ArticlePDF Available

Antiproton Flux, Antiproton-to-Proton Flux Ratio, and Properties of Elementary Particle Fluxes in Primary Cosmic Rays Measured with the Alpha Magnetic Spectrometer on the International Space Station

Authors:

Abstract

A precision measurement by AMS of the antiproton flux and the antiproton-to-proton flux ratio in primary cosmic rays in the absolute rigidity range from 1 to 450 GV is presented based on 3.49 × 10^5 antiproton events and 2.42 × 10^9 proton events. The fluxes and flux ratios of charged elementary particles in cosmic rays are also presented. In the absolute rigidity range ∼60 to ∼500 GV, the antiproton p-, proton p, and positron e+ fluxes are found to have nearly identical rigidity dependence and the electron e− flux exhibits a different rigidity dependence. Below 60 GV, the (p-/p), (p-/e+ ), and (p/e+ ) flux ratios each reaches a maximum. From ∼60 to ∼500 GV, the (p-/p), (p-/e+ ), and (p/e+ ) flux ratios show no rigidity dependence. These are new observations of the properties of elementary particles in the cosmos.
Antiproton Flux, Antiproton-to-Proton Flux Ratio, and Properties of Elementary Particle
Fluxes in Primary Cosmic Rays Measured with the Alpha Magnetic Spectrometer
on the International Space Station
M. Aguilar,26 L. Ali Cavasonza,1B. Alpat,32 G. Ambrosi,32 L. Arruda,24 N. Attig,21 S. Aupetit,17 P. Azzarello,16
A. Bachlechner,1F. Barao,24 A. Barrau,17 L. Barrin,15 A. Bartoloni,38 L. Basara,36 S. Başegˇmez-du Pree,6M. Battarbee,46
R. Battiston,36,37,a J. Bazo,32,b U. Becker,9M. Behlmann,9B. Beischer,1J. Berdugo,26 B. Bertucci,32,33 V. Bindi,19
G. Boella,28,29 W. de Boer,22 K. Bollweg,20 V. Bonnivard,17 B. Borgia,38,39 M. J. Boschini,28 M. Bourquin,16 E. F. Bueno,40
J. Burger,9F. Cadoux,16 X. D. Cai,9M. Capell,9S. Caroff,3J. Casaus,26 G. Castellini,14 I. Cernuda,26 F. Cervelli,34
M. J. Chae,41 Y. H. Chang,10 A. I. Chen,9G. M. Chen,6H. S. Chen,6L. Cheng,42 H. Y. Chou,10 E. Choumilov,9V. Choutko,9
C. H. Chung,1C. Clark,20 R. Clavero,23 G. Coignet,3C. Consolandi,19 A. Contin,7,8 C. Corti,19 B. Coste,36 W. Creus,10
M. Crispoltoni,32,33,15 Z. Cui,42 Y. M. Dai,5C. Delgado,26 S. Della Torre,28 M. B. Demirköz,2L. Derome,17 S. Di Falco,34
F. Dimiccoli,36,37 C. Díaz,26 P. von Doetinchem,19 F. Dong,30 F. Donnini,32,33 M. Duranti,32,33 D. DUrso,32,c A. Egorov,9
A. Eline,9T. Eronen,46 J. Feng,45,d E. Fiandrini,32,33 E. Finch,31 P. Fisher,9V. Formato,32,15 Y. Galaktionov,9G. Gallucci,34
B. García,26 R. J. García-López,23 C. Gargiulo,15 H. Gast,1I. Gebauer,22 M. Gervasi,28,29 A. Ghelfi,17 F. Giovacchini,26
P. Goglov,9D. M. Gómez-Coral,27 J. Gong,30 C. Goy,3V. Grabski,27 D. Grandi,28 M. Graziani,32,33 I. Guerri,34,35
K. H. Guo,18 M. Habiby,16 S. Haino,45 K. C. Han,25 Z. H. He,18 M. Heil,9J. Hoffman,19 T. H. Hsieh,9H. Huang,45,e
Z. C. Huang,18 C. Huh,13 M. Incagli,34 M. Ionica,32 W. Y. Jang,13 H. Jinchi,25 S. C. Kang,13 K. Kanishev,36,37 G. N. Kim,13
K. S. Kim,13 Th. Kirn,1C. Konak,2O. Kounina,9A. Kounine,9V. Koutsenko,9M. S. Krafczyk,9G. La Vacca,28 E. Laudi,15
G. Laurenti,7I. Lazzizzera,36,37 A. Lebedev,9H. T. Lee,44 S. C. Lee,45 C. Leluc,16 H. S. Li,43 J. Q. Li,9,f J. Q. Li,30 Q. Li,30
T. X. Li,18 W. Li,4Z. H. Li,6Z. Y. Li,45,d S. Lim,13 C. H. Lin,45 P. Lipari,38 T. Lippert,21 D. Liu,45 Hu Liu,26,g S. Q. Lu,45,d
Y. S. Lu,6K. Luebelsmeyer,1F. Luo,42 J. Z. Luo,30 S. S. Lv,18 R. Majka,31 C. Mañá,26 J. Marín,26 T. Martin,20 G. Martínez,26
N. Masi,7D. Maurin,17 A. Menchaca-Rocha,27 Q. Meng,30 D. C. Mo,18 L. Morescalchi,34,h P. Mott,20 T. Nelson,19 J. Q. Ni,18
N. Nikonov,1F. Nozzoli,32,c P. Nunes,24 A. Oliva,26 M. Orcinha,24 F. Palmonari,7,8 C. Palomares,26 M. Paniccia,16
M. Pauluzzi,32,33 S. Pensotti,28,29 R. Pereira,19 N. Picot-Clemente,12 F. Pilo,34 C. Pizzolotto,32,c V. Plyaskin,9M. Pohl,16
V. Poireau,3A. Putze,3,i L. Quadrani,7,8 X. M. Qi,18 X. Qin,32,j Z. Y. Qu,45,k T. Räihä,1P. G. Rancoita,28 D. Rapin,16
J. S. Ricol,17 I. Rodríguez,26 S. Rosier-Lees,3A. Rozhkov,9D. Rozza,28 R. Sagdeev,11 J. Sandweiss,31 P. Saouter,16
S. Schael,1S. M. Schmidt,21 A. Schulz von Dratzig,1G. Schwering,1E. S. Seo,12 B. S. Shan,4J. Y. Shi,30 T. Siedenburg,1
D. Son,13 J. W. Song,42 W. H. Sun,9,l M. Tacconi,28 X. W. Tang,6Z. C. Tang,6L. Tao,3D. Tescaro,23 Samuel C. C. Ting,9,15
S. M. Ting,9N. Tomassetti,17 J. Torsti,46 C. Türkoğlu,2T. Urban,20 V. Vagelli,32 E. Valente,38,39 C. Vannini,34 E. Valtonen,46
M. Vázquez Acosta,23 M. Vecchi,40 M. Velasco,26 J. P. Vialle,3V. Vitale,32,c S. Vitillo,16 L. Q. Wang,42 N. H. Wang,42
Q. L. Wang,5X. Wang,9X. Q. Wang,6Z. X. Wang,18 C. C. Wei,45,m Z. L. Weng,9K. Whitman,19 J. Wienkenhöver,1
M. Willenbrock,9H. Wu,30 X. Wu,16 X. Xia,26,j R. Q. Xiong,30 W. Xu,9Q. Yan,9J. Yang,41 M. Yang,6Y. Yang,43 H. Yi,30
Y. J. Yu,5Z. Q. Yu,6S. Zeissler,22 C. Zhang,6J. Zhang,9,e J. H. Zhang,30 S. D. Zhang,9,f S. W. Zhang,6Z. Zhang,9
Z. M. Zheng,4Z. Q. Zhu,9,n H. L. Zhuang,6V. Zhukov,1A. Zichichi,7,8 N. Zimmermann,1and P. Zuccon9
(AMS Collaboration)
1I. Physics Institute and JARA-FAME, RWTH Aachen University, D52056 Aachen, Germany
2Department of Physics, Middle East Technical University (METU), 06800 Ankara, Turkey
3Laboratoire dAnnecyleVieux de Physique des Particules (LAPP), CNRS/IN2P3 and Université Savoie Mont Blanc,
F74941 AnnecyleVieux, France
4Beihang University (BUAA), Beijing, 100191, China
5Institute of Electrical Engineering (IEE), Chinese Academy of Sciences, Beijing, 100190, China
6Institute of High Energy Physics (IHEP), Chinese Academy of Sciences, Beijing, 100049, China
7INFN Sezione di Bologna, I40126 Bologna, Italy
8Università di Bologna, I40126 Bologna, Italy
9Massachusetts Institute of Technology (MIT), Cambridge, Massachusetts 02139, USA
10National Central University (NCU), ChungLi, Tao Yuan, 32054, Taiwan
11EastWest Center for Space Science, University of Maryland, College Park, Maryland 20742, USA
12IPST, University of Maryland, College Park, Maryland 20742, USA
13CHEP, Kyungpook National University, 41566 Daegu, Korea
PRL 117, 091103 (2016) PHYSICAL REVIEW LETTERS week ending
26 AUGUST 2016
0031-9007=16=117(9)=091103(10) 091103-1 Published by the American Physical Society
14CNRIROE, I-50125 Firenze, Italy
15European Organization for Nuclear Research (CERN), CH1211 Geneva 23, Switzerland
16DPNC, Université de Genève, CH1211 Genève 4, Switzerland
17Laboratoire de Physique Subatomique et de Cosmologie (LPSC), CNRS/IN2P3 and Université GrenobleAlpes,
F38026 Grenoble, France
18Sun YatSen University (SYSU), Guangzhou, 510275, China
19Physics and Astronomy Department, University of Hawaii, Honolulu, Hawaii 96822, USA
20National Aeronautics and Space Administration Johnson Space Center (JSC), Jacobs Engineering,
and Business Integra, Houston, Texas 77058, USA
21Jülich Supercomputing Centre and JARA-FAME, Research Centre Jülich, D52425 Jülich, Germany
22Institut für Experimentelle Kernphysik, Karlsruhe Institute of Technology (KIT), D76128 Karlsruhe, Germany
23Instituto de Astrofísica de Canarias (IAC), E38205 La Laguna, and Departamento de Astrofísica, Universidad de La Laguna,
E38206 La Laguna, Tenerife, Spain
24Laboratório de Instrumentação e Física Experimental de Partículas (LIP), P1000 Lisboa, Portugal
25National ChungShan Institute of Science and Technology (NCSIST), Longtan, Tao Yuan, 32546, Taiwan
26Centro de Investigaciones Energéticas, Medioambientales y Tecnológicas (CIEMAT), E28040 Madrid, Spain
27Instituto de Física, Universidad Nacional Autónoma de México (UNAM), México, D. F., 01000 Mexico
28INFN Sezione di MilanoBicocca, I20126 Milano, Italy
29Università di MilanoBicocca, I20126 Milano, Italy
30Southeast University (SEU), Nanjing, 210096, China
31Physics Department, Yale University, New Haven, Connecticut 06520, USA
32INFN Sezione di Perugia, I06100 Perugia, Italy
33Università di Perugia, I06100 Perugia, Italy
34INFN Sezione di Pisa, I56100 Pisa, Italy
35Università di Pisa, I56100 Pisa, Italy
36INFN TIFPA, I38123 Povo, Trento, Italy
37Università di Trento, I38123 Povo, Trento, Italy
38INFN Sezione di Roma 1, I00185 Roma, Italy
39Università di Roma La Sapienza, I00185 Roma, Italy
40Instituto de Fìsica de São Carlos, Universidade de São Paulo, CP 369, 13560-970, São Carlos, São Paulo, Brazil
41Department of Physics, Ewha Womans University, Seoul, 120-750, Korea
42Shandong University (SDU), Jinan, Shandong, 250100, China
43National Cheng Kung University, Tainan, 70101, Taiwan
44Academia Sinica Grid Center (ASGC), Nankang, Taipei 11529, Taiwan
45Institute of Physics, Academia Sinica, Nankang, Taipei 11529, Taiwan
46Space Research Laboratory, Department of Physics and Astronomy, University of Turku, FI20014 Turku, Finland
(Received 6 May 2016; revised manuscript received 14 June 2016; published 26 August 2016)
A precision measurement by AMS of the antiproton flux and the antiproton-to-proton flux ratio in
primary cosmic rays in the absolute rigidity range from 1 to 450 GV is presented based on 3.49 ×105
antiproton events and 2.42 ×109proton events. The fluxes and flux ratios of charged elementary particles
in cosmic rays are also presented. In the absolute rigidity range 60 to 500 GV, the antiproton ¯
p, proton
p, and positron eþfluxes are found to have nearly identical rigidity dependence and the electron eflux
exhibits a different rigidity dependence. Below 60 GV, the ( ¯
p=p), ( ¯
p=eþ), and (p=eþ) flux ratios each
reaches a maximum. From 60 to 500 GV, the ( ¯
p=p), ( ¯
p=eþ), and (p=eþ) flux ratios show no rigidity
dependence. These are new observations of the properties of elementary particles in the cosmos.
DOI: 10.1103/PhysRevLett.117.091103
We report on the measurement of the antiproton flux and
of the antiproton-to-proton flux ratio in primary cosmic
rays in the absolute rigidity range from 1 to 450 GV based
on 3.49 ×105antiproton events and 2.42 ×109proton
events collected by the Alpha Magnetic Spectrometer,
AMS, on the International Space Station, ISS, from May
19, 2011 to May 26, 2015. Of the four charged elementary
particles traveling through the cosmosprotons, electrons,
positrons, and antiprotonsthe experimental data on anti-
protons are limited because for each antiproton there are
approximately 104protons. Since the observation of anti-
protons in cosmic rays [1], many studies of cosmic ray
antiprotons have been performed [26].However,to
measure the antiproton flux to 1% accuracy requires a
Published by the American Physical Society under the terms of
the Creative Commons Attribution 3.0 License. Further distri-
bution of this work must maintain attribution to the author(s) and
the published articles title, journal citation, and DOI.
PRL 117, 091103 (2016) PHYSICAL REVIEW LETTERS week ending
26 AUGUST 2016
091103-2
separation power of 106. The sensitivity of antiprotons to
cosmic phenomena [710] is complementary to the sensi-
tivity of the measurements of cosmic ray positrons. For
example, AMS has accurately measured the excess in the
positron fraction to 500 GeV [11]. This data generated
many interesting theoretical models including collisions of
dark matter particles [12], astrophysical sources [13], and
collisions of cosmic rays [14]. Some of these models also
include specific predictions on the antiproton flux and the
antiproton-to-proton flux ratio in cosmic rays.
Simultaneously with these ¯
pmeasurements, the electron
eand positron eþfluxes [15] and the proton pflux [16]
have been analyzed by AMS. Hence, in this Letter we also
report the accurate study of the rigidity dependence of
elementary particle fluxes and their ratios in primary
cosmic rays. These measurements, performed with the
same detector, provide precise experimental information
over an extended energy range in the study of elementary
particles traveling through the cosmos.
Detector.The description of the AMS detector is
presented in Refs. [17,18]. All detector elements are used
for particle identification in the present analysis: the silicon
tracker [19], the permanent magnet [5,20], the time of flight
counters TOF [21], the anticoincidence counters ACC [22],
the transition radiation detector TRD [23], the ring imaging
Čerenkov detector RICH [24], and the electromagnetic
calorimeter ECAL [25].
To measure the rigidity R(momentum per unit of charge)
of cosmic rays and to differentiate between positive and
negative particles, the tracker has nine layers. The first (L1)
is at the top of the detector, the second (L2) just above the
magnet, six (L3 to L8) within the bore of the magnet, and
the last (L9) just above the ECAL. L2 to L8 constitute the
inner tracker. For jZ1particles the maximum detectable
rigidity, MDR, is 2 TV and the charge resolution is
ΔZ¼0.05. The TOF measures jZjand velocity with a
resolution of Δβ=β2¼4%. The ACC has 0.99999 effi-
ciency to reject cosmic rays entering the inner tracker from
the side.
The TRD separates ¯
pand pfrom eand eþusing the
ΛTRD estimator constructed from the ratio of the log-like-
lihood probability of the ehypothesis to that of the ¯
por p
hypothesis in each layer [11]. Antiprotons and protons,
which have ΛTRD 1, are efficiently separated from elec-
trons and positrons, which have ΛTRD 0.5. The RICH has a
velocity resolution Δβ=β¼0.1% for jZ1to ensure
separation of ¯
pand pfrom light particles (eand π)
below 10 GV. The ECAL is used to separate ¯
pand pfrom e
and eþwhen the event can be measured by the ECAL.
To distinguish antiprotons from charge confusion pro-
tons, that is, protons which are reconstructed in the tracker
with negative rigidity due to the finite tracker resolution or
due to interactions with the detector materials, a charge
confusion estimator ΛCC is defined using the boosted
decision tree technique [26]. The estimator combines
information from the tracker such as the track χ2=d:o:f:,
rigidities reconstructed with different combinations of
tracker layers, the number of hits in the vicinity of the
track, and the charge measurements in the TOF and the
tracker. With this method, antiprotons, which have
ΛCC þ1, are efficiently separated from charge confusion
protons, which have ΛCC ∼−1.
Event selection and data samples.Over 65 billion
cosmic ray events have been recorded in the first 48 months
of AMS operations. Only events collected during normal
detector operating conditions are used in this analysis. This
includes the time periods when the AMS zaxis is pointing
within 40° of the local zenith and when the ISS is not in the
South Atlantic Anomaly. Data analysis is performed in 57
absolute rigidity bins. The same binning as in our proton
flux measurement [16] is chosen below 80.5 GV. Above
80.5 GV two to four bins from Ref. [16] are combined to
ensure sufficient antiproton statistics.
Events are selected requiring a track in the TRD and in
the inner tracker and a measured velocity β>0.3in the
TOF corresponding to a downward-going particle. To
maximize the number of selected events while maintaining
an accurate rigidity measurement, the acceptance is
increased by releasing the requirements on the external
tracker layers, L1 and L9. Below 38.9 GV neither L1 nor
L9 is required. From 38.9 to 147 GV either L1 or L9 is
required. From 147 to 175 GV only L9 is required. Above
175 GV both L1 and L9 are required. In order to maximize
the accuracy of the track reconstruction, the χ2=d:o:f:of the
reconstructed track fit is required to be less than 10 both in
the bending and nonbending projections. The dE=dx
measurements in the TRD, the TOF, and the inner tracker
must be consistent with jZ1. To select only primary
cosmic rays, the measured rigidity is required to exceed the
maximum geomagnetic cutoff by a factor of 1.2 for either
positive or negative particles within the AMS field of view.
The cutoff is calculated by backtracing [16,27] using the
most recent IGRF geomagnetic model [28].
Events satisfying the selection criteria are classified into
two categoriespositive and negative rigidity events. A
total of 2.42 ×109events with positive rigidity are selected
as protons. They are 99.9% pure protons with almost no
background. Deuterons are not distinguished from protons,
their contribution decreases with rigidity: at 1 GV it is less
than 2% and at 20 GV it is 0.6% [5,29]. The effective
acceptance of this selection for protons is larger than in our
proton flux publication [16]. This is because there is no
strict requirement that selected particles pass through the
tracker layers L1 and L9 (see above) leading to a much
larger field of view at low rigidities and, therefore, to a
significant increase in the number of protons.
The negative rigidity event category comprises both
antiprotons and several background sources: electrons,
light negative mesons (πand a negligible amount of
K) produced in the interactions of primary cosmic rays
PRL 117, 091103 (2016) PHYSICAL REVIEW LETTERS week ending
26 AUGUST 2016
091103-3
with the detector materials, and charge confusion protons.
The contributions of the different background sources vary
with rigidity. For example, light negative mesons are
present only at rigidities below 10 GV, whereas charge
confusion becomes noticeable only at high rigidities.
Electron background is present at all rigidities. The
combination of information from the TRD, TOF, tracker,
RICH, and ECAL enables the efficient separation of the
antiproton signal events from these background sources
using a template fitting technique. The number of observed
antiproton signal events and its statistical error in the
negative rigidity sample are determined in each bin by
fitting signal and background templates to data by varying
their normalization. As discussed below, the template
variables used in the fit are constructed using information
from the TOF, tracker, and TRD. The distribution of the
variables for the template definition is the same for
antiprotons and protons if they are both reconstructed with
a correct charge-sign. This similarity has been verified with
the Monte Carlo simulation [30] and the antiproton and
proton data of 2.97 jRj<18.0GV. Therefore, the signal
template is always defined using the high-statistics proton
data sample. Three overlapping rigidity regions with differ-
ent types of template function are defined to maximize the
accuracy of the analysis: low absolute rigidity region (1.00
4.02 GV), intermediate region (2.9718.0 GV), and high
absolute rigidity region (16.6450 GV). In the overlapping
rigidity bins, the results with the smallest error are selected.
At low rigidities, a cut on the TRD estimator ΛTRD and
the velocity measurement in the TOF are important to
differentiate antiprotons from light particles (eand π).
Therefore, the mass distribution, calculated from the
rigidity measurement in the inner tracker and the velocity
measured by the TOF, is used to construct the templates and
to differentiate between the antiproton signal and the
background. The background eand πtemplates are
defined from the data sample selected using information
from the TRD, the RICH, and also the ECAL, when the
event can be measured by the ECAL.
At intermediate rigidities, ΛTRD and the velocity mea-
sured with the RICH βRICH are used to separate the
antiproton signal from light particles (eand π). As an
example, Fig. 1(a) shows that the antiproton signal and the
background are well separated in the (βRICH ΛTRD ) plane
for the absolute rigidity range 5.46.5 GV. To determine the
number of antiproton signal events, the πbackground is
removed by a rigidity dependent βRICH cut and the ΛTRD
distribution is used to construct the templates and to
differentiate between the ¯
psignal and ebackground.
The background template is defined from the edata
sample selected using ECAL. The Monte Carlo simulation
matches the data for eevents inside the ECAL. The
Monte Carlo simulation was then used to verify that the e
template shape outside the ECAL and inside the ECAL are
identical.
In the high rigidity region, the two-dimensional
(ΛTRD ΛCC) distribution is used to determine the number
of antiproton signal events. The lower bound of ΛCC is
chosen for each bin to optimize the accuracy of the fit. For
example, for jRj>175 GV, ΛCC ≥−0.6. Variation of the
lower bound is included in the systematic errors discussed
below. To fit the data three template shapes are defined. The
first two are for antiprotons and electrons with correctly
reconstructed charge sign and the last one is for charge
confusion protons. The background templates (i.e., elec-
trons and charge confusion protons) are from the
Monte Carlo simulation. The Monte Carlo simulation of
the change confusion was verified with the 400 GV proton
test beam data. An example of the fit to the data is shown in
Figs. 1(b) and 1(c) for the rigidity bin 175211 GV. The
distribution of data in the (ΛTRD ΛCC) plane is shown in
Fig. 1(b) and the fit results showing the signal and back-
ground distributions is highlighted in Fig. 1(c). The χ2of
the fit is 138 for 154 degrees of freedom.
Overall, results for all 57 rigidity bins give a total of
3.49 ×105antiproton events in the data.
Analysis.The isotropic antiproton flux for the absolute
rigidity bin Riof width ΔRiis given by
Λ ×sign(R)
-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5
RICH
β
0.96
0.98
1.0
1
10
2
10
3
10
4
10
5
10
pp
-
e
+
π
+
e
-
π
TRD
Λ
TRD
Λ
CC
Λ
CC
Λ
TRD
0.500.5 1
0.5
1
Events
0
5
10
15
Fit p
0.5
00.5
0.5
1
Events
0
10
15
Data
(a)
(b) (c)
5
-
e
p
Events
0
10
15
5
FIG. 1. (a) Negative rigidity and positive rigidity data samples
in the [βRICH signðRÞ×ΛTRD] plane for the absolute rigidity
range 5.46.5 GV. The contributions of ¯
p,p,eþ,e,πþ, and π
are clearly seen. The antiproton signal is well separated from the
backgrounds. (b) For negative rigidity events, the distribution of
data events in the (ΛTRD ΛCC) plane for the absolute rigidity bin
175211 GV. (c) Fit with χ2=d:o:f:¼138=154 of the antiproton
signal template (magenta), the electron background template
(blue), and the charge confusion proton background template
(green) to the data in (b).
PRL 117, 091103 (2016) PHYSICAL REVIEW LETTERS week ending
26 AUGUST 2016
091103-4
Φ¯
p
i¼N¯
p
i
A¯p
iTiΔRi
;ð1Þ
where the rigidity is defined on top of the AMS, N¯
p
iis the
number of antiprotons in the rigidity bin icorrected with
the rigidity resolution function (see below), A¯p
iis the
corresponding effective acceptance that includes geometric
acceptance as well as the trigger and selection efficiency,
and Tiis the exposure time.
Detector resolution effects cause migration of events N¯p
i
from rigidity bin Rito the measured rigidity bins ~
Rj
resulting in the observed number of events ~
N¯p
j.To
account for this event migration, an iterative unfolding
procedure [31] is used to correct the number of observed
events. It is described in detail in our publication on
the proton flux [16]. At each iteration, the folded accep-
tance is defined using the Monte Carlo simulation:
~
A¯
p
i¼ð1=Φ¯
p
iÞPjΦ¯
p
jA¯
p
jM¯
p
ij, where M¯
p
ij is the migration
probability from bin jto bin i. Parametrizing ~
A¯
pwith a
spline function, the number of events is corrected bin by bin
by a factor A¯
p=~
A¯pand the flux is recalculated according to
Eq. (1). The iteration proceeds until the fluxes calculated
for two consecutive steps agree within 0.1% and the
measured flux can be expressed as
Φ¯
p
i¼
~
N¯
p
i
~
A¯
p
iTiΔRi
:ð2Þ
The same procedure is used to unfold the observed number
of 2.42 ×109proton events in this analysis.
The ( ¯
p=p) flux ratio is defined for each absolute rigidity
bin by
¯
p
pi
Φ¯p
i
Φp
i
¼
~
N¯p
i
~
Np
i
~
Ap
i
~
A¯p
i
:ð3Þ
With 3.49 ×105antiproton events, the detailed study of
systematic errors of the ¯
pflux and ( ¯
p=p) flux ratio is the
key part of the present analysis.
There are four sources of systematic errors on the ¯
pflux
and ( ¯
p=p) flux ratio. The first source affects mostly ~
N¯
p
i
and, to a much lesser extent, ~
Np
i. It includes uncertainties in
the definition of the geomagnetic cutoff factor, in the event
selection, and in the shape of the templates. The second
source affects ~
A¯
p
iand ~
Ap
i. It includes uncertainties in the
inelastic cross sections of protons and antiprotons in the
detector materials and in the migration matrices M¯
p
ij and
Mp
ij. The third source is the uncertainty in the absolute
rigidity scale. The fourth source, relevant only for the ¯
p
flux, is the uncertainty in the normalization of the effective
folded acceptance ~
A¯
p
i. Contributions of these four sources
to the systematic errors on the ¯
pflux and ( ¯
p=p) flux ratio
are discussed below. They are added quadratically to arrive
at the systematic errors.
Variation of the geomagnetic cutoff factor in the range
1.2 to 1.4 shows a systematic uncertainty of 1% at 1 GV
and negligible above 2 GV for both ~
N¯
p
iand ~
Np
i. To evaluate
systematic uncertainties related to the event selection, the
analysis is repeated in each rigidity bin 1000 times with
different sets of cut values, such that the selection efficiency
varies up to 10% (see Ref. [11] for details). This uncertainty
in ~
N¯
p
iamounts to 4% at 1 GV, 0.5% at 10 GV, and rises to
6% at 450 GV. The uncertainty in ~
Np
iis negligible over the
entire rigidity range.
Uncertainty in the shape of the fit templates affects only
~
N¯
p
i. It becomes particularly important at high rigidities
(>150 GV) where charge confusion protons enter the
antiproton region. Three template shapes are used for
the fit in this regionthe antiproton signal template, the
electron background template, and the charge confusion
proton template. The antiproton signal template has the same
shape as protons reconstructed with correct charge-sign (see
above). It is extracted from high statistics proton data,
therefore the systematic effects are negligible. The electron
template from the Monte Carlo simulation is validated with
electron data and does not contribute to the systematic error.
The systematic error due to the uncertainty in the shape of
the charge confusion proton template originates from the
uncertainties of the proton flux in the TV region and from
the uncertainties of the proton rigidity resolution function.
The former is estimated by varying the spectral index of the
proton flux within the accuracy of our proton measurement
[16]. The later is estimated by comparing the charge
confusion amount predicted by the Monte Carlo simulation
with the one obtained from the fit of the three template
shapes to the negative rigidity data. Overall, the systematic
error from the templates is estimated to be 12% at 450 GV,
decreasing to <1% below 30 GV.
The systematic errors on the folded acceptances ~
A¯
pand
~
Aporiginate from the uncertainties in the interaction cross
sections for protons and antiprotons in the detector materi-
als [32,33] and the uncertainties in the migration matrices
M¯
p
ij and Mp
ij. The systematic error from the cross section
uncertainties is estimated by varying the ¯
pand pinter-
action cross sections in the Monte Carlo simulation within
the accuracy of the cross section measurements [32,33].
The corresponding systematic error on ~
A¯pis found to be
4% at 1 GV and 1% above 50 GV. The error on ~
Apis
found to be 2.5% at 1 GV and 1% above 50 GV. These
values are larger than those in Ref. [16] due to the larger
acceptance. The systematic errors on ~
A¯
pand ~
Apdue to the
cross section uncertainties are independent and they are
added in quadrature to get the systematic error on the
~
Ap=~
A¯pratio. The systematic uncertainty in the migration
matrix Mp
ij is studied by comparing the test beam data at
PRL 117, 091103 (2016) PHYSICAL REVIEW LETTERS week ending
26 AUGUST 2016
091103-5
400 GV with the Monte Carlo simulation. The rigidity
resolution function Δð1=RÞhas a pronounced Gaussian
core of width σand non-Gaussian tails spanning beyond
2.5σfrom the center [16]. Uncertainties in the core and in
the tails of the antiproton migration matrix M¯
p
ij are assumed
to be the same as for Mp
ij. Varying both the width of the core
and the amount of non-Gaussian tails as described in
Ref. [16] yields the systematic error of 1% below
200 GV and 1.5% at 450 GV for both ~
A¯
pand ~
Ap.
These systematic errors partially cancel in the ~
Ap=~
A¯
pratio,
yielding uncertainties of 1% at 1 GV and <0.5% above
2 GV. We note that the ~
Ap=~
A¯
pratio decreases from 1.15 at
1 GV to 1.04 at 450 GV due to the varying difference of
interaction cross sections for protons and antiprotons, and
due to bin-to-bin event migration.
The error on the absolute rigidity scale due to the residual
misalignment of the tracker planes was estimated by
comparing the electron and positron energies measured
in the ECAL with the momentum measured in the tracker to
be 1=26 TV1, see Ref. [16] for details. The corresponding
errors on the antiproton and proton fluxes are negligble
below 10 GV and gradually increase to 1% at 450 GV.
This error has opposite effects on the measured proton and
antiproton fluxes, therefore the error on the ( ¯
p=p) flux ratio
gradually increases to 2% at 450 GV.
The systematic error on the normalization of the effective
folded acceptance is due to small differences between the
data and the Monte Carlo samples [16]. This error is
relevant only for the ¯
pflux; it cancels in the ( ¯
p=p) flux
ratio. It is estimated to be 5% at 1 GV decreasing gradually
to 2% above 20 GV.
As stated above, charge confusion protons are protons
which are reconstructed in the tracker with negative
rigidity due to the finite tracker resolution or due to
interactions with the detector materials. The uncertainty
from the charge confusion proton template becomes sig-
nificant for jRj>30 GV. To ensure that the shape of
the charge confusion proton background templates from the
Monte Carlo simulation does not introduce bias into the
antiproton identification, we also performed a completely
independent data driven analysis based on the linear
regression method [34] for jRj>30 GV over the accep-
tance which includes L1, L9, and ECAL. In this analysis a
weighted sum is constructed from a set of measured
quantities sensitive to the finite tracker resolution and
interactions with the detector materials. This set includes
information from the TRD, TOF, tracker, ACC, and ECAL.
In particular, the tracker information includes the χ2=d:o:f:,
the rigidity reconstructed using different combinations of
tracker layers, and the maximum distance of the tracker hits
from the fitted track. The weights of individual quantities are
optimized to maximize the separation between the antiproton
signal and the charge confusion background. The resultant
sums are nearly Gaussian for both the antiproton signal and
the charge confusion proton background. This allows con-
struction of the template (i.e., a Gaussian distribution with
1% non-Gaussian tails) without relying on the Monte Carlo
simulation. The non-Gaussian tails are evaluated using
iterative fits with the subtraction of both the antiproton
and charge confusion proton templates from the negative
rigidity data. The details of the analysis used in this
technique will be presented in a separate publication. The
results of this data driven analysis agree within the system-
atic errors with those presented in this Letter.
Most importantly, in addition to the linear regression
analysis, several other independent analyses were per-
formed on the same data sample by different study groups.
The results of those analyses are consistent with this Letter.
Results.The measured antiproton flux and antiproton-to-
proton flux ratio together with their statistical and systematic
errors are presented in Table I of Supplemental Material [18] as
a function of the absolute rigidity value at the top of the AMS
detector along with the number of observed antiproton events
~
N¯
p. The statistical errors are from the fit errors on the signal.
As seen from this table, the statistical and systematic error
contributions to the total error in the flux and flux ratio vary
with rigidity. For 1.00 jRj<1.33 GV the statistical error
dominates, for 1.33 jRj<1.71 GV the errors are compa-
rable, for 1.71 jRj<48.5GV the systematic error domi-
nates, for 48.5jRj<108 GV the errors are comparable,
and for 108 jRj<450 GV statistical error dominates.
AMS has now measured the fluxes of electrons and
positrons [15] and of protons [16] with the same detector
and same time period, May 19, 2011 to November 26,
2013. The antiprotons in this Letter were measured from
May 19, 2011 to May 26, 2015. We have studied the time
dependent solar effects on these fluxes during the interval
covered by this Letter [35]. Within our current accuracy, the
time dependent solar effects are observable for antiprotons,
electrons, and positrons with jRj<10 GV and for protons
with jRj<20 GV. For the study of the flux ratio depend-
ence of elementary particles (see below), we chose jRj>
10 GV where the time dependent solar effects for protons
are small and the uncertainties are dominated by the
accuracies of the measurements of electrons, positrons,
and antiprotons. This enables us to study the overall rigidity
dependent behavior of different fluxes as shown in Fig. 2
above 10 GV. The points are placed along the abscissa at
ˆ
R
calculated for a flux jRj2.7[36]. As seen from Fig. 2, the
rigidity dependence of the fluxes for antiprotons, positrons,
and protons are nearly identical above 60 GV whereas the
rigidity dependence of the electron flux is different.
Analysis of the antiproton spectral index γ¯
pis performed
over independent rigidity intervals with a variable width to
have sufficient sensitivity to γ¯p. The spectral index is
calculated from γ¯p¼d½logðΦ¯pÞ=d½logðjRjÞ. The result is
presented in Fig. 3 of Supplemental Material [18] com-
pared with our result on the proton spectral index [16].As
seen, over the four lowest rigidity points, jRj<60.3GV,
PRL 117, 091103 (2016) PHYSICAL REVIEW LETTERS week ending
26 AUGUST 2016
091103-6
the antiproton spectral index decreases more rapidly than
the proton spectral index and for the highest rigidity
interval, 60.3jRj<450 GV, the antiproton spectral
index is consistent with the proton spectral index.
Figure 3(a) presents the measured ( ¯
p=p) flux ratio.
Compared with earlier experiments [2,6], the AMS results
extend the rigidity range to 450 GV with increased
precision. Figure 2 of Supplemental Material [18] shows
the low energy (<10 GeV) part of our measured ( ¯
p=p)
flux ratio. To minimize the systematic error for this flux
ratio we have used the 2.42 ×109protons selected with the
same acceptance, time period, and absolute rigidity range
as the antiprotons. From 10 to 450 GV, the values of the
proton flux are identical to 1% to those in our publication
[16]. As seen from Fig. 3(a), above 60 GV the ratio
appears to be rigidity independent.
To estimate the lowest rigidity above which the ( ¯
p=p)
flux ratio is rigidity independent, we use rigidity intervals
with starting rigidities from 10 GV and increasing bin by
bin. The ending rigidity for all intervals is fixed at 450 GV.
Each interval is split into two sections with a boundary
between the starting rigidity and 450 GV. Each of the two
sections is fit with a constant and we obtain two mean
values of the ( ¯
p=p) flux ratio. The lowest starting rigidity of
the interval that gives consistent mean values at the
90% C.L. for any boundary defines the lowest limit.
This yields 60.3 GV as the lowest rigidity above which
the ( ¯
p=p) flux ratio is rigidity independent with a mean
value of ð1.81 0.04Þ×104. To further probe the behav-
ior of the flux ratio we define the best straight line fit over a
rigidity interval as
ð¯
p=pÞ¼CþkðjRjR0Þ;ð4Þ
where Cis the value of the flux ratio at R0,kis the slope, and
R0is chosen to minimize the correlation between the fitted
values of Cand k, i.e., the mean of jRjover the interval
weighted with the statistical and uncorrelated systematic
errors. The solid red line in Fig. 3(a) shows this best straight
line fit above 60.3 GV, as determined above, together with
the 68% C.L. range of the fit parameters (shaded region).
Above 60.3 GV, R0¼91 GV. The fitted value of the slope,
k¼ð0.70.9Þ×107GV1, is consistent with zero.
With the AMS measurements on the fluxes of all charged
elementary particles in cosmic rays, ¯
p,p,eþ, and e,we
can now study the rigidity dependent behavior of different
flux ratios. The flux ratios and errors are tabulated in Tables
II and III of Supplemental Material [18]. For the antiproton-
to-positron ratio the rigidity independent interval is 60.3
jRj<450 GV with a mean value of 0.479 0.014. Fitting
Eq. (4) over this interval yields kð¯
p=eþÞ¼ð2.83.2Þ×
104GV1. For the proton-to-positron ratio, the rigidity
independent interval is 59.13 jRj<500 GV with a mean
value of ð2.67 0.05Þ×103and kðp=eþÞ¼ð0.9
1.0ÞGV1. Both results are shown in Fig. 3(b) together
with the 68% C.L. range of the fit parameters (shaded
regions). In the study of the ratios, we have taken into
account the correlation of the errors due to uncertainty in
the ECAL energy scale in Φe[15].
In Fig. 4 of Supplemental Material [18] we present our
measured antiproton-to-electron and proton-to-electron
flux ratios. Both of these flux ratios exhibit rigidity
2
10
10 3
10
5
10
15
20
25
3
10×
2
4
6
8
20
40
60
80
100
120
1
2
3
4
e-
e+
p
p
]
1.7
GV
-1
s
-1
sr
-2
[m
2.7
R
Φ
|Rigidity| [GV]
FIG. 2. The measured antiproton flux (red, left axis) compared
to the proton flux (blue, left axis) [16], the electron flux (purple,
right axis), and the positron flux (green, right axis) [15]. All the
fluxes are multiplied by
ˆ
R2.7. The fluxes show different behavior
at low rigidities while at jRjabove 60 GV the functional
behavior of the antiproton, proton, and positron fluxes are nearly
identical and distinctly different from the electron flux. The error
bars correspond to the quadratic sum of the statistical and
systematic errors.
|Rigidity| [GV]
100 200 300 400 500
-5
10
-1
10
-2
10
4
10
3
10
-4
10
AMS-02
PAMELA
(a)
(b)
+
0
p
Φ /Φ ratio
e
+
p
Φ /Φ
e
+
p
Φ /Φ
e
+
p
Φ /Φ ratio
e
Φ /Φ ratio
pp
FIG. 3. (a) The measured ( ¯
p=p) flux ratio as a function of the
absolute value of the rigidity from 1 to 450 GV. The PAMELA [6]
measurement is also shown. (b) The measured ( ¯
p=eþ) (red, left
axis) and (p=eþ) (blue, right axis) flux ratios. The solid lines show
the best fit of Eq. (4) to the data above the lowest rigidity consistent
with rigidity independence together with the 68% C.L. ranges of
the fit parameters (shaded regions). For the AMS data, the error
bars are the quadratic sum of statistical and systematic errors.
Horizontally, the data points are placed at the center of each bin.
PRL 117, 091103 (2016) PHYSICAL REVIEW LETTERS week ending
26 AUGUST 2016
091103-7
behavior which is distinct from that observed in the
antiproton-to-proton, antiproton-to-positron, and proton-
to-positron flux ratios.
To examine the rigidity dependence of the flux ratios
shown in Figs. 3(a) and 3(b) quantitatively in a model
independent way, Eq. (4) is fit to the flux ratios over their
rigidity ranges with a sliding window. For each flux ratio,
the width of the window varies with rigidity to have
sufficient sensitivity to the slope ksuch that each window
covers between four and eight bins. The variations of Cand
slope kfor the ( ¯
p=p) flux ratio are shown in Fig. 4.Atlow
rigidity the slope kcrosses zero, that is, the ratio reaches a
maximum at 20 GV as also clearly seen in the parameter
C. As seen from Fig. 5 of Supplemental Material [18], the
rigidity dependence of the ( ¯
p=eþ) and (p=eþ) flux ratios
are nearly identical to that of the ( ¯
p=p) flux ratio. Also
shown in Fig. 4, as well as in Fig. 5 of the Supplemental
Material [18], are the mean values of the flux ratios over the
intervals where they are rigidity independent.
In conclusion, with this measurement of the antiproton
flux and the ( ¯
p=p) flux ratio, AMS has simultaneously
measured all the charged elementary particle cosmic ray
fluxes and flux ratios. In the absolute rigidity range 60 to
500 GV, the antiproton, proton, and positron fluxes are
found to have nearly identical rigidity dependence and the
electron flux exhibits a different rigidity dependence. In the
absolute rigidity range below 60 GV, the ( ¯
p=p), ( ¯
p=eþ),
and (p=eþ) flux ratios each reaches a maximum. In the
absolute rigidity range 60 to 500 GV, the ( ¯
p=p),
(¯
p=eþ), and (p=eþ) flux ratios show no rigidity depend-
ence. These are new observations of the properties of
elementary particles in the cosmos.
We thank former NASA Administrator Daniel S. Goldin
for his dedication to the legacy of the ISS as a scientific
laboratory and his decision for NASA to fly AMS as a
DOE payload. We also acknowledge the continuous
support of the NASA leadership including Charles
Bolden and William H. Gerstenmaier and of the JSC
and MSFC flight control teams which has allowed
AMS to operate optimally on the ISS for five years.
We are grateful for the support of Jim Siegrist and his staff
of the DOE. We also acknowledge the continuous support
from MIT and its School of Science, Michael Sipser, Marc
Kastner, Ernest Moniz, Richard Milner, and Boleslaw
Wyslouch. Research supported by São Paulo Research
Foundation (FAPESP) Grants No. 2014/19149-7,
No. 2014/50747-8, and No. 2015-50378-5, Brazil;
CAS, NSFC, MOST, NLAA, the provincial governments
of Shandong, Jiangsu, Guangdong, and the China
Scholarship Council, China; the Finnish Funding
Agency for Innovation (Tekes) Grants No. 40361/01
and No. 40518/03 and the Academy of Finland Grant
No. 258963, Finland; CNRS, IN2P3, CNES, Enigmass,
and the ANR, France; Pascale Ehrenfreund, DLR, and
JARA-HPC under Project No. JARA0052, Germany;
INFN and ASI under ASI-INFN Agreements No. 2013-
002-R.0 and No. 2014-037-R.0, Italy; CHEP Grants
No. NRF-2009-0080142 and No. NRF-2012-010226 at
Kyungpook National University and No. NRF-2013-
004883 at Ewha Womans University, Korea; the
Consejo Nacional de Ciencia y Tecnología and UNAM,
Mexico; FCT under Grant No. PTDC/FIS/122567/2010,
Portugal; CIEMAT, IAC, CDTI, and SEIDI-MINECO
under Grants No. AYA2012-39526-C02-(01/02),
No. ESP2015-71662-C2-(1-P/2-P), No. SEV-2011-0187,
No. SEV-2015-0548, and No. MDM-2015-0509, Spain;
the Swiss National Science Foundation (SNSF), federal
and cantonal authorities, Switzerland; Academia Sinica
and the Ministry of Science and Technology (MOST)
under Grants No. 103-2112-M-006-018-MY3, No. 104-
2112-M-001-027, and No. CDA-105-M06, former
President of Academia Sinica Yuan-Tseh Lee, and former
Ministers of MOST Maw-Kuen Wu and Luo-Chuan Lee,
Taiwan; the Turkish Atomic Energy Authority at METU,
Turkey; and NSF Grant No. 1455202, Wyle Laboratories
Grant No. 2014/T72497, and NASA NESSF Grant
No. HELIO15F-0005, USA. We gratefully acknowledge
the strong support from CERN including Rolf-Dieter
Heuer and Fabiola Gianotti, from the CERN IT depart-
ment and Bernd Panzer-Steindel, and from the European
Space Agency including Johann-Dietrich Wörner and
Simonetta Di Pippo. We are grateful for important dis-
cussions with Fiorenza Donato, Jonathan Ellis, Jonathan
Feng, Igor Moskalenko, Michael Salamon, Subir Sarkar,
Joachim Trümper, Michael S. Turner, Steven Weinberg,
and Arnold Wolfendale.
aAlso at ASI, I00133 Roma, Italy.
bPresent address: Departamento de Ciencias, Pontifica
Universidad Católica del Perú (PUCP), Lima 32, Peru.
1.0
1.2
1.4
1.6
1.8
2.0
2.2 1.0
10
C [10 ]
10
0.0
0.2
0.4
0.6
0.8
C
k
-4
6
-1-
k[10 GV ]
|Rigidity| [GV]
pp
Φ /Φ
ratio fit
2
FIG. 4. Sliding window fits of Eq. (4) to the ( ¯
p=p) flux ratio
measured by AMS with parameter C(green, left axis) and the
slope k(blue, right axis). The green and blue shaded regions
indicate that the errors are correlated between adjacent points.
The points are placed at R0. The dashed blue line at k¼0is to
guide the eye. The black arrow indicates the lowest rigidity above
which the flux ratio is consistent with being rigidity independent
and the black horizontal band shows the mean value and the
1-sigma error of the flux ratio above this rigidity.
PRL 117, 091103 (2016) PHYSICAL REVIEW LETTERS week ending
26 AUGUST 2016
091103-8
cAlso at ASI Science Data Center (ASDC), I00133 Roma,
Italy.
dAlso at Sun YatSen University (SYSU), Guangzhou,
510275, China.
eAlso at Wuhan University, Wuhan 430072, China.
fAlso at Harbin Institute of Technology (HIT), Harbin
150001, China.
gAlso at Huazhong University of Science and Technology
(HUST), Wuhan, 430074, China.
hAlso at Università di Siena, I-53100 Siena, Italy.
iAlso at Laboratoire dAnnecyleVieux de Physique Thé-
orique (LAPTh), CNRS and Université Savoie Mont Blanc,
F-74941 AnnecyleVieux, France.
jAlso at Shandong University (SDU), Jinan, Shandong,
250100, China.
kAlso at Nankai University, Tianjin 300071, China.
lAlso at Southeast University (SEU), Nanjing, 210096,
China.
mAlso at Institute of Theoretial Physics, Chinese Academy
of Sciences, Beijing 100190, China.
nAlso at Jilin University, Jilin 130012, China.
[1] E. A. Bogomolov et al.,inProceedings of the 16th
International Cosmic Ray Conference, University of Tokio,
Tokio, Japan (1979) Vol. 1, p. 330; R. L. Golden, S. Horan,
B. G. Mauger, G. D. Badhwar, J. L. Lacy, S. A. Stephens, R. R
Daniel, and J. E. Zipse, Phys. Rev. Lett. 43, 1196 (1979);A.
Buffington and S. M. Shindler, Astrophys. J. 247,L105
(1981).
[2] K. Yoshimura et al.,Phys. Rev. Lett. 75, 3792 (1995);S.
Orito et al.,Phys. Rev. Lett. 84, 1078 (2000); Y. Asaoka
et al.,Phys. Rev. Lett. 88, 051101 (2002); K. Abe et al.,
Phys. Lett. B 670, 103 (2008);Phys. Rev. Lett. 108, 051102
(2012);Astrophys. J. 822, 65 (2016).
[3] M. Boezio et al.,Astrophys. J. 487, 415 (1997);561, 787
(2001).
[4] A. S. Beach et al.,Phys. Rev. Lett. 87, 271101 (2001).
[5] M. Aguilar et al.,Phys. Rep. 366, 331 (2002).
[6] O. Adriani et al.,Phys. Rev. Lett. 102, 051101 (2009);105,
121101 (2010);JETP Lett. 96, 621 (2013).
[7] C. S. Shen and G. B. Berkey, Phys. Rev. 171, 1344 (1968);
I. V. Moskalenko and A. W. Strong, Astrophys. J. 493, 694
(1998); M. Simon, A. Molnar, and S. Roesler, Astrophys. J.
499, 250 (1998); I. V. Moskalenko, A. W. Strong, J. F.
Ormes, and M. S. Potgieter, Astrophys. J. 565, 280
(2002); V. S. Ptuskin, I. V. Moskalenko, F. C. Jones, A.
W. Strong, and V. N. Zirakashvili, Astrophys. J. 642, 902
(2006); F. Donato, D. Maurin, P. Brun, T. Delahaye, and P.
Salati, Phys. Rev. Lett. 102, 071301 (2009).
[8] G. Jungman, M. Kamionkowski, and K. Griest, Phys. Rep.
267, 195 (1996); L. Bergström, Rep. Prog. Phys. 63, 793
(2000); G. Bertone, D. Hooper, and J. Silk, Phys. Rep. 405,
279 (2005); G. Kane, R. Lu, and S. Watson, Phys. Lett. B 681,
151 (2009); F. Donato, D. Maurin, P. Brun, T. Delahaye, and
P. Salati, Phys. Rev. Lett. 102, 071301 (2009).
[9] P. Blasi and P. D. Serpico, Phys. Rev. Lett. 103, 081103
(2009); Y. Fujita, K. Kohri, R. Yamazaki, and K. Ioka, Phys.
Rev. D 80, 063003 (2009).
[10] S. W. Hawking, Nature (London) 248, 30 (1974); P. Kiraly,
J. Szabelski, J. Wdowczyk, and A. W. Wolfendale, Nature
(London) 293, 120 (1981).
[11] M. Aguilar et al.,Phys. Rev. Lett. 110, 141102 (2013);
L. Accardo et al.,Phys. Rev. Lett. 113, 121101 (2014).
[12] J. Kopp, Phys. Rev. D 88, 076013 (2013); L. Feng, R-Z.
Yang, H-N. He, T-K. Dong, Y-Z. Fan, and J. Chang, Phys.
Lett. B 728, 250 (2014); L. Bergström, T. Bringmann, I.
Cholis, D. Hooper, and C. Weniger, Phys. Rev. Lett. 111,
171101 (2013); C. H. Chen, C. W. Chiang, and T. Nomura,
Phys. Lett. B 747, 495 (2015); H. B. Jin, Y. L. Wu, and Y.-F.
Zhou, Phys. Rev. D 92, 055027 (2015).
[13] I. Cholis and D. Hooper, Phys. Rev. D 88, 023013 (2013);T.
Linden and S. Profumo, Astrophys. J. 772, 18 (2013);P.
Mertsch and S. Sarkar, Phys. Rev. D 90, 061301 (2014);K.
Kohri, K. Ioka, Y. Fujita, and R. Yamazaki, Prog. Theor.
Exp. Phys. 2016, 021E01 (2016).
[14] K. Blum, B. Katz, and E. Waxman, Phys. Rev. Lett. 111,
211101 (2013); R. Cowsik, B. Burch, and T. Madziwa-
Nussinov, Astrophys. J. 786, 124 (2014); R. Kappl and
M. W. Winkler, J. Cosmol. Astropart. Phys. 09 (2014) 051.
[15] M. Aguilar et al.,Phys. Rev. Lett. 113, 121102 (2014).
[16] M. Aguilar et al.,Phys. Rev. Lett. 114, 171103 (2015).
[17] A. Kounine, Int. J. Mod. Phys. E 21, 1230005 (2012);S.
Rosier-Lees, in Proceedings of the 19th Symposium on
Astroparticle Physics in the Netherlands, Beekbergen, 2014
(unpublished); S. C. C. Ting, Nucl. Phys. B, Proc. Suppl.
243244, 12 (2013); S. C. Lee, in Proceedings of the 20th
International Conference on Supersymmetry and Unifica-
tion of Fundamental Interactions (SUSY 2012), Beijing,
2012 (unpublished); M. Aguilar, in Proceedings of the XL
International Meeting on Fundamental Physics, Centro de
Ciencias de Benasque Pedro Pascual, 2012 (unpublished);
S. Schael, in Proceedings of the 10th Symposium on Sources
and Detection of Dark Matter and Dark Energy in the
Universe, Los Angeles, 2012 (unpublished); B. Bertucci,
Proc. Sci., EPS-HEP (2011) 67; M. Incagli, AIP Conf. Proc.
1223, 43 (2010); R. Battiston, Nucl. Instrum. Methods Phys.
Res., Sect. A 588, 227 (2008).
[18] See Supplemental Material at http://link.aps.org/
supplemental/10.1103/PhysRevLett.117.091103 for details
of the AMS detector, its test beam calibration, and the
simulation; for the tabulated ¯
pflux Φ¯
pand the ( ¯
p=p),
(¯
p=eþ), ( ¯
p=e), (p=eþ), and (p=e) flux ratios; and for
figures of an antiproton event display, the low energy
antiproton-to-proton flux ratio, the antiproton spectral in-
dex, the ( ¯
p=e) and (p=e) flux ratios, and sliding window
fit parameters of the ( ¯
p=eþ) and (p=eþ) flux ratios.
[19] B. Alpat et al.,Nucl. Instrum. Methods Phys. Res., Sect. A
613, 207 (2010).
[20] K. Lübelsmeyer et al.,Nucl. Instrum. Methods Phys. Res.,
Sect. A 654, 639 (2011).
[21] V. Bindi et al.,Nucl. Instrum. Methods Phys. Res., Sect. A
743, 22 (2014), and references therein.
[22] Ph. von Doetinchem, W. Karpinski, Th. Kirn, K.
Lübelsmeyer, St. Schael, and M. Wlochal, Nucl. Phys. B,
Proc. Suppl. 197, 15 (2009).
[23] F. Hauler et al.,IEEE Trans. Nucl. Sci. 51, 1365 (2004); Ph.
Doetinchem et al.,Nucl. Instrum. Methods Phys. Res., Sect.
A558, 526 (2006); Th. Kirn, Nucl. Instrum. Methods Phys.
Res., Sect. A 706, 43 (2013).
[24] M. Aguilar-Benitez et al.,Nucl. Instrum. Methods
Phys. Res., Sect. A 614, 237 (2010); F. Giovacchini,
PRL 117, 091103 (2016) PHYSICAL REVIEW LETTERS week ending
26 AUGUST 2016
091103-9
Nucl. Instrum. Methods Phys. Res., Sect. A 766,57
(2014).
[25] C. Adloff et al.,Nucl. Instrum. Methods Phys. Res., Sect. A
714, 147 (2013).
[26] B. P. Roe, H. J. Yang, J. Zhu, Y. Liu, I. Stancu, and G.
McGregor, Nucl. Instrum. Methods Phys. Res., Sect. A 543,
577 (2005).
[27] J. Alcaraz et al.,Phys. Lett. B 484, 10 (2000).
[28] C. C. Finlay et al.,Geophys. J. Int. 183, 1216 (2010);E.
Thébault et al.,Earth Planets Space 67, 79 (2015).
[29] P. Papini et al.,Astrophys. J. 615, 259 (2004).
[30] S. Agostinelli et al.,Nucl. Instrum. Methods Phys. Res.,
Sect. A 506, 250 (2003); J. Allison et al.,IEEE Trans. Nucl.
Sci. 53, 270 (2006).
[31] A. Kondor, Nucl. Instrum. Methods Phys. Res. 216, 177
(1983); V. Blobel, Report No. DESY-84-118 (1984);
G. DAgostini, Nucl. Instrum. Methods Phys. Res., Sect.
A362, 487 (1995).
[32] T. Bowen, M. Di Corato, W. H. Moore, and G. Tagliaferri, II
Nuovo Cimento 9, 908 (1958); G. Bellettini, G. Cocconi,
A. N. Diddens, E. Lillethun, G. Matthiae, J. P. Scanlon,
and A. M. Wetherell, Nucl. Phys. 79, 609 (1966);S.P.
Denisov, S. V. Donskov, Yu. P. Gorin, R. N. Krasnokutsky,
A. I. Petrukhin, Yu. D. Prokoshkin, and D. A. Stoyanova,
Nucl. Phys. B61, 62 (1973); A. S. Carroll et al.,Phys. Lett.
B80, 319 (1979); J. R. Letaw, R. Silberberg, and C. H.
Tsao, Astrophys. J. Suppl. Ser. 51, 271 (1983); N. Abgrall
et al.,Phys. Rev. C 84, 034604 (2011).
[33] B. Cork, G. R. Lambertson, O. Piccioni, and W. A. Wenzel,
Phys. Rev. 107, 248 (1957); J. V. Allaby et al., Yad. Fiz. 12,
538 (1970); R. J. Abrams, R. L. Cool, G. Giacomelli, T. F.
Kycia, B. A. Leontic, K. K. Li, A. Lundby, D. N. Michael,
and J. Teiger, Phys. Rev. D 4, 3235 (1971); H. Aihara et al.,
Nucl. Phys. A360, 291 (1981); K. Nakamura, J. Chiba,
T. Fujii, H. Iwasaki, T. Kageyama, S. Kuribayashi, T.
Sumiyoshi, T. Takeda, H. Ikeda, and Y. Takada, Phys.
Rev. Lett. 52, 731 (1984).
[34] X. Yan and X. G. Su, Linear Regression Analysis (World
Scientific, Singapore, 2009), ISBN-13 978-981 283 410 2;
A. Hoecker et al.,arXiv:physics/0703039v5.
[35] AMS Collaboration, Measurement of the Time Dependent
Solar Modulation of Primary Cosmic Rays Fluxes with the
Alpha Magnetic Spectrometer on the International Space
Station (to be published).
[36] G. D. Lafferty and T. R. Wyatt, Nucl. Instrum. Methods
Phys. Res., Sect. A 355, 541 (1995). We have used Eq. (6)
with
ˆ
Rxlw.
PRL 117, 091103 (2016) PHYSICAL REVIEW LETTERS week ending
26 AUGUST 2016
091103-10
... On the other hand, describing the production and propagation of these charged hadrons is a challenging task, and any statement regarding DM annihilation rests on our ability to understand the associated astrophysical uncertainties. The Alpha Magnetic Spectrometer (AMS-02) experiment has provided the most precise measurements of the cosmic ray proton and antiproton flux to date [86], and here we will explore the implications of this data on our neutrino portal DM scenario. Since DM annihilates to RHNs, which subsequently decay via W , Z, and Higgs bosons, the resulting cascade decay, showering and hadronization produce a variety of hadronic final states including antiprotons. ...
... AMS-02 has provided precise measurements of the proton flux, Φ p (K) [90], and the antiproton-to-proton flux ratio, r(K) [86], which can be used to place constraints on DM annihilation. To proceed, we require an estimate of the secondary background antiproton flux originating from astrophysical sources. ...
... where i runs over energy bins, and σ i is the reported uncertainty of the flux ratio [86]. Following Ref. [91], we define a limit on σv as a function of m χ , m N according to the ...
Preprint
We investigate the feasibility of the indirect detection of dark matter in a simple model using the neutrino portal. The model is very economical, with right-handed neutrinos generating neutrino masses through the Type-I seesaw mechanism and simultaneously mediating interactions with dark matter. Given the small neutrino Yukawa couplings expected in a Type-I seesaw, direct detection and accelerator probes of dark matter in this scenario are challenging. However, dark matter can efficiently annihilate to right-handed neutrinos, which then decay via active-sterile mixing through the weak interactions, leading to a variety of indirect astronomical signatures. We derive the existing constraints on this scenario from Planck cosmic microwave background measurements, Fermi dwarf spheroidal galaxies and Galactic Center gamma-rays observations, and Alpha Magnetic Spectrometer - 02 antiprotons observations, and also discuss the future prospects of Fermi and the Cherenkov Telescope Array. Thermal annihilation rates are already being probed for dark matter lighter than about 50 GeV, and this can be extended to dark matter masses of 100 GeV and beyond in the future. This scenario can also provide a dark matter interpretation of the Fermi Galactic Center gamma ray excess, and we confront this interpretation with other indirect constraints. Finally we discuss some of the exciting implications of extensions of the minimal model with large neutrino Yukawa couplings and Higgs portal couplings.
... As a result, after rescaling by f 2 T 0 , the overall thermally averaged annihilation cross section, ⟨σv⟩ W + W − , is suppressed and remains below the experimental sensitivity reaches, as illustrated in Fig. 2 13 . In this Figure 2. Bounds on the scalar Triplet mass m T 0 from indirect searches of W + W − channel by Fermi-LAT [330,333] (depicted by golden-yellow coloured patterned region) and AMS-02 [345] (depicted by pink coloured patterned region). The two differently coloured dashed lines correspond to two different choices of the triplet SM-like Higgs coupling parameter, λ HT , as depicted on the plot. ...
... for two different values of λ HT = 0.01 (green coloured line) and 0.30 (hot-pink coloured line), with the upper bounds on ⟨σv⟩ W + W − derived from satellite-based Fermi-LAT data [330,333] (represented by the golden-yellow coloured patterned region) and AMS-02 data [345] (depicted by the pink-coloured patterned region). The upward trend of the cross-section lines in Fig. 2 can be attributed to the fact that, as m T 0 increases, the relic density also rises, leading to an increase in the ratio f T 0 (see Eq. (3.17)). ...
Preprint
Full-text available
We explore the gravitational wave probes of a two-component dark matter framework, consisting of an SU(2)LSU(2)_L triplet scalar and a Standard Model singlet fermion. The triplet scalar dark matter typically remains underabundant in the region below 1.9\sim 1.9 TeV, due to the strong SU(2)LSU(2)_L gauge mediated interactions. We introduce a second dark matter component, an SU(2)LSU(2)_L singlet vector-like Dirac fermion, to address this deficit in the dark matter relic abundance within a sub-TeV range. A key aspect of the proposed setup is the potential dark matter inter-conversion between the two components, which impacts the dark matter freeze-out dynamics and relic density of individual dark matter components. In such a scenario, we examine the properties of electroweak phase transition and identify the regions of parameter space that exhibit strong first-order phase transition. We estimate the resulting gravitational wave spectrum and its detectability, which could be probed through the conventional power-law-integrated sensitivity limits and the recently proposed peak-integrated sensitivity curves. Our analysis reveals that a novel region of the model's parameter space, compatible with dark matter observables, can generate a detectable gravitational wave spectrum, observable by upcoming space-based gravitational wave detectors such as LISA, BBO, DECIGO, and DECIGOcorr, while also offering complementary detection prospects in the dark matter and collider experiments.
... Charged particles are strongly affected by the magnetic fields and interstellar medium so that their propagation is quite complex to model. Nevertheless stringent constraint on the DM annihilation cross section was obtained in [62] using the antiproton data measured by Alpha Magnetic Spectrometer (AMS-02) [63]. Energy injection from the DM annihilation in the early Universe affects the anisotropies of the cosmic microwave background (CMB). ...
Article
Full-text available
According to recent measurements, the dark matter magnetic dipole moment is strongly constrained. In the composite Higgs models the magnetic dipole moment of the Dirac dark matter fermion and its mass can be suppressed by the approximate U ( 1 ) symmetry. We consider the E 6 inspired composite Higgs model ( E 6 CHM ) with U ( 1 ) symmetry violating operators, which give rise to dark matter’s mass and coupling constant to Higgs boson. The dependence of the spin-independent dark matter-nucleon scattering cross section on the E 6 CHM parameters is explored. We argue that there are regions of the parameter space which are still safe from all current constraints and may lead to spectacular LHC signatures. Published by the American Physical Society 2024
... The latest debates seem to refocus on a perplexing rise of the e + /e − fraction in the 10-300 GeV range, and possibly related flatness ofp/p and e + /p ratios. They gained widespread interest after the AMS-02 team has published the most accurate elemental and energy spectra in this range [3]. ...
Preprint
Recent observations of galactic cosmic rays (CR) in the 1-500 GeV energy range have revealed striking deviations from what deemed "standard." The anomalies cut across hadronic and leptonic CRs. I discuss findings that challenge physical mechanisms long held responsible for the CR production in galactic supernova remnants (SNR). I also consider some new physics of particle acceleration in SNR shocks that is not part of conventional models but may explain the anomalies. However, a possible 20-30\% excess remains unaccounted for in the e+/e+e^{+}/e^{+} ratio over the range of a few 100 GeV. If not explained by future models, it suggests an additional source of positrons such as a dark matter decay/annihilation or pulsar contribution. Earlier efforts to explain both the e+/ee^{+}/e^{-} and p/He anomalies with the "standard" models by adjusting the SNR environmental parameters and multiple sources are critically assessed.
... The antiproton component in cosmic rays has been recognized since decades as a potentially important channel for DM searches. The recent release [144] of high-precision data by the Ams-02 experiment concerning the antiproton flux and antiproton to proton ratio has on one side strengthened the relevance of antiprotons and on the other side pointed to a need for a better determination of the astrophysical background (called 'secondary antiprotons'). At the moment, such background can explain the data, within the rather large uncertainties, without the need for an additional exotic component. ...
Preprint
Dark matter (DM) charged under a dark U(1) force appears in many extensions of the Standard Model, and has been invoked to explain anomalies in cosmic-ray data, as well as a self-interacting DM candidate. In this paper, we perform a comprehensive phenomenological analysis of such a model, assuming that the DM abundance arises from the thermal freeze-out of the dark interactions. We include, for the first time, bound-state effects both in the DM production and in the indirect detection signals, and quantify their importance for Fermi, AMS, and CMB experiments. We find that DM in the mass range 1 GeV to 100 TeV, annihilating into dark photons of MeV to GeV mass, is in conflict with observations. Instead, DM annihilation into heavier dark photons is viable. We point out that the late decays of multi-GeV dark photons can produce significant entropy and thus dilute the DM density. This can lower considerably the dark coupling needed to obtain the DM abundance, and in turn relax the existing constraints.
... • Other indirect detection signals: Fermi-LAT and DES measurements of dwarf spheroidal galaxies can be relevant particularly at low DM mass [132,133], and large positron signals [134] can be constrained by AMS-02 [135]. Also note that Fermi-LAT observed the Sun, searching for longlived mediators directly decaying to electrons in the DM mass range 70-2000 GeV, which are stronger than the gamma-ray limits [136]. ...
Preprint
Dark matter capture and annihilation in the Sun can produce detectable high-energy neutrinos, providing a probe of the dark matter-proton scattering cross section. We consider the case when annihilation proceeds via long-lived dark mediators, which allows gamma rays to escape the Sun and reduces the attenuation of neutrinos. For gamma rays, there are exciting new opportunities, due to detailed measurements of GeV solar gamma rays with Fermi, and unprecedented sensitivities in the TeV range with HAWC and LHAASO. For neutrinos, the enhanced flux, particularly at higher energies (\simTeV), allows a more sensitive dark matter search with IceCube and KM3NeT. We show that these search channels can be extremely powerful, potentially improving sensitivity to the dark matter spin-dependent scattering cross section by several orders of magnitude relative to present searches for high-energy solar neutrinos, as well as direct detection experiments.
Article
Full-text available
In Radio Super Novae (RSNe) a magnetic field of ( B × r ) = 1 0 16.0 ± 0.12 Gauss × cm is observed; these are the same numbers for Blue Super Giant (BSG) star explosions as for Red Super Giant (RSG) star explosions, despite their very different wind properties. The EHT data for M87 as well for low power radio galaxies all show consistency with just this value of the quantity ( B × r ) , key for angular momentum and energy transport, and can be derived from the radio jet data. We interpret this as a property of the near surroundings of a black hole (BH) at near maximal rotation, independent of BH mass. In the commonly used green onion model, in which a 2 π flow changes over to a jet flow we interpret this as a wind emanating from the BH/accretion disk system and its surroundings. Near the BH collisions in the wind can produce a large fraction of anti-protons. In this scenario the cosmic Ray (CR) population from the wind/jet is proposed to be visible as EeV protons and anti-protons in the CR data to EeV energy, with a E − 7 / 3 spectrum. This can be connected to a concept of inner and outer Penrose zones in the ergo-region. The observed numbers for the magnetic field imply the Planck time as the governing time scale: A BH rotating near maximum can accept a proton per log bin of energy in an extended spectrum with the associated pions every Planck time.
Article
Full-text available
It is conceivable that a bosonic dark matter (DM) with non-gravitational interactions with SM particles will be accumulated at the center of a neutron star (NS) and can lead to black hole formation. In contrast to previous works with a fixed NS temperature, we dynamically determine the formation of Bose-Einstein condensate (BEC) for a given set of DM parameters, namely the DM-neutron scattering cross-section (σχn), the thermal average of DM annihilation cross-section (⟨σv⟩) and the DM mass (mχ). For both non-annihilating and annihilating DM with ⟨σv⟩ ≲ 10⁻²⁶ cm³ s⁻¹, the BEC forms for mχ ≲ 10 TeV. In case of non-annihilating DM, observations of old NS allows σχn ≲ 10⁻⁵² cm² for 10 MeV ≤ mχ ≲ 10 GeV (with BEC) and σχn ≲ 10⁻⁴⁷ cm² for 5 TeV ≲ mχ ≲ 30 PeV (without BEC). This analysis shows that the electroweak mass window, 10 GeV ≲ mχ ≲ 5 TeV is essentially unconstrained by NS observations and therefore is subject only to direct detection experiments. In the annihilating DM scenario, the exclusion limits on DM parameters become weaker and even vanish for typical WIMP annihilation cross-section. However, the late-time heating of the NS enables us to probe the region with σχn ≳ 10⁻⁴⁷ cm², using the James Webb Space Telescope in the foreseeable future. When our results are viewed in the context of indirect searches of DM, it provides a lower limit on the ⟨σv⟩, which is sensitive to the DM thermal state.
Article
Full-text available
The 12th generation of the International Geomagnetic Reference Field (IGRF) was adopted in December 2014 by the Working Group V-MOD appointed by the International Association of Geomagnetism and Aeronomy (IAGA). It updates the previous IGRF generation with a definitive main field model for epoch 2010.0, a main field model for epoch 2015.0, and a linear annual predictive secular variation model for 2015.0-2020.0. Here, we present the equations defining the IGRF model, provide the spherical harmonic coefficients, and provide maps of the magnetic declination, inclination, and total intensity for epoch 2015.0 and their predicted rates of change for 2015.0-2020.0. We also update the magnetic pole positions and discuss briefly the latest changes and possible future trends of the Earth’s magnetic field.
Article
Full-text available
The BESS-Polar Collaboration measured the energy spectra of cosmic-ray protons and helium during two long-duration balloon flights over Antarctica in December 2004 and December 2007, at substantially different levels of solar modulation. Proton and helium spectra probe the origin and propagation history of cosmic rays in the galaxy, and are essential to calculations of the expected spectra of cosmic-ray antiprotons, positrons, and electrons from interactions of primary cosmic-ray nuclei with the interstellar gas, and to calculations of atmospheric muons and neutrinos. We report absolute spectra at the top of the atmosphere for cosmic-ray protons in the kinetic energy range 0.2-160 GeV and helium nuclei 0.2-80 GeV/nucleon. The corresponding magnetic rigidity ranges are 0.6-160 GV for protons and 1.3-160 GV for helium. These spectra are compared to measurements from previous BESS flights and from AMS-01, ATIC-2, PAMELA, and AMS-02. We also report the ratio of the proton and helium fluxes from 1.3 GV to 160 GV and compare to ratios from PAMELA and AMS-02.
Article
Full-text available
We explain the excess of the antiproton fraction recently reported by the AMS-02 experiment by considering collisions between cosmic-ray protons accelerated by a local supernova remnant (SNR) and the surrounding dense cloud. The same "pp collisions" provide the right branching ratio to fit the observed positron excess simultaneously without a fine tuning. The supernova happened in relatively lower metalicity than the major cosmic-ray sources. The cutoff energy of electrons marks the supernova age of ~10^{5} years, while the antiproton excess may extend to higher energy. Both antiproton and positron fluxes are completely consistent with our predictions in Fujita, Kohri, Yamazaki and Ioka (2009).
Article
Full-text available
A precise measurement of the proton flux in primary cosmic rays with rigidity (momentum/charge) from 1 GV to 1.8 TV is presented based on 300 million events. Knowledge of the rigidity dependence of the proton flux is important in understanding the origin, acceleration, and propagation of cosmic rays. We present the detailed variation with rigidity of the flux spectral index for the first time. The spectral index progressively hardens at high rigidities.
Article
Full-text available
We propose a dark matter explanation to simultaneously account for the excess of antiproton-to-proton and positron power spectra observed in the AMS-02 experiment while having the right dark matter relic abundance and satisfying the current direct search bounds. We extend the Higgs triplet model with a hidden gauge symmetry of SU(2)XSU(2)_X that is broken to Z3Z_3 by a quadruplet scalar field, rendering the associated gauge bosons stable weakly-interacting massive particle dark matter candidates. By coupling the complex Higgs triplet and the SU(2)XSU(2)_X quadruplet, the dark matter candidates can annihilate into triplet Higgs bosons each of which in turn decays into lepton or gauge boson final states. Such a mechanism gives rise to correct excess of positrons and antiprotons with an appropriate choice of the triplet vacuum expectation value. Besides, the model provides a link between neutrino mass and dark matter phenomenology.
Article
Full-text available
AMS-02 was successfully installed on the International Space Station (ISS) in May 2011, to perform precise measurements of galactic cosmic rays in the 100 MV to few TV magnetic rigidity range. Among several specialized sub-detectors, AMS-02 includes a Ring Imaging Cherenkov detector (RICH), which provides a precise measurement of the particle charge and velocity. The Cherenkov light is produced in a radiator made of silica aerogel and sodium fluoride and collected by means of an array of photomultiplier tubes. Since its launch to space, the detector has been taking data without failures; its Functionality and data integrity are monitored and show stable response. In order to achieve the optimal detector performance, calibrations have been performed to account for the dependence of the photocletectors response on temperature and for effective non-uniformities in the detector. The knowledge gathered of the photon yield at the percent level resulted in a charge resolution of 0.3 charge units for He and 0.5 charge units for Si ions. The required precision in the measurements of the particle velocity at the per mil level demanded a more accurate determination of the aerogel refractive index. A map of the aerogel radiator refractive index has been directly inferred from in-flight high statistics data with a precision of Delta n/n < 2 x 10(-5) on average and its stability. with time has also been checked. Finally, a velocity resolution of similar to 0.8 x 10(-3) for He and similar to 0.5 x 10(-3) for Z > 5 ions has been obtained.
Article
Full-text available
A precision measurement by AMS of the positron fraction in primary cosmic rays in the energy range from 0.5 to 500 GeV based on 10.9 million positron and electron events is presented. This measurement extends the energy range of our previous observation and increases its precision. The new results show, for the first time, that above ∼200 GeV the positron fraction no longer exhibits an increase with energy.
Article
Full-text available
Precision measurements by the Alpha Magnetic Spectrometer on the International Space Station of the primary cosmic-ray electron flux in the range 0.5 to 700 GeV and the positron flux in the range 0.5 to 500 GeV are presented. The electron flux and the positron flux each require a description beyond a single power-law spectrum. Both the electron flux and the positron flux change their behavior at ∼30 GeV but the fluxes are significantly different in their magnitude and energy dependence. Between 20 and 200 GeV the positron spectral index is significantly harder than the electron spectral index. The determination of the differing behavior of the spectral indices versus energy is a new observation and provides important information on the origins of cosmic-ray electrons and positrons.
Article
The first measurement on the antiproton to proton ratio made by the AMS-02 collaboration agrees with the expection from conventional cosmic-ray secondaries in the kinetic energy range 10100\sim 10-100 GeV, which can be turned into stringent upper limits on the dark matter (DM) annihilation cross sections above 300\sim 300 GeV. Using the GALPROP code, we derive the upper limits in various propagation models and DM profiles. We show that in the "conventional" propagation model, for the qqˉq\bar q, bbˉb\bar b, and WW final states, the constraints can be more stringent than that derived from the recent Ferm-LAT gamma-ray data on the dwarf spheroidal satellite galaxies. Making use of the typical minimal, median and maximal models obtained from a previous global fit, we show that the variation of the upper limits is around a factor of five.
Article
This volume presents in detail the fundamental theories of linear regression analysis and diagnosis, as well as the relevant statistical computing techniques so that readers are able to actually model the data using the methods and techniques described in the book. It covers the fundamental theories in linear regression analysis and is extremely useful for future research in this area. The examples of regression analysis using the Statistical Application System (SAS) are also included. This book is suitable for graduate students who are either majoring in statistics/biostatistics or using linear regression analysis substantially in their subject fields. © 2009 by World Scientific Publishing Co. Pte. Ltd. All rights reserved.