or more as shown
in Fig. 10. One can also use the precision measurement
of g − 2 of the electron as a sensitive probe of new
physics [130, 131].
In summary the discovery of the Higgs boson and
the measurement of its mass at ∼ 126 GeV has very
significant implications for new physics beyond the
Standard Model. The fact that a 126 GeV Higgs boson
mass in the framework of the Standard Model makes the
[1] CMS Collaboration, Phys. Lett. B 716 (2012) 30–61,
arXiv:1207.7235.
[2] ATLAS Collaboration, Phys. Lett. B 716 (2012)
1–29, arXiv:1207.7214.
[3] CMS Collaboration, Science 338 (2012) 1569–1575.
vacuum unstable provides yet another reason why new
physics beyond the Standard Model must exist. The
most promising candidate for such physics is supersymmetry. Specifically within the concrete framework of
supergravity grand unification one finds that the Higgs
boson mass is predicted to lie below ∼ 130 GeV. The fact
that the observed HIggs boson mass respects this bound
is a significant support for SUGRA GUT. Further, the
Higgs boson mass of ∼ 126 GeV requires the average
SUSY scale to be high, i.e., in the TeV region. This high
scale explains why we have seen no significant deviation
from the Standard Model prediction in FCNC processes
such as b → sγ and Bs → µ+ µ− . Further, the same high
SUSY scale explains
√
√ the non-observation of sparticles in
s = 7 TeV and s = 8 TeV data at RUN I of the LHC.
As discussed in Section III one important issue pertains
to the Brookhaven experiment which sees a 3σ deviation
in gµ −2 from the Standard Model prediction. This effect
is difficult to understand within the supergravity model
with universal boundary conditions. However, it is not
difficult to explain the observed phenomenon within supergravity unified models with non-universal boundary
conditions. Here it is possible to have light electroweak
gauginos and light sleptons while the squarks are heavy.
In this case one can explain the Brookhaven gµ − 2 result
as well as achieve a Higgs boson mass consistent with
experiment. The discovery of the Higgs boson mass is
important not only because one has found the last missing piece of the Standard Model but also because it is
likely the first piece of a new class of models such as
supersymmetric models which require the existence of a
whole new set of particles. It is hoped that LHC RUN II
will reveal some of these.
ACKNOWLEDGMENTS
This research is supported in part by grants NSF
grants PHY-1314774 and PHY-0969739, and by XSEDE
grant TG-PHY110015. This research used resources of
the National Energy Research Scientific Computing Center, which is supported by the Office of Science of the U.S.
Department of Energy under Contract No. DE-AC0205CH11231
[4] ATLAS Collaboration, Science 338 (2012)
1576–1582.
[5] F. Englert and R. Brout, Phys. Rev. Lett. 13 (1964)
321–323.
[6] P. W. Higgs, Phys. Lett. 12 (1964) 132–133.
Angular momentum decomposition from a QED example
Tianbo Liu1, 2, ∗ and Bo-Qiang Ma1, 3, 4, †
arXiv:1412.7775v1 [hep-ph] 25 Dec 2014
1
School of Physics and State Key Laboratory of Nuclear Physics and Technology, Peking University, Beijing 100871, China
2
INFN, Sezione di Pavia, via Bassi 6, 27100 Pavia, Italy
3
Collaborative Innovation Center of Quantum Matter, Beijing, China
4
Center for High Energy Physics, Peking University, Beijing 100871, China
We investigate the angular momentum decomposition with a quantum electrodynamics example to
clarify the proton spin decomposition debates. We adopt the light-front formalism where the parton
model is well defined. We prove that the sum of fermion and boson angular momenta is equal to
half the sum of the two gravitational form factors A(0) and B(0), as is well known. However, the
suggestion to make a separation of the above relation into the fermion and boson pieces, as a way
to measure the orbital angular momentum of fermions or bosons, respectively, is not justified from
our explicit calculation.
PACS numbers: 11.15.-q, 12.20.-m, 13.40.-f, 14.20.Dh
Angular momentum decomposition of a relativistic
composite system is a fundamental problem in physics
and is one of the most active frontiers in recent years.
Although the total angular momentum of an isolated
system is well defined, its decomposition into the spin
and orbital angular momentum of each component in interaction theories is nontrivial and of great interest. In
quantum chromodynamics (QCD) where we have no free
quarks or gluons, it is challenging to relate each term in
the decomposition, especially the orbital angular momentum (OAM), to physical observables. This is of physical
significance in understanding the nucleon structure.
The famous “proton spin crisis,” i.e., the observation
that the quark spin only contributes a small fraction [1]
(about 30% from recent analyses [2]) to the proton spin,
puzzled the whole physics society. This result severely
deviates from the naive quark model. A straightforward
understanding is to attribute the remaining proton spin
to OAM and/or gluon helicity. Because of the Wigner
rotation [3] that relates spin states between instant form
and light-front form (or those between rest frame and
infinite momentum frame) [4], one can obtain a nonvanishing OAM contribution from the extension of a nonrelativistic s-wave quark model to a relativistic light-front
treatment [5]. Therefore, the measurement of the OAM is
important, though the gluon may also contribute a large
portion to the proton spin [6]. However, the definition of
OAM in a gauge field theory is still under debate.
A most intuitive decomposition, given by Jaffe and
Manohar [7], breaks the gauge invariance and therefore
seems unmeasurable. Then a manifest gauge invariant
decomposition was proposed [8], and in this decomposition the total angular momentum of each parton flavor is
related to the sum of two gravitational form factors, A(0)
and B(0), which can be measured through deeply virtual
Compton scattering (DVCS) processes. These kinds of
∗
†
[email protected]
[email protected]
relations can shed light on the measurement of OAM if
they are generally valid, but we will find in this report
that this relation is unjustified.
By splitting the gauge potential into pure gauge and
physical terms, Chen et al. suggested a decomposition [9]
in which the operators of each term are gauge invariant
and satisfy the angular momentum commutation relations. Then many decomposition versions are proposed
with this approach [10]. In these decompositions, it
seems that a special gauge in which the pure gauge term
vanishes is still implied. In fact, it comes from the socalled Stueckelberg symmetry, which copies the group
of gauge symmetry but acts on the fields in a different
manner, in separating the pure gauge term. Hence the
approach of Chen et al. can be viewed as a gauge invariant extension (GIE) based on a Stueckelberg symmetry
fixing [11]. This fixing procedure is essentially a choice of
the separation for the pure gauge and physical terms, and
thus may result in different decomposition versions which
actually correspond to different physical objects [12]. Because of the longitudinal boost invariance, the light-front
gauge motivated choice is favored by the parton language.
Nowadays, all of the decompositions are usually divided
into two classes [13], the canonical class, e.g., Jaffe and
Manohar’s (JM’s), and the kinetic or mechanical class,
e.g., Ji’s. Both of them share the same term for fermion
spin, and the main difference between them is the definition of fermion OAM. From the GIE procedure, the
OAMs defined in both classes are, in principle, measurable without gauge invariance breaking, but the connection of the OAM to physical observables is still open to
challenge.
Recently, some quark model calculations show that
canonical and kinetic OAMs are different even in no
gauge field models [14, 15], in which cases one believes
that the two definitions should coincide with each other
and give the same results. In this report, we perform
explicit calculations in quantum electrodynamics (QED)
in order to avoid model assumptions. Since perturbative
calculations with QED are quite reliable and have been
precisely tested, it is an ideal theoretical laboratory to
The PADME experiment at LNF
Mauro Raggi1∗, Venelin Kozhuharov2 and P. Valente3
Laboratori Nazionali di Frascati - INFN, Frascati (Rome), Italy
2
University of Sofia “St. Kl. Ohridski”, Sofia, Bulgaria
3
INFN Sezione di RomaI, P.le Aldo Moro, 2 - 00185 Roma, Italy
1
arXiv:1501.01867v1 [hep-ex] 8 Jan 2015
7/01/2015
to be published in the DHF2014 proceedings
Abstract
Massive photon-like particles are predicted in many extensions of the Standard Model. They
have interactions similar to the photon, are vector bosons, and can be produced together with
photons. The PADME experiment proposes a search for the dark photon (A0 ) in the e+ e− → γA0
process in a positron-on-target experiment, exploiting the positron beam of the DAΦNE linac at
the Laboratori Nazionali di Frascati, INFN. In one year of running a sensitivity in the relative
interaction strength down to 10−6 is achievable, in the mass region from 2.5 MeV < MA0 < 22.5
MeV. The proposed experimental setup and the analysis technique is discussed.
1
Introduction
For MeV-GeV scale dark matter the direct detection techniques are notoriously difficult. Nevertheless some of the most appealing dark matter scenarios predict the possibility of observing
new particles in such a small mass range. This is the case of models in which the new states
are hidden not because of their high mass but due to a very small coupling to the Standard
Model. Despite attaining the highest energy ever reached at accelerators, LHC has not yet
been able to provide evidence for WIMP like particles, strongly constraining this class of dark
matter models. This largely open field of GeV-scale dark matter has recently revived models
postulating the existence of a hidden sector[1] interacting through a messenger with the visible
one and offers a well-motivated opportunity for experimental exploration. Dark sector models
have been used to explain different anomalies recently observed in particles and astroparticle
physics: the excess of positrons in cosmic rays observed by PAMELA in 2008[2] and confirmed
by AMS[3] and the present three sigma discrepancy between experiment and theory in the muon
anomalous magnetic moment aµ = (gµ − 2)/2[4].
The simplest hidden sector model just introduces one extra U(1) gauge symmetry and a
corresponding gauge boson: the “dark photon” (DP ). As in QED, this will generate interactions
of the types
L ∼ g 0 qf ψ¯f γ µ ψf Uµ0 ,
(1)
where g 0 is the universal coupling constant of the new interaction and qf are the corresponding
charges of the interacting fermions. Not all the Standard Model particles need to be charged
under this new U(1) symmetry thus leading in general to a different (and sometimes vanishing)
interaction strength for quarks and leptons. In the case of zero U(1) charge of the quarks [5], the
new gauge boson can be directly produced in hadron collisions or meson decays. The coupling
constant and the charges can in alternative be generated effectively through the so called kinetic
∗
corresponding author: [email protected]
1
Belle preprint 2014-21
KEK preprint 2014-37
arXiv:1501.01702v1 [hep-ex] 8 Jan 2015
¯ → Xs+d γ
Measurement of the direct CP asymmetry in B
decays with a lepton tag
L. Pes´
antez,3 P. Urquijo,35 J. Dingfelder,3 A. Abdesselam,55 I. Adachi,12, 9 K. Adamczyk,44 H. Aihara,60
S. Al Said,55, 26 K. Arinstein,4 D. M. Asner,47 V. Aulchenko,4 T. Aushev,37, 21 R. Ayad,55 S. Bahinipati,14
A. M. Bakich,54 V. Bansal,47 E. Barberio,35 V. Bhardwaj,40 B. Bhuyan,15 A. Bobrov,4 A. Bondar,4 G. Bonvicini,65
5
ˇ
A. Bozek,44 M. Braˇcko,33, 22 T. E. Browder,11 D. Cervenkov,
V. Chekelian,34 A. Chen,41 B. G. Cheon,10
K. Chilikin,21 R. Chistov,21 K. Cho,27 V. Chobanova,34 Y. Choi,53 D. Cinabro,65 J. Dalseno,34, 57 Z. Doleˇzal,5
Z. Dr´
asal,5 A. Drutskoy,21, 36 D. Dutta,15 S. Eidelman,4 H. Farhat,65 J. E. Fast,47 T. Ferber,7 O. Frost,7
V. Gaur,56 N. Gabyshev,4 S. Ganguly,65 A. Garmash,4 D. Getzkow,8 R. Gillard,65 Y. M. Goh,10 B. Golob,31, 22
J. Haba,12, 9 J. Hasenbusch,3 H. Hayashii,40 X. H. He,48 A. Heller,24 T. Horiguchi,59 W.-S. Hou,43
M. Huschle,24 T. Iijima,39, 38 K. Inami,38 A. Ishikawa,59 R. Itoh,12, 9 Y. Iwasaki,12 I. Jaegle,11 D. Joffe,25
T. Julius,35 K. H. Kang,29 E. Kato,59 T. Kawasaki,45 C. Kiesling,34 D. Y. Kim,52 J. B. Kim,28 J. H. Kim,27
K. T. Kim,28 M. J. Kim,29 S. H. Kim,10 Y. J. Kim,27 B. R. Ko,28 P. Kodyˇs,5 S. Korpar,33, 22 P. Kriˇzan,31, 22
P. Krokovny,4 B. Kronenbitter,24 T. Kuhr,24 T. Kumita,62 A. Kuzmin,4 Y.-J. Kwon,67 J. S. Lange,8
I. S. Lee,10 Y. Li,64 L. Li Gioi,34 J. Libby,16 D. Liventsev,12 P. Lukin,4 D. Matvienko,4 K. Miyabayashi,40
H. Miyata,45 R. Mizuk,21, 36 G. B. Mohanty,56 A. Moll,34, 57 H. K. Moon,28 E. Nakano,46 M. Nakao,12, 9
T. Nanut,22 Z. Natkaniec,44 M. Nayak,16 C. Ng,60 N. K. Nisar,56 S. Nishida,12, 9 S. Ogawa,58 S. Okuno,23
S. L. Olsen,51 C. Oswald,3 P. Pakhlov,21, 36 G. Pakhlova,21 C. W. Park,53 H. Park,29 T. K. Pedlar,32 R. Pestotnik,22
M. Petriˇc,22 L. E. Piilonen,64 E. Ribeˇzl,22 M. Ritter,34 A. Rostomyan,7 M. Rozanska,44 Y. Sakai,12, 9 S. Sandilya,56
L. Santelj,12 T. Sanuki,59 Y. Sato,38 V. Savinov,49 O. Schneider,30 G. Schnell,1, 13 C. Schwanda,18 A. J. Schwartz,6
K. Senyo,66 O. Seon,38 M. E. Sevior,35 V. Shebalin,4 C. P. Shen,2 T.-A. Shibata,61 J.-G. Shiu,43 B. Shwartz,4
A. Sibidanov,54 F. Simon,34, 57 Y.-S. Sohn,67 A. Sokolov,19 E. Solovieva,21 M. Stariˇc,22 M. Steder,7 T. Sumiyoshi,62
U. Tamponi,20, 63 N. Taniguchi,12 G. Tatishvili,47 Y. Teramoto,46 K. Trabelsi,12, 9 M. Uchida,61 T. Uglov,21, 37
Y. Unno,10 S. Uno,12, 9 Y. Usov,4 C. Van Hulse,1 P. Vanhoefer,34 G. Varner,11 A. Vinokurova,4 V. Vorobyev,4
M. N. Wagner,8 B. Wang,6 C. H. Wang,42 M.-Z. Wang,43 P. Wang,17 Y. Watanabe,23 K. M. Williams,64
E. Won,28 J. Yamaoka,47 S. Yashchenko,7 Y. Yook,67 Z. P. Zhang,50 V. Zhilich,4 V. Zhulanov,4 and A. Zupanc22
(The Belle Collaboration)
1
University of the Basque Country UPV/EHU, 48080 Bilbao
2
Beihang University, Beijing 100191
3
University of Bonn, 53115 Bonn
4
Budker Institute of Nuclear Physics SB RAS and Novosibirsk State University, Novosibirsk 630090
5
Faculty of Mathematics and Physics, Charles University, 121 16 Prague
6
University of Cincinnati, Cincinnati, Ohio 45221
7
Deutsches Elektronen–Synchrotron, 22607 Hamburg
8
Justus-Liebig-Universit¨
at Gießen, 35392 Gießen
9
The Graduate University for Advanced Studies, Hayama 240-0193
10
Hanyang University, Seoul 133-791
11
University of Hawaii, Honolulu, Hawaii 96822
12
High Energy Accelerator Research Organization (KEK), Tsukuba 305-0801
13
IKERBASQUE, Basque Foundation for Science, 48013 Bilbao
14
Indian Institute of Technology Bhubaneswar, Satya Nagar 751007
15
Indian Institute of Technology Guwahati, Assam 781039
16
Indian Institute of Technology Madras, Chennai 600036
17
Institute of High Energy Physics, Chinese Academy of Sciences, Beijing 100049
18
Institute of High Energy Physics, Vienna 1050
19
Institute for High Energy Physics, Protvino 142281
20
INFN - Sezione di Torino, 10125 Torino
21
Institute for Theoretical and Experimental Physics, Moscow 117218
Typeset by REVTEX
1
22
J. Stefan Institute, 1000 Ljubljana
Kanagawa University, Yokohama 221-8686
Institut f¨
ur Experimentelle Kernphysik, Karlsruher Institut f¨
ur Technologie, 76131 Karlsruhe
25
Kennesaw State University, Kennesaw GA 30144
26
Department of Physics, Faculty of Science, King Abdulaziz University, Jeddah 21589
27
Korea Institute of Science and Technology Information, Daejeon 305-806
28
Korea University, Seoul 136-713
29
Kyungpook National University, Daegu 702-701
30 ´
Ecole Polytechnique F´ed´erale de Lausanne (EPFL), Lausanne 1015
31
Faculty of Mathematics and Physics, University of Ljubljana, 1000 Ljubljana
32
Luther College, Decorah, Iowa 52101
33
University of Maribor, 2000 Maribor
34
Max-Planck-Institut f¨
ur Physik, 80805 M¨
unchen
35
School of Physics, University of Melbourne, Victoria 3010
36
Moscow Physical Engineering Institute, Moscow 115409
37
Moscow Institute of Physics and Technology, Moscow Region 141700
38
Graduate School of Science, Nagoya University, Nagoya 464-8602
39
Kobayashi-Maskawa Institute, Nagoya University, Nagoya 464-8602
40
Nara Women’s University, Nara 630-8506
41
National Central University, Chung-li 32054
42
National United University, Miao Li 36003
43
Department of Physics, National Taiwan University, Taipei 10617
44
H. Niewodniczanski Institute of Nuclear Physics, Krakow 31-342
45
Niigata University, Niigata 950-2181
46
Osaka City University, Osaka 558-8585
47
Pacific Northwest National Laboratory, Richland, Washington 99352
48
Peking University, Beijing 100871
49
University of Pittsburgh, Pittsburgh, Pennsylvania 15260
50
University of Science and Technology of China, Hefei 230026
51
Seoul National University, Seoul 151-742
52
Soongsil University, Seoul 156-743
53
Sungkyunkwan University, Suwon 440-746
54
School of Physics, University of Sydney, NSW 2006
55
Department of Physics, Faculty of Science, University of Tabuk, Tabuk 71451
56
Tata Institute of Fundamental Research, Mumbai 400005
57
Excellence Cluster Universe, Technische Universit¨
at M¨
unchen, 85748 Garching
58
Toho University, Funabashi 274-8510
59
Tohoku University, Sendai 980-8578
60
Department of Physics, University of Tokyo, Tokyo 113-0033
61
Tokyo Institute of Technology, Tokyo 152-8550
62
Tokyo Metropolitan University, Tokyo 192-0397
63
University of Torino, 10124 Torino
64
CNP, Virginia Polytechnic Institute and State University, Blacksburg, Virginia 24061
65
Wayne State University, Detroit, Michigan 48202
66
Yamagata University, Yamagata 990-8560
67
Yonsei University, Seoul 120-749
23
24
(Dated: January 9, 2015)
2
Precision Measurement of the Mass of the D ∗0 Meson and the Binding Energy of the
X(3872) Meson as a D 0 D ∗0 Molecule
A. Tomaradze,1 S. Dobbs,1 T. Xiao,1 and Kamal K. Seth1
arXiv:1501.01658v1 [hep-ex] 7 Jan 2015
1
Northwestern University, Evanston, Illinois 60208, USA
(Dated: January 9, 2015)
A precision measurement of the mass difference between
the D0 and D∗0 mesons has been made
√
−1
+ −
using 316 pb
of e e annihilation data taken at s = 4170 MeV using the CLEO-c detector.
We obtain ∆M ≡ M (D∗0 ) − M (D0 ) = 142.007 ± 0.015(stat) ± 0.014(syst) MeV, as the average
for the two decays, D0 → K − π + and D0 → K − π + π − π + . The new measurement of ∆M leads
to M (D∗0 ) = 2006.850 ± 0.049 MeV, and the currently most precise measurement of the binding
energy of the “exotic” meson X(3872) if interpreted as a D0 D∗0 hadronic molecule, Eb (X(3872)) ≡
M (D0 D∗0 ) − M (X(3872)) = 3 ± 192 keV.
Of all the claims and counterclaims for the so-called
“exotic” mesons, which do not fit in the pictures of
conventional q q¯ mesons [1], the most intriguing one is
X(3872). Its existence has been confirmed from numerous measurements, by Belle [2], CDF [3], D0 [4],
BaBar [5], LHCb [6] and CMS [7], and its mass, width,
and spin are respectively, M (X(3872)) = 3871.69 ±
0.17 MeV, Γ(X(3872)) < 1.2 MeV, and J P C = 1++ [8].
Although many different suggestions for the structure of
X(3872) exist in the literature [9–12], the closeness of
the X(3872) mass to the sum of the masses of the D0
and D∗0 mesons, and the smallness of its width have
made the suggestion that it is a weakly bound hadronic
molecule made of the D0 and D∗0 mesons extremely attractive [13]. To submit this provocative suggestion to
experimental test it is important to measure the binding
energy of X(3872), indeed to determine if it is bound at
all. This Letter reports on the results of just such a measurement. Throughout this Letter we use the PDG [8]
convention for units, with masses in MeV and momenta
in MeV/c, and inclusion of charge-conjugate states is implied.
A measurement of the binding energy of X(3872)
requires the knowledge of three masses, M (X(3872)),
M (D0 ), and M (D∗0 ), with the most accurately determined value of M (D∗0 ) obtained by measuring the mass
difference, ∆M ≡ M (D∗0 ) − M (D0 ). Since the discovery of X(3872) in 2003, the precision in the value
of the mass of the X(3872) has steadily improved from
±800 keV, originally, to the present average with error
of ±170 keV [8] because of numerous improved measurements. Similarly, the precision of the value M (D0 ) has
improved, from ±1000 keV, originally, to ±180 keV by
a CLEO measurement of M (D0 ) in 2007 [14], and to
±40 keV due to two recent higher-precision measurements of M (D0 ) by BaBar [15], and our recent publication [16]. As a consequence, the determination of
the binding energy of X(3872) as a D0 D∗0 molecule has
changed from (600 ± 600) keV in 2007 [14] to (126 ± 204)
keV in 2014 [16].
Through all these improvements, the mass difference
∆M has remained fixed at the value, ∆M = 142.120 ±
0.070 MeV as measured by CLEO in 1992 using data
taken at the Υ(4S) resonance [17]. To determine the
binding energy of X(3872) with the highest possible precision, it has become imperative to make a new higherprecision measurement of ∆M ≡ M (D∗0 ) − M (D0 ).
In this Letter we report on such a measurement using data taken at the ψ(4160) resonance, which decays
into D∗ D∗ , D∗ D, and DD. We analyze D∗0 → D0 π 0 ,
and D0 decays, D0 → K − π + (henceforth Kπ), and
D0 → K − π + π + π − (henceforth K3π).
−1
We
of e+ e− annihilation data taken
√ use 316 pb
at s = 4170 MeV with the CLEO-c detector. The
CLEO-c detector [19] consists of a CsI(Tl) electromagnetic calorimeter, an inner vertex drift chamber, a central
drift chamber, and a ring-imaging Cherenkov (RICH)
detector, all inside a superconducting solenoid magnet providing a 1.0 Tesla magnetic field. The acceptance for charged and neutral particles is | cos θ| <
0.93. Charged-particle momentum resolution is σp /p =
0.6% @ 1 GeV/c. Photon energy resolution is σE /E =
2.2% @ 1 GeV, and 5% @ 100 MeV. The detector response was studied using a GEANT-based [20] Monte
Carlo simulation.
We select events with well-measured tracks by requiring that they be fully contained in the barrel region
(| cos θ| < 0.8) of the detector, and have transverse momenta > 120 MeV/c.
In our previous article on the precision measurement
of the mass of the D0 meson [16], we made a precision
recalibration of the CLEO-c solenoid magnetic field and
determined a correction of (2.9 ± 0.4) × 10−4 in the default calibration of the CLEO-c magnetic field. The data
we use in the present investigation were taken just after
this recalibration. We use the same corrected field as
determined in our previous article in the present Letter.
Charged pions and kaons were identified using information from both the drift chamber dE/dx and the RICH
detector. First, it was required that the dE/dx of the
charged particle track be consistent within 3σ of the respective pion or kaon hypothesis. For tracks with mo-
tat
io
Pa nal T
rti he
cle or
Ph etic
ysi al
cs
B
AC
Co
m
pu
KA
SFB TR9
SFB/CPP-14-100
DESY 15-005
HU-15/02
Non-perturbative computation of the strong coupling constant on the lattice
arXiv:1501.01861v1 [hep-lat] 8 Jan 2015
Rainer Sommera,b , Ulli Wolffb
a John
von Neumann Institute for Computing (NIC), DESY, Platanenallee 6, 15738 Zeuthen, Germany
b Institut f¨
ur Physik, Humboldt Universi¨at, Newtonstr. 15, 12489 Berlin, Germany
Abstract
We review the long term project of the ALPHA collaboration to compute in QCD the running coupling constant
and quark masses at high energy scales in terms of low energy hadronic quantities. The adapted techniques required
to numerically carry out the required multiscale non-perturbative calculation with our special emphasis on the control
of systematic errors are summarized. The complete results in the two dynamical flavor approximation are reviewed
and an outlook is given on the ongoing three flavor extension of the programme with improved target precision.
Keywords: QCD, running coupling, quark mass, lattice QCD, Monte Carlo, Schr¨odinger functional, gradient flow
1. Introduction
Quantum Chromo Dynamics (QCD) is the renormalizable quantum field theory containing gluon and quark
fields that interact in a unique way dictated by SU(3)
gauge invariance. It may be seen as arising from the
standard model of elementary particles in a limit where
all other fields, including their interactions with quarks
and gluons, are stripped away. Strong interactions and
confinement are the characteristics of this sector which
hence calls for non-perturbative evaluations and is in the
focus of lattice formulations and simulations.
We here consider QCD with a free number of Nf color
triplets (flavors) of quark species. In Nature we see the
case Nf = 6 with the flavors up, down, strange, charm,
bottom and top in order of ascending mass. The species
beyond light up and down quarks come with characteristic scales of the order of 0.1 GeV, 1 GeV, 4 GeV,
175 GeV. Therefore it makes sense to consider effective theories with Nf < 6 to describe physics with characteristic energies significantly below the scales of the
dropped degrees of freedom. They then enter only indirectly into the determination of the free parameters of
the effective theory. In lattice simulations the modelling
of the precise flavor content is technically very demanding. Therefore a lot of studies are found and will also
be discussed here that refer to Nf = 2 and Nf = 3 where
the latter number is the minimum to allow for real applications as an effective theory [1, 2, 3, 4, 5]. In any
case generalized QCD has Nf + 1 free parameters given
by one quark mass per species and in addition the gauge
coupling.
The two light species are in most studies, including
those described here, approximated to be degenerate.
The value zero for some or even all Nf quark masses
is theoretically nice as it enhances the chiral symmetry of the model and is thus stabilized under renormalization. The renormalization of the coupling can be
defined in this massless limit and we then speak of a
massless renormalization scheme. Such schemes are
technically convenient in nontrivial perturbative as well
as non-perturbative calculations. The renormalization
of the coupling can be left unchanged as quark masses
are ‘turned on later’. To define a renormalized coupling
constant in a massless scheme an additional scale µ enters via the renormalization conditions. The resulting
scale dependent ‘running’ coupling g¯ (µ) obeys a Callan-
Parity violation in neutron capture on the proton:
determining the weak pion-nucleon coupling
J. de Vries1 , N. Li1 , Ulf-G. Meißner1,2,3,4 ,
A. Nogga1,2,3 , E. Epelbaum5 , N. Kaiser6
arXiv:1501.01832v1 [nucl-th] 8 Jan 2015
1
2
Institute for Advanced Simulation, Institut f¨
ur Kernphysik, and J¨
ulich Center for
Hadron Physics, Forschungszentrum J¨
ulich, D-52425 J¨
ulich, Germany
JARA - Forces and Matter Experiments, Forschungszentrum J¨
ulich, D-52425 J¨
ulich,
Germany
3
4
JARA - High Performance Computing, Forschungszentrum J¨
ulich, D-52425 J¨
ulich,
Germany
Helmholtz-Institut f¨
ur Strahlen- und Kernphysik and Bethe Center for Theoretical
Physics, Universit¨at Bonn, D-53115 Bonn, Germany
5
Institut f¨
ur Theoretische Physik II, Ruhr-Universit¨at Bochum, 44780 Bochum,
Germany
6
Physik Department T39, Technische Universit¨at M¨
unchen, D-85747 Garching,
Germany
Abstract
We investigate the parity-violating analyzing power in neutron capture on the proton
at thermal energies in the framework of chiral effective field theory. By combining this
analysis with a previous analysis of parity violation in proton-proton scattering, we are able
to extract the size of the weak pion-nucleon coupling constant. The uncertainty is significant
and dominated by the experimental error which is expected to be reduced soon.
Predicting positive parity Bs mesons from lattice QCD
C. B. Lang,1, ∗ Daniel Mohler,2, † Sasa Prelovsek,3, 4, ‡ and R. M. Woloshyn5, §
1
Institute of Physics, University of Graz, A–8010 Graz, Austria
Fermi National Accelerator Laboratory, Batavia, Illinois 60510-5011, USA
3
Department of Physics, University of Ljubljana, 1000 Ljubljana, Slovenia
4
Jozef Stefan Institute, 1000 Ljubljana, Slovenia
5
TRIUMF, 4004 Wesbrook Mall Vancouver, BC V6T 2A3, Canada
(Dated: January 9, 2015)
arXiv:1501.01646v1 [hep-lat] 7 Jan 2015
2
∗
We determine the spectrum of Bs 1P states using lattice QCD. For the Bs1 (5830) and Bs2
(5840)
mesons, the results are in good agreement with the experimental values. Two further mesons
are expected in the quantum channels J P = 0+ and 1+ near the BK and B ∗ K thresholds. A
combination of quark-antiquark and B (∗) meson-kaon structures are used to determine the mass of
(∗)
two QCD bound states below the B (∗) K threshold, with the assumption that mixing with Bs η and
(∗)
P
+
isospin-violating decays to Bs π are negligible. We predict a J = 0 bound state Bs0 with mass
mBs0 = 5.711(13)(19) GeV. With further assumptions motivated theoretically by the heavy quark
limit, a bound state with mBs1 = 5.750(17)(19) GeV is predicted in the J P = 1+ channel. The
results from our first principles calculation are compared to previous estimates based on models.
PACS numbers: 11.15.Ha, 12.38.Gc
Over the years experiments have uncovered a number
of mesons involving heavy quarks that do not seem to fit
the simple quark-antiquark picture suggested by quark
models. Examples of these include states in the charmonium and bottomonium spectrum [1] as well as the
∗
charm-strange Ds0
(2317) and Ds1 (2460) [2]. The latter
states are identified with the j = 21 heavy-quark multiplet
[3] and were predicted to be broad states above thresholds
∗
in potential models. However, the observed Ds0
(2317)
and Ds1 (2460) are narrow states below the DK or D∗ K
thresholds [2], and it has been suggested that the thresholds play an important role in lowering the mass of the
physical states [4]. In a recent lattice QCD simulation
[5–7] these states are seen as QCD bound states below
threshold with a mass in good agreement with experiment.
In the Bs meson spectrum only two positive parity
states are known from experiment [8–10], the Bs1 (5830)
∗
and Bs2
(5840). The LHCb experiment should be able
to see the remaining two states (0+ and 1+ ), which are
expected to decay into s-wave states by emitting either
a photon or a π 0 [11]. On the theory side there are a
number of phenomenological model and EFT mass determinations [11–18], a determination using Unitarized
EFT based on low energy constants extracted from lattice QCD simulations [19], and some lattice QCD calculations in the static limit [20–24]. The HPQCD collaboration has published a prediction [25] taking into account
explicitly only quark-antiquark interpolating fields and
extracting only the ground states in the system. This
strategy can lead to inaccurate results in the vicinity of
thresholds where meson-meson scattering can have a significant effect. None of the previous lattice simulations
clearly establish the states in question as either QCD
bound states below threshold or resonances above thresh-
old. It is this gap which we aim to fill with the current
publication.
In this letter we present results for masses of the pwave states of bottom-strange mesons with quantum
numbers J P = 0+ , 1+ , 2+ . For the heavy-quark doublet with j P = 23 masses determined using only quarkantiquark interpolating fields agree with those of the ob∗
served Bs1 (5830) and Bs2
(5840). This, as well as calculated mass differences between heavy-light mesons, verifies our computational setup. Then, using the extracted
pole positions of observed QCD bound states relative to
the B (∗) K thresholds, the bound state masses and estimates of systematic uncertainties are determined.
The gauge configurations are from the PACS-CS collaboration [26]. They have Nf = 2 + 1 dynamical quarks
(up/down, strange); the bottom quark is implemented as
a valence quark. The light and strange quarks are nonperturbatively improved Wilson fermions. The lattice
spacing is 0.0907(13) fm and the Pion mass is 156(7)(2)
MeV. Due to the large lattice size 323 × 64 and larger
physical volume we use stochastic distillation [27] for the
quark propagation as in our analysis of the Ds mesons
[5–7]. Further details including the u, d, and s quark
parameters can be found in [6].
The dynamic strange quark mass used in [26] differs
significantly from the physical value. We therefore use a
partially quenched strange quark mval
6= msea
s
s . Different
determinations agree very well and yield the value for
κs [6] which leads to the kaon mass mK = 504(1)(7) MeV.
The bottom quark is treated as a valence quark and
the Fermilab method [28, 29] is used. See Ref. [6, 30]
for details of our implementation. In the simplified form
that we use [31, 32], only the bottom quark hopping parameter κb is tuned non-perturbatively, while the clover
coefficients cE and cB are set to the tadpole improved
January 2015
arXiv:1501.01642v1 [astro-ph.CO] 7 Jan 2015
Anisotropies in Non-Thermal Distortions of
Cosmic Light from Photon-Axion Conversion
Guido D’Amicoa,1 and Nemanja Kaloperb,2
a
Center for Cosmology and Particle Physics, New York University, New York, NY 10003
b
Department of Physics, University of California, Davis, CA 95616
ABSTRACT
Ultralight axions which couple sufficiently strongly to photons can leave imprints on the sky
at diverse frequencies by mixing with cosmic light in the presence of background magnetic
fields. We explore such direction dependent grey-body distortions of the CMB spectrum,
enhanced by resonant conditions in the IGM plasma. We also find that if such axions are
produced in the early universe and represent a subdominant dark radiation component today,
they could convert into X-rays in supervoids, and brighten them at X-ray frequencies.
1
2
[email protected]
[email protected]
De-Confinement and Clustering of Color Sources in
Nuclear Collisions
arXiv:1501.01524v1 [nucl-th] 7 Jan 2015
M. A. Brauna , J. Dias de Deusb , A. S. Hirschc , C. Pajaresd , R. P.
Scharenbergc , B. K. Srivastavac,∗
a
Department of High Energy Physics, Saint-Petersburg State University, S. Petersburg,
Russia
b
CENTRA, Instituto Superior Tecnico, 1049-001 Lisboa, Portugal
c
Department of Physics and Astronomy, Purdue University, West Lafayette, IN-47907,
USA
d
Departamento de Fisica de Particulas, Universidale de Santiago de Compostela and
Instituto Galego de Fisica de Atlas Enerxias(IGFAE), 15782 Santiago, de Compostela,
Spain
Abstract
A brief introduction of the relationship of string percolation to the
Quantum Chromo Dynamics (QCD) phase diagram is presented. The behavior of the Polyakov loop close to the critical temperature is studied in terms of
the color fields inside the clusters of overlapping strings, which are produced
in high energy hadronic collisions. The non-Abelian nature of the color fields
implies an enhancement of the transverse momentum and a suppression of
the multiplicities relative to the non overlapping case. The prediction of this
framework are compared with experimental results from the SPS, RHIC and
LHC for pp and AA collisions. Rapidity distributions, probability distributions of transverse momentum and multiplicities, Bose-Einstein correlations,
elliptic flow and ridge structures are used to evaluate these comparison.
The thermodynamical quantities, the temperature, and energy density derived from RHIC and LHC data and Color String Percolation Model (CSPM)
are used to obtain the shear viscosity to entropy density ratio (η/s). It
was observed that the inverse of (η/s) represents the trace anomaly ∆ =
(ε − 3P )/T 4 . Thus the percolation approach within CSPM can be successfully used to describe the initial stages in high energy heavy ion collisions
∗
Corresponding author
Email address: [email protected] (B. K. Srivastava)
January 8, 2015
Simulating net particle production and chiral magnetic current in a CP-odd domain
Kenji Fukushima
Department of Physics, The University of Tokyo,
7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan
arXiv:1501.01940v1 [hep-ph] 8 Jan 2015
We elucidate the numerical formulation to simulate net production of particles and anomalous currents with CP-breaking background fields which cause an imbalance of particles over anti-particles.
For a concrete demonstration we numerically impose pulsed electric and magnetic fields to observe
that the dynamical chiral magnetic current follows together with the net particle production. The
produced particle density is quantitatively consistent with the axial anomaly, while the chiral magnetic current is suppressed by a delay before the the onset of the current generation.
Introduction: In many quantum problems in physics
it is highly demanded to establish a numerical framework
to simulate full dynamical processes of particle production out of equilibrium. Unsettled and important problems include the leptogenesis and the baryogenesis for
the explanation of the baryon asymmetry of the Universe
(BAU). It is well-known that Sakharov’s three conditions
are needed for the BAU; namely, B-breaking process, C
and CP violation, and out-of-equilibrium. The particle
production on top of the electroweak sphaleron [1] accommodates a process with ∆(B + L) 6= 0, which means
that B + L decays toward chemical-equilibrium [and this
is why the idea based on the SU(5) grand unified theory [2] does not work out for the BAU]. Therefore, if
the leptogenesis with some (B − L)-breaking process beyond the Standard Model (see Ref. [3] for example) generates L 6= 0 initially, it would amount to B 6= 0 finally
in chemical-equilibrium. The particle production associated with the sphaleron transition has a character of
topological invariance and the index theorem relates the
topological number to the change in the particle number. We must recall, however, that the sphaleron is a
special static saddle-point configuration and for general
gauge configurations with C and CP violation we have to
consider microscopic simulations with Weyl fermions.
The sphaleron transition rate is proportional to T 4
where T is the temperature [4], and so such topological fluctuations should be abundant also in the strong
interaction when the physical system is heated up to
T > ΛQCD as argued in Ref. [5]. Such high temperature is indeed realized in the quark-gluon plasma created
in the relativistic heavy-ion collision (HIC) experiment.
Although the strong interaction does not break CP (except for negligibly small strong θ), we can still anticipate
local violation of P and CP before thermalization [6].
This possibility is nowadays referred to commonly as the
local parity violation (LPV). Theoretical and experimental challenges are still ongoing about the LPV detection.
Along these lines the recognition of the interplay with
pulsed magnetic field B between two heavy ions was a
major breakthrough [7]; B on top of a CP-odd domain
would induce an electric current j in parallel to B, which
is summarized in a compact formula of the chiral mag-
y
B
y
jy
x
jz
jx
Bz
Ez
z
FIG. 1. Schematic view of currents induced by By in a CPodd domain realized by parallel Ez and Bz . The currents flow
in all the x, y, and z directions; jx is the Hall current and jy
is the CME current.
netic effect (CME) with chiral chemical potential µ5 [8].
In a realistic situation of the HIC we can anticipate P
and CP violation not only from the QCD sphalerons but
also from the glasma [9] which is a description of the
initial condition of the HIC in terms of coherent fields.
Then, without introducing µ5 , we can (and should) directly compute the anomalous currents associated with
the particle production in a CP-odd domain with chromoelectric and chromo-magnetic fields (see Ref. [10]).
It would be an intriguing attempt to test the CME
in a table-top experiment in a way similar to the quantum Hall effect. The biggest difference between the CME
and the Hall effect arises from the fact that the carriers
of electric charge are not the ordinary particles but the
chirality; that is the excess of the right-handed particle number to the left-handed particle number [11]. It
is quite non-trivial how to implement such a chiral battery (usually represented by µ5 6= 0) in materials. Just
recently the CME may have been confirmed in a system
rather similar to the glasma, in which chirality imbalance
is imposed by background fields that break CP symmetry [12].
Figure 1 illustrates the schematic view of the CME
setup in the HIC and in the condensed matter experiment. The parallel E and B (in the z direction in
Fig. 1) form P- and CP-odd product E · B, and the
CME current j y is induced in a direction perpendicular
arXiv:1501.01938v1 [hep-ph] 8 Jan 2015
New physics in ∆Γd
Gilberto Tetlalmatzi-Xolocotzi∗
Institute for Particle Physics and Phenomenology, Durham University
E-mail: [email protected]
We analyze the possibility of having new physics effects in the decay rate difference, ∆Γd , of
neutral Bd mesons. Three different sources of enhancement are considered, CKM unitarity violations, beyond standard model effects in the tree-level dimension six operators (d¯p)( p¯0 b) with
¯ ττ).
¯
p, p0 = u, c; and large enhancements of the almost unconstrained operators (db)(
We find
that deviations of several hundred per cent from the standard model prediction of ∆Γd are not
excluded by current experimental data.
Flavorful Ways to New Physics - FWNP,
28-31 October 2014
Freudenstadt - Lauterbad, Germany
∗ Speaker.
c Copyright owned by the author(s) under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike Licence.
http://pos.sissa.it/
Gilberto Tetlalmatzi-Xolocotzi
New physics in ∆Γd
1. Introduction
There is a longstanding discrepancy between experimental results for the like-sign dimuon
asymmetry measured by the D0 Collaboration [1]-[4] and the corresponding standard model predictions [5]-[9]. In [10] an interesting connection between the dimuon asymmetry and the decay
rate difference of neutral Bd mesons was suggested: the measured enhancement of the dimuon
asymmetry could also be explained by an enhanced decay rate difference of the neutral Bd mesons.
Moreover, ∆Γd , is currently only weakly constrained by direct measurements. This was the motivation for the study in [11] where indirect experimental constraints on possible new physics enhancements of ∆Γd were studied; we present here the main findings of [11].
2. Neutral B mixing
Due to electroweak interactions the neutral states Bd and B¯ d oscillate into each other, the time
evolution of this system is given by solving the Schrödinger-like equation
i
d
dt
!
!
|Bd i
|Bd i
= Σd
¯
¯
|Bd i
|Bd i
where
i
Σd = Md − Γd =
2
Σd
iΓd11
2
d∗
d∗ − iΓ12
M12
2
d −
M11
iΓd12
2
d
d − iΓ11
M11
2
d −
M12
!
.
(2.1)
MHd , MLd
Diagonalizing
gives the physical eigenstates {|BH >, |BL >} with the masses
and the
decay rates ΓdH , ΓdL . To provide a mathematical description of the mixing process it is useful to
define the observables ∆Md = MHd − MLd and ∆Γd = ΓdH − ΓdL . Theoretically ∆Md and ∆Γd can be
calculated from the components of Σd according to the formulas
d
∆Md ≈ 2|M12
|
and
∆Γd ≈ 2|Γd12 |cos(φd )
where
Md φd = arg − d12 .
Γ12
(2.2)
Only ∆Md and ∆Γd are directly accessible in experiment, the phase φd can be calculated from the
measurement of the semileptonic asymmetry
Γd adsl = 12
).
d sin(φd
M12
(2.3)
By combining data from Belle, BABAR, D0, DELPHI and LHCb the following direct experimental
bound for ∆Γd is available
∆ΓHFAG
d
= (0.1 ± 1.0)%[12]
Γd
which can be compared to the standard model prediction
∆ΓSM
d
= (0.42 ± 0.08)%[5].
Γd
2
(2.4)
New Paradigm for Baryon and Lepton Number Violation
Pavel Fileviez P´erez
Particle and Astro-Particle Division
Max Planck Institute for Nuclear Physics, 69117 Heidelberg, Germany
arXiv:1501.01886v1 [hep-ph] 7 Jan 2015
Abstract
The possible discovery of proton decay, neutron-antineutron oscillation, neutrinoless beta decay
in low energy experiments, and exotic signals related to the violation of the baryon and lepton
numbers at collider experiments will change our understanding of the conservation of fundamental symmetries in nature. In this review we discuss the rare processes due to the existence
of baryon and lepton number violating interactions. The simplest grand unified theories and
the neutrino mass generation mechanisms are discussed. The theories where the baryon and
lepton numbers are defined as local gauge symmetries spontaneously broken at the low scale are
discussed in detail. The simplest supersymmetric gauge theory which predicts the existence of
lepton number violating processes at the low scale is investigated. The main goal of this review
is to discuss the main implications of baryon and lepton number violation in physics beyond
the Standard Model.
Preprint submitted to Elsevier
January 9, 2015
Early Universe in the SU(3)L ⊗ U(1)X electroweak models
H. N. Long∗
Institute of Physics, Vietnam Academy of Science and Technology,
10 Dao Tan, Ba Dinh, Hanoi, Vietnam
arXiv:1501.01852v1 [hep-ph] 8 Jan 2015
(Dated: January 9, 2015)
We present status of the 3-3-1 models and their implications to cosmological evolution such as inflation, phase transitions and sphalerons. The models are able to
provide quite good agreement with the Standard Cosmology: the inflation happens
in the GUT scale, while phase transition has two sequences corresponding two steps
of symmetry breaking in the models. Some bounds on the model parameters are
obtained.
PACS numbers: 11.30.Fs,11.15.Ex,98.80.Cq
I.
INTRODUCTION
It is well known that our Universe content is 68.3% of Dark Energy (DE), 26.8% of Dark
Matter (DM) and of 4.9% of luminous matter [1]. With the unique fact of accelerated
Universe, the core origin of Dark Energy is still under question, while the existence of
Dark Matter is unambiguous. According to the Standard Cosmology, in the moment at
10−36 s after the Big Bang (BB), there was inflation, and our Universe has been expanded
exponentially. The inflationary scenario solves a number of problems such as the Universe’s
flatness, horizon, primordial monopole, etc. It is well known that there is no anti-matter
in our Universe, or other word speaking: at present there exists a Baryon Asymmetry of
Universe (BAU). The baryon number vanishes (nB = 0) at the BB, and this conflicts with
the present BAU. Nowadays, the BAU is one of the greatest challenges in Physics and
any physical model has to give an explanation. The BAU is realized if three Sakharov’s
conditions are satisfied [2, 3]
1. B violation,
∗
Electronic address: [email protected]
Unitarity in composite Higgs approaches with vector resonances
Daniele Barducci1 , Haiying Cai2 , Stefania De Curtis3 , Felipe J. Llanes-Estrada4 and Stefano Moretti5
1
arXiv:1501.01830v1 [hep-ph] 8 Jan 2015
5
LAPTH, Universit´e de Savoie, CNRS, B.P.110, F-74941 Annecy-le-Vieux, France.
2
Universit´e de Lyon, F-69622 Lyon, France, Universit´e Lyon 1,
CNRS/IN2P3, UMR5822 IPNL, F-69622 Villeurbanne Cedex, France.
3
INFN, Sezione di Firenze, Via G. Sansone 1, 50019 Sesto Fiorentino, Italy.
4
Departamento de F´ısica Te´
orica I, Univ. Complutense de Madrid, 28040 Madrid, Spain.
School of Physics and Astronomy, University of Southampton, Southampton SO17 1BJ, U.K.
(Dated: January 9, 2015)
We examine a simple Composite Higgs Model (CHM) with vector resonances in addition to the
Standard Model (SM) fields in perturbation theory by using the K-matrix method to implement
unitarity constraints. We find that the WL WL scattering amplitude has an additional scalar pole
(analogous to the σ meson of QCD) as in generic strongly interacting extensions of the SM. The
mass and width of this dynamically generated scalar resonance are large and the mass behaves
contrary to the vector one, so that when the vector resonance is lighter, the scalar one is heavier,
and vice versa. We also attempt an interpretation of this new resonance. Altogether, the presence
of the vector state with the symmetries of the CHM improve the low-energy unitarity behavior also
in the scalar-isoscalar channel.
LAPTH-003/15
I.
INTRODUCTION
The recent discovery of a Higgs-boson [1] has revived interest in the Electro-Weak Symmetry Breaking
(EWSB) sector of the Standard Model (SM) and beyond. If this Higgs boson is confirmed to have exactly
the couplings expected in the SM, a renormalizable theory of the EW interactions will be a closed chapter of
physics history. Nevertheless, for several reasons, the
particle physics community feels that there could be further new particles beyond the newly discovered Higgs boson. It is then interesting that its reported mass, about
125 GeV, is of the same magnitude as the EW gauge
bosons, MW ≃ 82 GeV and MZ ≃ 91 GeV, while no
new particles have been seen up to 600–700 GeV. Particularly stringent are the bounds on possible further W ′
or Z ′ vector bosons and other particles coupling to W W
and W Z pairs below about 1.5 TeV [2].
A natural scenario that theoretically fits this insight is
that of a Composite Higgs Model (CHM) in which the
Higgs state is a naturally light quasi-Nambu-Goldstone
Boson (qNGB) stemming, like the longitudinal components of the gauge bosons WL and ZL , from the spontaneous breaking of a higher energy symmetry [3].
While we do not really know what that symmetry
might be like, Occam’s razor dictates to examine first
those models with the minimum number of ingredients.
In the EWSB sector, this means the four Goldstone
bosons that seem to be the low-energy content of the
theory. A minimal such choice is the SO(5) → SO(4)
breaking, proposed in [4], that we spell out in section II.
Since our goal is to look forward to the TeV and multiTeV region where new vector resonances may hide, and
this is high-energy compared with the EW scale, we can
profit from the Equivalence Theorem (ET) [5] between
the longitudinal WL components and the π qNGB’s. The
Lagrangian density that controls their low-energy interactions is discussed in subsection II B.
We then dedicate section III to the extraction of the
scattering amplitudes among the low-energy particles
in Leading Order (LO) chiral perturbation theory, extended by new vector resonances, that would correspond
to the first accessible states (at the Large Hadron Collider
(LHC)) of the CHM considered here. The amplitudes
therefore include contact chiral interactions that are a
polynomial in s and Beyond the SM (BSM) gauge-like
interactions ρ-π-π entering through t- and u-channel vector exchanges, with ρ representing the accessible (spin-1
gauge) resonances. The polynomial terms imply strong
interactions in spite of the Higgs being light [6, 7]. We
find that the hh → hh scattering amplitude vanishes in
LO, thus we calculate the other three (elastic π i π j →
π k π l and inelastic π i π j → hh, π i π j → π 3 h) relevant
modes. Actually, we will prove that the latter scattering amplitude vanishes due to a cancellation between the
contributions from the two degenerate vector resonances
contributing to the process.
The amplitudes are projected over the few lowest partial waves in section IV, where we check the good convergence of the expansion at low-energy. While the vector
channel is well behaved due to the new spin-1 resonances
introduced in the CHM scenario, this is not the case for
the scalar-isoscalar partial wave: we note the breakdown
of unitarity by perturbation theory in the 2 TeV region
for values of the parameters that are still compatible with
current LHC bounds. It is well known, and continues being reinstated [8], that, generically, if the couplings of the
Higgs boson do not perfectly match the SM ones, unitarity violations are expected (see [9] for an exception).
A traditional way out is to restrict the analysis to
those values of the parameters f , gs , the ‘compositeness’
(energy) scale and the new gauge coupling, respectively
The properties of C-parameter and coupling constants
arXiv:1501.01761v1 [hep-ph] 8 Jan 2015
R. Saleh-Moghaddam, M. E. Zomorrodian
Department of Physics, Faculty of sciences, Ferdowsi university of Mashhad,
91775-1436, Mashhad, Iran
In this article, we present the properties of the C-parameter which is
one of event shape variables. We obtain the coupling constants both in
the perturbative and in the non-perturbative part of the QCD theory. To
achieve this we fit the dispersive model as well as the shape function model
with our data. Our results are consistent with the QCD predictions. We
explain more features of our results in the main text.
PACS numbers: 12.38.-t, 12.38.Bx, 12.38.Lg
1. Introduction
Event shape variables in e+ e− annihilation provide an ideal testing
ground to study Quantum Chromo-Dynamics (QCD) and have been measured and studied extensively in the last three decades. In particular, event
shape variables are interesting for studying the interplay between perturbative and non-perturbative dynamics [1]. One of the most common and
successful ways of testing QCD has been by investigating the distribution
of event shapes in e+ e− → hadrons, which have been measured accurately
over a range of centre-of-mass energies, and provide a useful way of evaluating the strong coupling constant αs . The main obstruction to obtaining
an accurate value of αs from distributions is not due to a lack of precise
data but to dominant errors in the theoretical calculation of the distributions. In particular, there are non-perturbative effects that cannot yet be
calculated from first principles but cause power-suppressed corrections that
can be signicant at experimentally accessible energy scales [2].
In this article, we show the cross section for C- parameter as an event
shape observable. Also we peruse both perturbative and non-perturbative
theory for calculating coupling constants using different models .It is mentioned that we have already done some analyses on some event shape observables in our previous publications [3, 4].
(1)
arXiv:???
charged Higgs pair production at the LHC as a probe of the
top-seesaw assisted technicolor models
Guo-Li Liu1∗ , Xiao-Fei Guo1 , Kun Wu1 , Ji Jiang1 , Ping Zhou1,2
1
Department of Physics, Zhengzhou University, Zhengzhou, 450001, China
arXiv:1501.01714v1 [hep-ph] 8 Jan 2015
2
National Space Science Center, Chinese Academy of Science,100190
Abstract
The top-seesaw assisted technicolor model, which was proposed recently to explain the 126
GeV Higgs mass discovered by the Large Hadron Colliders (LHC), can predicts some light and
heavy charged Higgs bosons in addition to the neutral Higgs. In this work we will study the
pair productions of the charged Higgs, light or heavy ones, at the LHC. For the productions at
the LHC we consider the processes proceeding through gluon-gluon fusion and quark-anti-quark
annihilation. We find that in a large part of parameter space the production cross sections of
the light charged Higgs at the LHC can be quite large compared with the low standard model
backgrounds, while the suppression of the heavy ones is just too strong to be detected. Therefore,
at the LHC future experiments, the light charged Higgs pair productions may be detectable, which
allows for probing this new technicolor models.
PACS numbers: 12.60.Nz, 14.80.Bn
∗
[email protected]
1