• Sonuç bulunamadı

Principal-Component Analysis Of Two-Particle Azimuthal Correlations İn PbPb And Ppb Collisions At CMS

N/A
N/A
Protected

Academic year: 2021

Share "Principal-Component Analysis Of Two-Particle Azimuthal Correlations İn PbPb And Ppb Collisions At CMS"

Copied!
40
0
0

Yükleniyor.... (view fulltext now)

Tam metin

(1)

CMS-HIN-15-010

Principal-component analysis of two-particle azimuthal

correlations in PbPb and pPb collisions at CMS

The CMS Collaboration

Abstract

For the first time a principle-component analysis is used to separate out different or-thogonal modes of the two-particle correlation matrix from heavy ion collisions. The analysis uses data from √sNN = 2.76 TeV PbPb and√sNN = 5.02 TeV pPb collisions collected by the CMS experiment at the LHC. Two-particle azimuthal correlations have been extensively used to study hydrodynamic flow in heavy ion collisions. Re-cently it has been shown that the expected factorization of two-particle results into a product of the constituent single-particle anisotropies is broken. The new infor-mation provided by these modes may shed light on the breakdown of flow factor-ization in heavy ion collisions. The first two modes (“leading” and “subleading”) of two-particle correlations are presented for elliptical and triangular anisotropies in PbPb and pPb collisions as a function of pT over a wide range of event activity. The

leading mode is found to be essentially equivalent to the anisotropy harmonic pre-viously extracted from two-particle correlation methods. The subleading mode rep-resents a new experimental observable and is shown to account for a large fraction of the factorization breaking recently observed at high transverse momentum. The principle-component analysis technique has also been applied to multiplicity fluctu-ations. These also show a subleading mode. The connection of these new results to previous studies of factorization is discussed.

Published in Physical Review C as doi:10.1103/PhysRevC.96.064902.

c

2017 CERN for the benefit of the CMS Collaboration. CC-BY-4.0 license

See Appendix A for the list of collaboration members

(2)
(3)

lider (RHIC) indicate that a strongly interacting QGP is produced in heavy ion collisions [1–4]. The presence of azimuthal anisotropy in the emission of final state hadrons revealed a strong collective flow behavior of this strongly coupled hot and dense medium [5, 6]. The significantly higher energies available at the CERN LHC compared to RHIC have allowed the ALICE, AT-LAS, and CMS experiments to make very detailed measurements of the QGP properties [7–15]. The collective expansion of the QGP can be described by hydrodynamic flow models [16–18]. In the context of these models, the azimuthal anisotropy of hadron emission is the response to the initial density profile of the overlap region of the colliding nuclei. Such anisotropic emis-sion, for a given event, can be quantified through a Fourier decomposition of the single-particle distribution dN dp = ∞

n=−∞ Vn(p)e−inφ, (1)

with Vn(p) = vn(p)einΨn(p), dppp = dpTdφ dη, and p being a shorthand notation for pT and

η. This single-particle distribution is the invariant yield of emitted particles N expressed in phase space pT, η and φ, i.e., transverse momentum, pseudorapidity, and azimuthal angle.

Here, vncorresponds to the real single-particle anisotropy andΨn(p)represents the nth order

event plane angle. Also, because of the reflection symmetry of the overlap region, the relation Vn∗ = V−nholds for the complex harmonics. Using this relation and integrating Eq. (1) over a

given pseudorapidity and pTwindow yields

dN = N  1+2 ∞

n=1 vn(p)cos[n(φ−Ψn(p))]  . (2)

Note that the single-particle anisotropy coefficient vnis generally a function of pTand η, which

is also the case for the event plane angle. The azimuthal correlation of Npairs emitted particle pairs (with particles labeled a and b) as a function of their azimuthal separation∆φab =φa−φb can be characterized by its own Fourier harmonics

dNpairs d∆φab = Npairs  1+2 ∞

n=1 Vn∆(pa, pb)cos(n∆φ)  , (3)

where Vn∆ is the two-particle harmonic. In a pure hydrodynamic picture, as a consequence

of independent particle emission, the flow hypothesis connects the single- and two-particle spatial anisotropies from Eqs. (2) and (3) through factorization. In other words, particles carry information only about their orientation with respect to the whole system and the two-particle distribution can therefore be factorized based on

hdN pairs d∆φab i = h dN a dN bi, (4)

with the bracket hirepresenting the average over all events of interest. This equality can be investigated by looking at the connection between the single- and two-particle harmonics

(4)

From Eq. (5) we infer that factorization is preserved when the cosine value equals unity. This scenario is possible only when the event plane angle acts as a global phase, lacking any pT

or η dependence for a given event. Thus, measurements of the momentum space fluctuations (correlations) constrain the initial state and properties of QGP expansion dynamics. Previous measurements have shown a significant breakdown of factorization at high pT in ultracentral

(i.e., almost head-on) PbPb collisions [15]. A smaller effect was also seen in high-multiplicity pPb collisions [19]. Furthermore, significant factorization breakdown effects as a function of η were observed in both PbPb and high-multiplicity pPb collisions [19]. Several possible explana-tions for the observed factorization breaking have been proposed. One expected contribution arises from nonflow effects, i.e. short-range correlations mainly due to jet fragmentation and resonance decays. However, factorization breaking is also possible in hydrodynamic models, once the effects of event-by-event initial-state fluctuations are taken into account [20, 21]. Such a nonuniform initial-state energy density can arise from fluctuations in the positions of nucle-ons within nuclei and/or the positinucle-ons of quark and gluon cnucle-onstituents inside each nucleon, giving rise to variations in the collision points when the two nuclei collide. The resulting fluc-tuating initial energy density profile creates nonuniformities in pressure gradients which push particles in different regions of phase space in directions that vary randomly about a mean an-gle, thereby imprinting these fluctuations on the final particle distributions. Consequently, the event plane angles estimated from particles in different pTand η ranges may vary with respect

to each other. By introducing such a dependence,Ψn= Ψn(pT, η), it is possible to describe the

resulting final-state particle distributions using hydrodynamical models [20, 21].

Principal-component analysis (PCA) is a multivariate technique that can separate out the dif-ferent orthogonal contributions (also known as modes) to the fluctuations. Using the method introduced in Ref. [22], this paper presents the first experimental use of applying PCA to two-particle correlations in order to study factorization breaking as a function of pT. This allows the

extraction of a new experimental observable, the subleading mode, which is directly connected to initial-state fluctuations and their effect on factorization breaking.

2

Experimental setup and data samples

The Compact Muon Solenoid (CMS) is an axially symmetric detector with an onion-like struc-ture, which consists of several subsystems concentrically placed around the interaction point. The CMS magnet is a superconducting solenoid providing a magnetic field of 3.8 T, which al-lows precise measurement of charged particle momentum. The muon chambers are placed outside the solenoid. In this analysis the data used is extracted from the silicon tracker, which is the closest subdetector to the interaction point. This detector consists of 1440 silicon pixel and 15 148 silicon strip detector modules that detect hit locations, from which the charged par-ticle trajectories are reconstructed. The silicon tracker covers charged parpar-ticles within the range

|η| < 2.5, and provides an impact parameter resolution of ∼15 µm and a pT resolution better

than 1.5% up to pT∼100 GeV/c.

The other two subdetectors located inside the solenoid, are the electromagnetic calorimeter (ECAL) and hadronic calorimeter (HCAL). The ECAL is constructed of 75 848 lead-tungstate crystals which are arranged in a quasi-projective geometry and cover a pseudorapidity range of |η| < 1.48 units in the barrel and two endcaps that extend|η|up to 3.0. The HCAL barrel and endcaps are sampling calorimeters constructed from brass and scintillator plates, covering

|η| < 3.0. Additional extension in|η|from 2.9 up to 5.2 is achieved with the iron and quartz-fiber ˇCerenkov Hadron Forward (HF) calorimeters on either side of the interaction region. The HF calorimeters are segmented into towers, each of which is a two-dimensional cell with a

(5)

Ref. [24].

This analysis is performed using data recorded by the CMS experiment during the LHC heavy ion runs in 2011 and 2013. The PbPb data set at a center-of-mass energy of√sNN = 2.76 TeV corresponds to an integrated luminosity of about 159 µb−1, while the pPb data set at √sNN = 5.02 TeV corresponds to about 35 nb−1. During the pPb run, the beam energies were 4 TeV for protons and 1.58 TeV per nucleon for lead nuclei.

3

Selection of events and tracks

Online triggers, track reconstruction, and offline event selections are the same as in Refs. [15, 19, 25] for PbPb and pPb data samples, and are summarized in the following sections.

3.1 The PbPb data

Minimum bias PbPb events were collected using coincident trigger signals from both ends of the detector in either BSCs or the HF calorimeters. Events affected by cosmic rays, detector noise, out-of-time triggers, and beam backgrounds were suppressed by requiring a coincidence of the minimum bias trigger with bunches colliding in the interaction region. The efficiency of the trigger is more than 97% in case of hadronic inelastic PbPb collisions. Because of hardware limits on the data acquisition rate, only a small fraction (2%) of all minimum bias events were recorded (i.e., the trigger is “prescaled”). To enhance the event sample for very central PbPb collisions, a dedicated online trigger was implemented by simultaneously requiring the HF transverse energy (ET) sum to be greater than 3260 GeV and the pixel cluster multiplicity to be

greater than 51400 (which approximately corresponds to 9500 charged particles over 5 units of η). The selected events correspond to the 0–0.2% most central PbPb collisions. Other standard PbPb centrality classes presented in this paper were determined based on the total energy de-posited in the HF calorimeters [13]. The inefficiencies of the minimum bias trigger and event selection for very peripheral events are taken into account.

In order to reduce further the background from single-beam interactions (e.g., beam gas and beam halo), cosmic muons, and ultraperipheral collisions leading to the electromagnetic breakup of one or both Pb nuclei [26], offline PbPb event selection criteria [13] were applied by requir-ing energy deposits in at least three towers in each of the HF calorimeters, with at least 3 GeV of energy in each tower, and the presence of a reconstructed primary vertex built of at least two tracks. The reconstructed primary vertex is required to be located within±15 cm of the average interaction point along the beam axis and within a radius of 0.2 cm in the transverse plane. Following the procedure developed in Ref. [15], events with large signals in both ZDCs and HFs are identified as having at least one additional interaction, or pileup events, and are thus rejected (about 0.1% of all events).

The reconstruction of the primary event vertex and of the trajectories of charged particles in PbPb collisions is based on signals in the silicon pixel and strip detectors and is described in detail in Ref. [13]. From studies based on PbPb events simulated usingHYDJET v1.8 [27], the combined geometrical acceptance and reconstruction efficiency of the primary tracks is about

(6)

70% at pT ∼1 GeV/c and|η| <1.0 for the most central (0–5%) PbPb events, but drops to about 50% for pT ∼0.3 GeV/c. The fraction of misidentified tracks is kept to be<5% over most of the

pT (>0.5 GeV/c) and |η|(<1.6) ranges. It increases to about 20% for very low pT (<0.5 GeV/c)

particles in the forward (|η| ≥2.0) region.

3.2 The pPb data

Minimum bias pPb events were triggered by requiring at least one track with pT > 0.4 GeV/c

to be found in the pixel tracker in coincidence with an LHC pPb bunch crossing. From all minimum bias triggered events, only a fraction of (∼10−3) was recorded. In order to select high-multiplicity pPb collisions, a dedicated trigger was implemented using the CMS level-1 (L1) and high-level trigger (HLT) systems. At L1, the total transverse energy summed over the ECAL and HCAL is required to be greater than a given threshold (20 or 40 GeV). The online track reconstruction for the HLT is based on the three layers of pixel detectors, and requires a track originated within a cylindrical region of length 30 cm along the beam and radius of 0.2 cm perpendicular to the beam. For each event, the vertex reconstructed with the highest number of pixel tracks is selected. The number of pixel tracks (Ntrkonline) with|η| < 2.4, pT > 0.4 GeV/c,

and having a distance of closest approach of 0.4 cm or less to this vertex, is determined for each event.

In the offline analysis, hadronic pPb collisions are selected by requiring a coincidence of at least one HF calorimeter tower with more than 3 GeV of total energy in each of the HF detectors. Events are also required to contain at least one reconstructed primary vertex within 15 cm of the nominal interaction point along the beam axis and within 0.15 cm transverse to the beam trajectory. At least two reconstructed tracks are required to be associated with the primary vertex. Beam-related background is suppressed by rejecting events for which less than 25% of all reconstructed tracks are of good quality (i.e., the tracks selected for physics analysis). The instantaneous luminosity provided by the LHC in the 2013 pPb run resulted in approxi-mately 3% probability of at least one additional interaction occurring in the same bunch cross-ing, i.e. pileup events. Pileup was rejected using a procedure based on the number of tracks in a given vertex and the distance between that an additional vertex (see Ref. [25]). The frac-tion of pPb events selected by these criteria, which have at least one particle (proper lifetime τ > 10−18s) with total energy E > 3 GeV in η range of−5 < η < −3 and at least one in the range 3< η<5 (selection referred to as “double-sided”) has been found to be 97–98% by using theEPOS[28] andHIJING[29] event generators.

In this analysis, the CMS highPurity [30] tracks are used. Additionally, a reconstructed track is only considered as a primary-track candidate if the significance of the separation along the beam axis (z) between the track and the best vertex, dz/σ(dz), and the significance of the impact parameter relative to the best vertex transverse to the beam, dT(dT), are less than 3 in each

case. The relative uncertainty of the pT measurement, σ(pT)/pT, is required to be less than

10%. To ensure high tracking efficiency and to reduce the rate of misidentified tracks, only tracks within|η| < 2.4 and with pT > 0.3 GeV/c are used in the analysis. The entire pPb data

set is divided into classes of reconstructed track multiplicity, Ntrkoffline, where primary tracks with|η| <2.4 and pT >0.4 GeV/c are counted. The multiplicity classification in this analysis is

(7)

h

d∆φ i = 1+n

=1υ

2

n{2}cos(n∆φ) , (6)

where υn{2}is the integrated reference flow calculated from the Vn∆as

υn{2} = q Vn∆(pref T , prefT ) q V0∆(prefT , prefT ) , (7) with, Vn∆(prefT , prefT ) ≡ h

i∈ref cos(n∆φi)i. (8)

Here, the label V0∆ for Npairs is used, since the sum over cosine counts the number of pairs for

the n=0 case. Calculating the differential flow one gets υn(pT){2}υn{2} = Vn∆(pT, prefT ) V0∆(pT, prefT ) , (9) or, υn(pT) = Vn∆(pT, prefT ) q Vn∆(prefT , prefT ) q V0∆(prefT , prefT ) V0∆(pT, prefT ) . (10)

The single-particle anisotropy definition in Eq. (10) includes the V0∆ terms to compensate for the fact that the Vn∆Fourier harmonics are calculated without per-event normalization by the number of pairs in the given bin [15, 19]. This way of calculating the cosine term is essential for the PCA to work, since it gives a weight to a bin that is of the order of the number of particles in it [22].

In a realistic experiment, the Vn∆ harmonics of Eq. (8) are affected by imperfections in the de-tector and take the following operational definition

Vn∆(paT, pbT) = hcos(n∆φ)iS− hcos(n∆φ)iB, n=1, 2, 3, . . . . (11)

Here, the first term in the right-hand side of Eq. (11),hcos(n∆φ)iS, is the two-particle anisotropic signal where the correlated particles belong to the same event. The second term,hcos(n∆φ)iB

is a background term that accounts for the nonuniform acceptance of the detector. This term is usually two orders of magnitude smaller than the corresponding signal. It is estimated by mixing particle tracks from two random events. These two events have the same 2 cm wide range of the primary vertex position in the z direction and belong to the same centrality (track multiplicity) class. For both terms, in order to suppress nonflow correlations, a pseudorapidity difference requirement between the two tracks|∆η| >2 is applied.

(8)

4.1 Factorization breaking

The PCA is a multivariate analysis that orders the fluctuations in the data by size. The or-dering is done through principal components that represent orthogonal eigenvectors of the corresponding covariance data matrix. In the context of flow fluctuations, the components should reveal any significant substructure caused by the fluctuating initial state geometry of colliding nuclei. Introducing PCA in terms of factorization breaking one can write the Pearson correlation coefficient used for measurement of the effect as in Ref. [19]

rn(paT, pbT) ≡

Vn∆(paT, pbT)

q

Vn∆(paT, pTa)Vn∆(pbT, pbT)

≈ hcos n(Ψ(paT) −Ψ(pTb))i. (12)

The ratio rnis approximated by the cosine term, giving unity if the event plane angle is a global

phase, as discussed previously. Expressing the ratio through the two-particle harmonic in com-plex form from Eq. (5), rncan only be unity if the complex flow coefficient Vn(pT)is generated

from one initial geometry. For instance, where the initial geometry of the overlap region is de-fined by some complex eccentricity (εn) and a fixed real function f(pT), i.e., Vn(pT) = f(pT)εn.

However, if events are described by multiple eccentricities then rnmay be less than unity and

the flow pattern displays factorization breaking [31]. This last statement can be generalized by expanding the complex flow coefficient using the principal components (Vn(1)(pT), V

(2)

n (pT), . . .)

as a basis built from a covariance data matrix of given size Nα×Nα

Vn(pT) =ξ(n1)V (1) n (pT) +ξ(n2)V (2) n (pT) + · · · +ξn(Nα)V (Nα) n (pT), (13)

where ξn(i)are complex uncorrelated variables with zero mean i.e.hξn(i)ξn(j)i =δij,hξ(ni)i =0, and

Nαrepresents the number of pTdifferential bins. Therefore, the two-particle harmonics are the

building elements of the covariance data matrix[Vˆn∆(paT, pbT)]Nα×Nα.

A covariance matrix is symmetrical and positive semidefinite (i.e., with eigenvalues λ ≥0). For the flow matrix, the last trait is valid if there are no nonflow contributions and no strong statistical fluctuations [22]. Now, calculating the two-particle harmonic using the expansion from Eq. (13) one gets

Vn∆(paT, paT) = Nα

α=1 Vn(α)(paT)V (α) n (pbT). (14)

Here, the principal components will be referred to as modes [22, 31, 32]. In order to calculate the modes the spectral decomposition is rewritten as:

Vn∆(paT, pbT) =

α λ(α)e(α)(pTa)e(α)(pTb), (15) which gives: Vn(α)(pT) = p λ(α)e(α)(pT), (16) where e(α)(p

T)are (α)index values of normalized eigenvectors and λ(α) eigenvalues that are sorted in a strict decreasing order λ(1) > λ(2) > · · · > λ(n). Eq. (14) shows directly that factorization holds only in the case where just one mode is present. If multiple modes are present in the data, Eqs. (15) and (16) allow to define a normalized orthogonal basis for the total vngiven in Eq. (10). These basis vectors are defined by:

v(α) n (pT) ≡

Vn(α)(pT)

V0(1)(pT)

(9)

by multiplying the Vn∆(pT, pT)with: ζ = DV|η|<2.4 0 (paT, pbT) V0|∆η|>2(pa T, pbT) E , (18)

ζ being the mean value of the ratio of number of all pairs and number of pairs after applying |∆η| > 2 selection for the given bins. If the η distribution of particles did not depend upon pT then ζ(paT, pbT) would be constant for all values of paT and pbT. In fact, ζ does have a slight

dependence on pTa and pbT with a maximum at low values of paT and pbT. As events become more central the center of gravity of zeta moves to higher pTvalues. Finally, after applying this

correction the eigenvalue problem is solved with new matrix elements ˜

Vn∆(paT, pbT)≡ζVn∆(paT, pbT). (19) Eq. (17) then becomes:

v(nα)(pT) =

˜

Vn(α)(pT)

hM(pT)i

. (20)

The leading (α = 1) and the subleading (α = 2) normalized modes (for simplicity, term modes will be used) can be thought of as new experimental observables. Given that the eigenvalues λ(α)are strongly ordered, two components typically describe the variance in the harmonic flow to high accuracy. The leading mode is strongly correlated with the event plane, and thus is essentially equivalent to the standard definition of the single-particle anisotropic flow, while the subleading mode is uncorrelated with the event plane, and thus quantifies the magnitude of the factorization breaking caused by the initial-state fluctuations.

4.2 Multiplicity fluctuations

The PCA can also be applied for investigating multiplicity fluctuations in heavy ion collisions. The multiplicity matrix that is used for extraction of the corresponding modes is built from the following matrix elements:

[Mˆ(paT, pTb)]Nα×Nα = hV0∆(p

a

T, pbT)i − hM(pTa)ihM(pbT)i, (21)

where the term V0∆(pTa, pbT)represents the number of pairs for the given bins and M(pT)the

given bin multiplicity. Unlike in the flow cases n = 2, 3, here no pseudorapidity requirement

|∆η| > 2 is applied when correlating tracks. Using the multiplicity matrix the modes defined by Eq. (16) are derived and the leading and subleading modes are calculated with Eq. (20), ex-cluding the multiplication step in Eq. (19). The leading mode represents the “total multiplicity fluctuations”, i.e., if the higher modes are zero, then v(01)would approximately be equal to the standard deviation of multiplicity for the given pT bin. The reconstructed subleading mode

represents a new observable of the multiplicity spectrum. The multiplicity results in Section 6 represent exploratory studies and are for simplicity only presented for PbPb.

(10)

5

Systematic uncertainties

Several sources of possible systematic uncertainties, such as the event selection, the dimension of the matrix, and the effect of the tracking efficiency were investigated. Among these sources, only the effect of the tracking efficiency had a noticeable influence on the results. For all the considered cases n=0, 2, 3 the systematic uncertainties were estimated from the full difference between the final result with and without the correction for the tracking efficiency. Each recon-structed track was weighted by the inverse of the efficiency factor, εtrk(pT, η), which is a

func-tion of transverse momentum and pseudorapidity. The efficiency weighting factor accounts for the detector acceptance A(pT, η)and the reconstruction efficiency, E(pT, η)trk= A E).

From Eqs. (16) and (20) it can be seen that modes are functions of the eigenvectors and eigenval-ues, i.e. e and λ, of the matrix, and of the differential multiplicity M(pT). When the efficiency

correction is applied to each track, a completely new matrix is produced and the multiplic-ity of tracks also increases. The principal components of this new matrix were then calculated and new modes derived. This procedure gives a robust test of how susceptible the modes are to strong changes in (λ, e, M). Table 1 summarizes the uncertainties of the subleading mode in the highest bin 2.5< pT <3.0 GeV/c for both the pPb and PbPb cases. The systematic uncertainties

are estimated values and are rounded to the nearest integer. For the leading mode, systematic uncertainties are significant only for n = 0, while for the subleading mode systematic uncer-tainties are larger for all the cases n = 0, 2, 3. In the lower pT range, for the multiplicity case

n=0, the systematic uncertainties of the subleading mode are strongly correlated.

Table 1: Summary of estimated systematic uncertainties relative to the given mode for the last pT bin 2.5< pT <3.0 GeV/c for PbPb and pPb data.

PbPb n=2 n=3 n=0 Centrality (%) α=1 α=2 α=1 α=2 α=1 α=2 0–0.2 1% 30% 1% 40% 40% 10% 0–5 1% 50% 1% 40% 15% 10% 0–10 1% 30% 1% 40% 10% 30% 10–20 1% 10% 1% 40% 10% 20% 20–30 1% 10% 1% 20% 10% 15% 30–40 1% 10% 1% 35% 10% 10% 40–50 1% 10% 1% 25% 10% 10% 50–60 1% 7% 1% 30% 10% 30% pPb n=2 n=3 Noffline trk α=1 α=2 α=1 α=2 [220, 260) 1% 1.5% 1% 20% [185, 220) 1% 2.0% 1% 20% [150, 185) 1% 2.0% 1% 20% [120, 150) 1% 2.0% 1% 20%

6

Results

Figure 1 shows leading and subleading modes for the elliptic case (n = 2) for eight central-ity regions in PbPb collisions at√sNN = 2.76 TeV as a function of pT. These centrality regions

range from ultracentral 0–0.2% to peripheral 50–60%. The data are binned into seven pT bins

covering the region 0.3 < pT < 3.0 GeV/c. The number of differential pT bins for

(11)

1 2 3 2 ) α ( v 0 0.1 0.2 0-0.2% Centrality: CMS 1 2 3 0 0.1 0.2 0-5% =1 α =2 α 1 2 3 0 0.1 0.2 0-10% , CMS |>2.0} η ∆ {| 2 v , ALICE |>0.8} η ∆ {| 2 v 1 2 3 0 0.1 0.2 10-20% (2.76 TeV PbPb) -1 b µ 159 (GeV/c) T p 1 2 3 2 ) α ( v 0 0.1 0.2 20-30% (GeV/c) T p 1 2 3 0 0.1 0.2 30-40% (GeV/c) T p 1 2 3 0 0.1 0.2 40-50% (GeV/c) T p 1 2 3 0 0.1 0.2 50-60%

Figure 1: Leading (α=1) and subleading (α=2) modes for n=2 as a function of pT, measured

in a wide centrality range of PbPb collisions at √sNN = 2.76 TeV. The results for the leading mode (α = 1) are compared to the standard elliptic flow magnitude measured by ALICE and CMS using the two-particle correlation method taken from Refs. [7, 15], respectively. The error bars correspond to statistical uncertainties and boxes to systematic ones.

(12)

1 2 3 3 ) α ( v 0 0.05 0.1 0-0.2% Centrality: CMS 1 2 3 0 0.05 0.1 0-5% =1 α =2 α 1 2 3 0 0.05 0.1 0-10% , CMS |>2.0} η ∆ {| 3 v , ALICE |>0.8} η ∆ {| 3 v 1 2 3 0 0.05 0.1 10-20% (2.76 TeV PbPb) -1 b µ 159 (GeV/c) T p 1 2 3 3 ) α ( v 0 0.05 0.1 20-30% (GeV/c) T p 1 2 3 0 0.05 0.1 30-40% (GeV/c) T p 1 2 3 0 0.05 0.1 40-50% (GeV/c) T p 1 2 3 0 0.05 0.1 50-60%

Figure 2: Leading (α=1) and subleading (α=2) modes for n=3 as a function of pT, measured

in a wide centrality range of PbPb collisions at √sNN = 2.76 TeV. The results for the leading mode (α = 1) are compared to the standard triangular flow magnitude measured by ALICE and CMS using the two-particle correlation method taken from Refs. [7, 15], respectively. The error bars correspond to statistical uncertainties and boxes to systematic ones.

(13)

1

2

3

2

)

α

(

v

0

0.1

0.2

0.3

< 260

trk

offline

N

220

CMS

1

2

3

0

0.1

0.2

0.3

< 220

trk

offline

N

185

(5.02 TeV pPb)

-1

35 nb

, CMS

|>2}

η

{2, |

2

v

=1

α

=2

α

(GeV/c)

T

p

1

2

3

2

)

α

(

v

0

0.1

0.2

0.3

< 185

trk

offline

N

150

(GeV/c)

T

p

1

2

3

0

0.1

0.2

0.3

< 150

trk

offline

N

120

Figure 3: Leading (α = 1) and subleading (α = 2) modes for n = 2 as a function of pT,

mea-sured in high-multiplicity pPb collisions at√sNN = 5.02 TeV, for four classes of reconstructed track multiplicity Ntrkoffline. The results for the leading mode (α = 1) are compared to the stan-dard elliptic flow magnitude taken from Ref. [25]. The error bars correspond to statistical uncertainties and boxes to systematic ones.

(14)

1

2

3

3

)

α

(

v

0

0.05

0.1

< 260

trk

offline

N

220

CMS

1

2

3

0

0.05

0.1

< 220

trk

offline

N

185

(5.02 TeV pPb)

-1

35 nb

, CMS

|>2}

η

{2, |

3

v

=1

α

=2

α

(GeV/c)

T

p

1

2

3

3

)

α

(

v

0

0.05

0.1

< 185

trk

offline

N

150

(GeV/c)

T

p

1

2

3

0

0.05

0.1

< 150

trk

offline

N

120

Figure 4: Leading (α = 1) and subleading (α = 2) modes for n = 3 as a function of pT,

mea-sured in high-multiplicity pPb collisions at√sNN = 5.02 TeV, for four classes of reconstructed track multiplicity Ntrkoffline. The results for the leading mode (α = 1) are compared to the stan-dard triangular flow magnitude taken from Ref. [25]. The error bars correspond to statistical uncertainties and boxes to systematic ones.

(15)

0 1 2

2

r

0.8 0.9 1 0-0.2% Centrality: CMS 0 1 2 0.8 0.9 1 0-5% 0 1 2 0.8 0.9 1 10-20% (2.76 TeV PbPb) -1 b µ 159 < 3.0 GeV/c T a 2.5 < p 2

r

=2 α PCA N

(GeV/c)

T b

- p

T a

p

0 1 2

2

r

0.8 0.9 1 20-30%

(GeV/c)

T b

- p

T a

p

0 1 2 0.8 0.9 1 30-40%

(GeV/c)

T b

- p

T a

p

0 1 2 0.8 0.9 1 40-50%

Figure 5: Comparison of the Pearson correlation coefficient r2 reconstructed with harmonic

decomposition using the leading and subleading modes and r2values from Ref. [19], as a

func-tion of pa

T−pbTin bin of pTa for six centrality classes in PbPb collisions at

sNN =2.76 TeV. The error bars correspond to statistical uncertainties and boxes to systematic ones.

(16)

0 1 2

3

r

0.85 0.9 0.95 1 1.05 0-0.2% Centrality: CMS 0 1 2 0.85 0.9 0.95 1 1.05 0-5% 3

r

=2 α PCA N 0 1 2 0.85 0.9 0.95 1 1.05 10-20% (2.76 TeV PbPb) -1 b µ 159 < 3.0 GeV/c T a 2.5 < p

(GeV/c)

T b

- p

T a

p

0 1 2

3

r

0.85 0.9 0.95 1 1.05 20-30%

(GeV/c)

T b

- p

T a

p

0 1 2 0.85 0.9 0.95 1 1.05 30-40%

(GeV/c)

T b

- p

T a

p

0 1 2 0.85 0.9 0.95 1 1.05 40-50%

Figure 6: Comparison of the Pearson correlation coefficient r3 reconstructed with harmonic

decomposition using the leading and subleading modes and r3values from Ref. [19], as a

func-tion of pa

T−pbTin bin of pTa for six centrality classes in PbPb collisions at

sNN =2.76 TeV. The error bars correspond to statistical uncertainties and boxes to systematic ones.

(17)

tracks |<2.4 η |

N

2 10 103 2 (1)

v

2 (2)

v

0 0.2 0.4 0.6 0.8 -1

= 35 nb

int

= 5.02 TeV, L

NN

s

pPb

-1

b

µ

= 159

int

= 2.76 TeV, L

NN

s

PbPb

PbPb centrality(%)

CMS

< 3.0 GeV/c

T

2.5 < p

0.1 2.5 7.5 15.0 25.0 35.0 45.0 55.0 tracks |<2.4 η |

N

2 10 103 3 (1)

v

3 (2)

v

0 0.2 0.4 0.6 0.8

Figure 7: The ratio between values of the subleading and leading modes, taken for the highest pT bin, as a function of centrality and of charged-particle multiplicity at midrapidity (double

axis). The PCA flow results for PbPb collisions at√sNN =2.76 TeV (filled blue squares) and for pPb collisions at√sNN = 5.02 TeV (filled red circles). The error bars correspond to statistical uncertainties and boxes to systematic ones.

(18)

1 2 3 0 ) α ( v 0 0.1 0.2 0.3 0-0.2% centrality CMS 1 2 3 0 0.1 0.2 0.3 0-5% 1 2 3 0 0.1 0.2 0.3 0-10% =1 α =2 α 1 2 3 0 0.1 0.2 0.3 10-20% (2.76 TeV PbPb) -1 b µ 159 (GeV/c) T p 1 2 3 0 ) α ( v 0 0.1 0.2 0.3 20-30% (GeV/c) T p 1 2 3 0 0.1 0.2 0.3 30-40% (GeV/c) T p 1 2 3 0 0.1 0.2 0.3 40-50% (GeV/c) T p 1 2 3 0 0.1 0.2 0.3 50-60%

Figure 8: Leading and subleading modes for n = 0, i.e. fluctuations in the total multiplicity, spanning eight centralities in PbPb collisions at√sNN =2.76 TeV. The error bars correspond to statistical uncertainties and boxes to systematic ones. The systematic uncertainties are strongly correlated bin-to-bin.

value within a given bin. For comparison, v(21) is plotted together with v2{2} from CMS for

ultracentral collisions [15] and from ALICE for midcentral collisions [7]. The leading mode, v(21), is dominant and is essentially equal to the single-particle anisotropy v2{2}extracted from

two-particle correlations. The subleading mode, v(22), is nonzero for all centrality classes and it tends to rise with pT. It has a small magnitude of about 0.02 for the highest pT bin and more

central collisions and then gradually increases up to 0.05 towards peripheral collisions.

Figure 2 shows leading and subleading modes for the triangular case (n = 3), using the same eight centrality classes in PbPb collisions at√sNN = 2.76 TeV. Similar to the n = 2 case, v(31) is plotted together with v3{2} from CMS for ultracentral collisions [15] and from ALICE for

midcentral collisions [7]. A very good agreement is found between v(31)and the standard v3{2}.

The subleading mode, v(32), is practically zero for ultracentral collisions but shows positive values for a range of centralities at high pT. From a hydrodynamical point of view the existence

of the subleading mode for n = 3 is the response to the first radial excitation of triangularity [32].

Figure 3 shows leading and subleading modes in the case of the elliptic harmonic (n = 2) in pPb collisions at√sNN = 5.02 TeV as a function of pT for four different classes of multiplicity.

The data are binned into six pT bins covering the region 0.3 < pT < 3.0 GeV/c. The number

of differential pT bins for constructing the covariance matrix is Nα = 6. As seen in PbPb

col-lisions, the leading mode is equal to standard v2{2}CMS results from Ref. [25]. Looking at

the subleading mode (α = 2) values close to zero are observed at low pT with a moderate

in-crease in magnitude towards high pT. For pTvalues close to 3.0 GeV/c the subleading mode υ

(2)

2

(19)

factoriza-bottom panel of Fig. 4 shows that v3 is close to zero for all values of pT. Quantitatively similar

behavior was seen for flow factorization breaking in Ref. [19]. Similarly to the elliptic case, the leading and subleading triangular modes are rather independent of multiplicity for pPb collisions.

The Pearson correlation coefficient defined in Eq. (12) measures the magnitude of factorization breaking. This coefficient depends upon the two-particle harmonics Vn∆that in turn are built

up from the complete set of modes as shown in Eq. (14). These harmonics are approximated by the sum of just the leading and subleading modes. The comparison between the values of the PCA r2 and of the r2 from Ref. [19] is shown in Fig. 5. Using only the leading and

subleading modes it is possible to reconstruct the shape of the r2. However r2is closer to unity

for the PCA results than for the previous measurements. This is expected because the Vn∆ values are constructed from only two of the modes. Figure 6 shows the n=3 case, again using the comparison with r3 from the previous two-particle correlation analysis [19]. Although the

errors are large it is clear that the principle-component analysis tracks the previously measured divergence of r3from unity at high pT.

The Pearson coefficient calculated from Eq. (12) can be expanded as a power series of ratios of modes. Figure 7 shows the ratio of the leading and subleading modes for both pPb and PbPb collisions as a function of centrality (track multiplicity). The ratios are calculated for the highest pT bin used in the analysis. The top panel shows the elliptic case while the bottom

panel shows the triangular case. For the elliptic case the ratio is clearly above zero, with pPb high-multiplicity values being above the peripheral PbPb ones. For the triangular case half of the individual points are consistent with zero within the uncertainties. However, the ensemble of all the points suggest that the ratio is above zero.

Finally, Fig. 8 shows leading and subleading modes for the multiplicity case (n = 0) for PbPb collisions as a function of pTfor eight regions of centrality. For all centralities the leading mode

depends only weakly on pT, while the subleading mode increases rapidly with pT except for

very central collisions. The observed increase of the subleading mode with pT for all

centrali-ties is a response to radial-flow fluctuations [22, 33]. From a hydrodynamical point of view, the number of particles at high-pT decreases exponentially as exp[pT(u−u0)/T]. Here, T is the

temperature, u is the maximum fluid velocity, and u0 =

1+u2. A small variation in u

pro-duces a relative yield that increases linearly with pT. Such behaviour is observed in the data for

more peripheral collisions. At a given pT the subleading mode increases strongly from central

to peripheral collisions. Since peripheral collisions correspond to smaller interaction volumes, it is expected that pT fluctuations are more important for peripheral than for central events.

7

Summary

For the first time the leading and subleading modes of elliptic and triangular flow have been measured for 5.02 TeV pPb and 2.76 TeV PbPb collisions. For PbPb collisions the leading and subleading modes of multiplicity fluctuations have also been measured. Since the principal component analysis uses all the information encoded in the covariance matrix, it provides

(20)

in-creased sensitivity to fluctuations. For a very wide range of pT and centrality, the leading

modes of the elliptic and triangular flow are found to be essentially equal to the anisotropy coefficients measured using the standard two-particle correlation method. For both the elliptic and triangular cases the subleading modes are non-zero and increase with pT. This behavior

reflects a breakdown of flow factorization at high pT in both the pPb and PbPb systems. For

charged-particle multiplicity both the leading and subleading modes increase steadily from central to peripheral PbPb events. The leading mode depends only weakly upon pT while the

subleading mode increases strongly with pT. This centrality and pT dependence is suggestive

of the presence of fluctuations in the radial flow.

In summary the subleading modes of the principal-component analysis capture new informa-tion from the spectra of flow and multiplicity fluctuainforma-tions and provide an efficient method to quantify the breakdown of factorization in two-particle correlations.

Acknowledgments

We congratulate our colleagues in the CERN accelerator departments for the excellent perfor-mance of the LHC and thank the technical and administrative staffs at CERN and at other CMS institutes for their contributions to the success of the CMS effort. In addition, we grate-fully acknowledge the computing centers and personnel of the Worldwide LHC Computing Grid for delivering so effectively the computing infrastructure essential to our analyses. Fi-nally, we acknowledge the enduring support for the construction and operation of the LHC and the CMS detector provided by the following funding agencies: BMWFW and FWF (Aus-tria); FNRS and FWO (Belgium); CNPq, CAPES, FAPERJ, and FAPESP (Brazil); MES (Bulgaria); CERN; CAS, MoST, and NSFC (China); COLCIENCIAS (Colombia); MSES and CSF (Croatia); RPF (Cyprus); SENESCYT (Ecuador); MoER, ERC IUT, and ERDF (Estonia); Academy of Fin-land, MEC, and HIP (Finland); CEA and CNRS/IN2P3 (France); BMBF, DFG, and HGF (Ger-many); GSRT (Greece); OTKA and NIH (Hungary); DAE and DST (India); IPM (Iran); SFI (Ireland); INFN (Italy); MSIP and NRF (Republic of Korea); LAS (Lithuania); MOE and UM (Malaysia); BUAP, CINVESTAV, CONACYT, LNS, SEP, and UASLP-FAI (Mexico); MBIE (New Zealand); PAEC (Pakistan); MSHE and NSC (Poland); FCT (Portugal); JINR (Dubna); MON, RosAtom, RAS, RFBR and RAEP (Russia); MESTD (Serbia); SEIDI, CPAN, PCTI and FEDER (Spain); Swiss Funding Agencies (Switzerland); MST (Taipei); ThEPCenter, IPST, STAR, and NSTDA (Thailand); TUBITAK and TAEK (Turkey); NASU and SFFR (Ukraine); STFC (United Kingdom); DOE and NSF (USA).

Individuals have received support from the Marie-Curie program and the European Research Council and Horizon 2020 Grant, contract No. 675440 (European Union); the Leventis Foun-dation; the A. P. Sloan FounFoun-dation; the Alexander von Humboldt FounFoun-dation; the Belgian Fed-eral Science Policy Office; the Fonds pour la Formation `a la Recherche dans l’Industrie et dans l’Agriculture (FRIA-Belgium); the Agentschap voor Innovatie door Wetenschap en Technologie (IWT-Belgium); the Ministry of Education, Youth and Sports (MEYS) of the Czech Republic; the Council of Science and Industrial Research, India; the HOMING PLUS program of the Foun-dation for Polish Science, cofinanced from European Union, Regional Development Fund, the Mobility Plus program of the Ministry of Science and Higher Education, the National Science Center (Poland), contracts Harmonia 2014/14/M/ST2/00428, Opus 2014/13/B/ST2/02543, 2014/15/B/ST2/03998, and 2015/19/B/ST2/02861, Sonata-bis 2012/07/E/ST2/01406; the National Priorities Research Program by Qatar National Research Fund; the Programa Clar´ın-COFUND del Principado de Asturias; the Thalis and Aristeia programs cofinanced by EU-ESF and the Greek NSRF; the Rachadapisek Sompot Fund for Postdoctoral Fellowship,

(21)
(22)

Chula-References

[1] BRAHMS Collaboration, “Quark gluon plasma and color glass condensate at RHIC? The perspective from the BRAHMS experiment”, Nucl. Phys. A 757 (2005) 1,

doi:10.1016/j.nuclphysa.2005.02.130, arXiv:nucl-ex/0410020.

[2] PHOBOS Collaboration, “The PHOBOS perspective on discoveries at RHIC”, Nucl. Phys. A 757 (2005) 28, doi:10.1016/j.nuclphysa.2005.03.084,

arXiv:nucl-ex/0410022.

[3] STAR Collaboration, “Experimental and theoretical challenges in the search for the Quark Gluon Plasma: The STAR Collaboration’s critical assessment of the evidence from RHIC collisions”, Nucl. Phys. A 757 (2005) 102,

doi:10.1016/j.nuclphysa.2005.03.085, arXiv:nucl-ex/0501009. [4] PHENIX Collaboration, “Formation of dense partonic matter in relativistic

nucleus-nucleus collisions at RHIC: Experimental evaluation by the PHENIX collaboration”, Nucl. Phys. A 757 (2005) 184,

doi:10.1016/j.nuclphysa.2005.03.086, arXiv:nucl-ex/0410003.

[5] J.-Y. Ollitrault, “Anisotropy as a signature of transverse collective flow”, Phys. Rev. D 46 (1992) 2290, doi:10.1103/PhysRevD.46.229.

[6] P. F. Kolb, J. Sollfrank, and U. Heinz, “Anisotropic transverse flow and the quark hadron phase transition”, Phys. Rev. C 62 (2000) 054909,

doi:10.1103/PhysRevC.62.054909, arXiv:hep-ph/0006129.

[7] ALICE Collaboration, “Harmonic decomposition of two-particle angular correlations in Pb-Pb collisions at√sNN= 2.76 TeV”, Phys. Lett. B 708 (2012) 249,

doi:10.1016/j.physletb.2012.01.060, arXiv:1109.2501.

[8] ALICE Collaboration, “Elliptic flow of identified hadrons in Pb-Pb collisions at√sNN= 2.76 TeV”, JHEP 06 (2015) 190, doi:10.1007/JHEP06(2015)190,

arXiv:1405.4632.

[9] ATLAS Collaboration, “Measurement of the azimuthal anisotropy for charged particle production in√sNN= 2.76 TeV lead-lead collisions with the ATLAS detector”, Phys. Rev. C 86 (2012) 014907, doi:10.1103/PhysRevC.86.014907, arXiv:1203.3087. [10] ATLAS Collaboration, “Measurement of the distributions of event-by-event flow

harmonics in lead-lead collisions at√sNN= 2.76 TeV with the ATLAS detector at the LHC”, JHEP 11 (2013) 183, doi:10.1007/JHEP11(2013)183, arXiv:1305.2942. [11] ATLAS Collaboration, “Measurement of event-plane correlations in√sNN= 2.76 TeV

lead-lead collisions with the ATLAS detector”, Phys. Rev. C 90 (2014) 024905, doi:10.1103/PhysRevC.90.024905, arXiv:1403.0489.

[12] CMS Collaboration, “Centrality dependence of dihadron correlations and azimuthal anisotropy harmonics in PbPb collisions at√sNN = 2.76 TeV”, Eur. Phys. J. C 72 (2012) 2012, doi:10.1140/epjc/s10052-012-2012-3, arXiv:1201.3158.

[13] CMS Collaboration, “Measurement of the elliptic anisotropy of charged particles produced in PbPb collisions at√sNN = 2.76 TeV”, Phys. Rev. C 87 (2013) 014902, doi:10.1103/PhysRevC.87.014902, arXiv:1204.1409.

(23)

arXiv:1312.1845.

[16] B. Schenke, S. Jeon, and C. Gale, “(3+1)D hydrodynamic simulation of relativistic heavy-ion collisions”, Phys. Rev. C 82 (2010) 014903,

doi:10.1103/PhysRevC.82.014903, arXiv:1004.1408.

[17] C. Shen et al., “The iEBE-VISHNU code package for relativistic heavy-ion collisions”, Comput. Phys. Commun. 199 (2016) 61, doi:10.1016/j.cpc.2015.08.039, arXiv:1409.8164.

[18] K. Dusling and D. Teaney, “Simulating elliptic flow with viscous hydrodynamics”, Phys. Rev. C 77 (2008) 034905, doi:10.1103/PhysRevC.77.034905, arXiv:0710.5932. [19] CMS Collaboration, “Evidence for transverse momentum and pseudorapidity dependent

event plane fluctuations in PbPb and pPb collisions”, Phys. Rev. C 92 (2015) 034911, doi:10.1103/PhysRevC.92.034911, arXiv:1503.01692.

[20] F. G. Gardim, F. Grassi, M. Luzum, and J.-Y. Ollitrault, “Breaking of factorization of two-particle correlations in hydrodynamics”, Phys. Rev. C 87 (2013) 031901,

doi:10.1103/PhysRevC.87.031901, arXiv:1211.0989.

[21] U. Heinz, Z. Qiu, and C. Shen, “Fluctuating flow angles and anisotropic flow

measurements”, Phys. Rev. C 87 (2013) 034913, doi:10.1103/PhysRevC.87.034913, arXiv:1302.3535.

[22] R. S. Bhalerao, J.-Y. Ollitrault, S. Pal, and D. Teaney, “Principal component analysis of event-by-event fluctuations”, Phys. Rev. Lett. 114 (2015) 152301,

doi:10.1103/PhysRevLett.114.152301, arXiv:1410.7739.

[23] O. A. Grachov et al., “Performance of the combined zero degree calorimeter for CMS”, in XIII Int. Conf. on Calorimetry in High Energy Physics (CALOR 2008), edited by Michele Livan. arXiv:0807.0785. J. Phys.: Conf. Series, 160 (2009) 012059.

doi:10.1088/1742-6596/160/1/012059.

[24] CMS Collaboration, “The CMS experiment at the CERN LHC”, JINST 3 (2008) S08004, doi:10.1088/1748-0221/3/08/S08004.

[25] CMS Collaboration, “Multiplicity and transverse momentum dependence of two- and four-particle correlations in pPb and PbPb collisions”, Phys. Lett. B 724 (2013) 213, doi:10.1016/j.physletb.2013.06.028, arXiv:1305.0609.

[26] Ø. Djuvsland and J. Nystrand, “Single and double photonuclear excitations in Pb+Pb collisions at√sNN = 2.76 TeV at the CERN Large Hadron Collider”, Phys. Rev. C 83 (2011) 041901, doi:10.1103/PhysRevC.83.041901, arXiv:1011.4908.

[27] I. P. Lokhtin and A. M. Snigirev, “A model of jet quenching in ultrarelativistic heavy ion collisions and high-pT hadron spectra at RHIC”, Eur. Phys. J. C 45 (2005) 211,

(24)

[28] S. Porteboeuf, T. Pierog, and K. Werner, “Producing hard processes regarding the complete event: The EPOS event generator”, (2011). arXiv:1006.2967.

[29] M. Gyulassy and X.-N. Wang, “HIJING 1.0: A Monte Carlo program for parton and particle production in high energy hadronic and nuclear collisions”, Comput. Phys. Commun. 83 (1994) 307, doi:10.1016/0010-4655(94)90057-4,

arXiv:nucl-th/9502021.

[30] CMS Collaboration, “Tracking and vertexing results from first collisions”, CMS Physics Analysis Summary CMS-PAS-TRK-10-001, 2010.

[31] A. Mazeliauskas and D. Teaney, “Fluctuations of harmonic and radial flow in heavy ion collisions with principal components”, Phys. Rev. C 93 (2016) 024913,

doi:10.1103/PhysRevC.93.024913, arXiv:1509.07492.

[32] A. Mazeliauskas and D. Teaney, “Subleading harmonic flows in hydrodynamic simulations of heavy ion collisions”, Phys. Rev. C 91 (2015) 044902,

doi:10.1103/PhysRevC.91.044902, arXiv:1501.03138.

[33] N. Borghini and J.-Y. Ollitrault, “Momentum spectra, anisotropic flow, and ideal fluids”, Phys. Lett. B 642 (2006) 227, doi:10.1016/j.physletb.2006.09.062,

(25)

M. Flechl, M. Friedl, R. Fr ¨uhwirth1, V.M. Ghete, J. Grossmann, J. Hrubec, M. Jeitler1, A. K ¨onig, N. Krammer, I. Kr¨atschmer, D. Liko, T. Madlener, I. Mikulec, E. Pree, D. Rabady, N. Rad, H. Rohringer, J. Schieck1, R. Sch ¨ofbeck, M. Spanring, D. Spitzbart, J. Strauss, W. Waltenberger, J. Wittmann, C.-E. Wulz1, M. Zarucki

Institute for Nuclear Problems, Minsk, Belarus

V. Chekhovsky, V. Mossolov, J. Suarez Gonzalez

Universiteit Antwerpen, Antwerpen, Belgium

E.A. De Wolf, X. Janssen, J. Lauwers, M. Van De Klundert, H. Van Haevermaet, P. Van Mechelen, N. Van Remortel, A. Van Spilbeeck

Vrije Universiteit Brussel, Brussel, Belgium

S. Abu Zeid, F. Blekman, J. D’Hondt, I. De Bruyn, J. De Clercq, K. Deroover, G. Flouris, S. Lowette, S. Moortgat, L. Moreels, A. Olbrechts, Q. Python, K. Skovpen, S. Tavernier, W. Van Doninck, P. Van Mulders, I. Van Parijs

Universit´e Libre de Bruxelles, Bruxelles, Belgium

H. Brun, B. Clerbaux, G. De Lentdecker, H. Delannoy, G. Fasanella, L. Favart, R. Goldouzian, A. Grebenyuk, G. Karapostoli, T. Lenzi, J. Luetic, T. Maerschalk, A. Marinov, A. Randle-conde, T. Seva, C. Vander Velde, P. Vanlaer, D. Vannerom, R. Yonamine, F. Zenoni, F. Zhang2

Ghent University, Ghent, Belgium

A. Cimmino, T. Cornelis, D. Dobur, A. Fagot, M. Gul, I. Khvastunov, D. Poyraz, C. Roskas, S. Salva, M. Tytgat, W. Verbeke, N. Zaganidis

Universit´e Catholique de Louvain, Louvain-la-Neuve, Belgium

H. Bakhshiansohi, O. Bondu, S. Brochet, G. Bruno, A. Caudron, S. De Visscher, C. Delaere, M. Delcourt, B. Francois, A. Giammanco, A. Jafari, M. Komm, G. Krintiras, V. Lemaitre, A. Magitteri, A. Mertens, M. Musich, K. Piotrzkowski, L. Quertenmont, M. Vidal Marono, S. Wertz

Universit´e de Mons, Mons, Belgium

N. Beliy

Centro Brasileiro de Pesquisas Fisicas, Rio de Janeiro, Brazil

W.L. Ald´a J ´unior, F.L. Alves, G.A. Alves, L. Brito, M. Correa Martins Junior, C. Hensel, A. Moraes, M.E. Pol, P. Rebello Teles

Universidade do Estado do Rio de Janeiro, Rio de Janeiro, Brazil

E. Belchior Batista Das Chagas, W. Carvalho, J. Chinellato3, A. Cust ´odio, E.M. Da Costa,

G.G. Da Silveira4, D. De Jesus Damiao, S. Fonseca De Souza, L.M. Huertas Guativa, H. Malbouisson, M. Melo De Almeida, C. Mora Herrera, L. Mundim, H. Nogima, A. Santoro, A. Sznajder, E.J. Tonelli Manganote3, F. Torres Da Silva De Araujo, A. Vilela Pereira

Universidade Estadual Paulistaa, Universidade Federal do ABCb, S˜ao Paulo, Brazil

S. Ahujaa, C.A. Bernardesa, T.R. Fernandez Perez Tomeia, E.M. Gregoresb, P.G. Mercadanteb, C.S. Moona, S.F. Novaesa, Sandra S. Padulaa, D. Romero Abadb, J.C. Ruiz Vargasa

(26)

Institute for Nuclear Research and Nuclear Energy of Bulgaria Academy of Sciences

A. Aleksandrov, R. Hadjiiska, P. Iaydjiev, M. Misheva, M. Rodozov, S. Stoykova, G. Sultanov, M. Vutova

University of Sofia, Sofia, Bulgaria

A. Dimitrov, I. Glushkov, L. Litov, B. Pavlov, P. Petkov

Beihang University, Beijing, China

W. Fang5, X. Gao5

Institute of High Energy Physics, Beijing, China

M. Ahmad, J.G. Bian, G.M. Chen, H.S. Chen, M. Chen, Y. Chen, C.H. Jiang, D. Leggat, Z. Liu, F. Romeo, S.M. Shaheen, A. Spiezia, J. Tao, C. Wang, Z. Wang, E. Yazgan, H. Zhang, J. Zhao

State Key Laboratory of Nuclear Physics and Technology, Peking University, Beijing, China

Y. Ban, G. Chen, Q. Li, S. Liu, Y. Mao, S.J. Qian, D. Wang, Z. Xu

Universidad de Los Andes, Bogota, Colombia

C. Avila, A. Cabrera, L.F. Chaparro Sierra, C. Florez, C.F. Gonz´alez Hern´andez, J.D. Ruiz Alvarez

University of Split, Faculty of Electrical Engineering, Mechanical Engineering and Naval Architecture, Split, Croatia

B. Courbon, N. Godinovic, D. Lelas, I. Puljak, P.M. Ribeiro Cipriano, T. Sculac

University of Split, Faculty of Science, Split, Croatia

Z. Antunovic, M. Kovac

Institute Rudjer Boskovic, Zagreb, Croatia

V. Brigljevic, D. Ferencek, K. Kadija, B. Mesic, T. Susa

University of Cyprus, Nicosia, Cyprus

M.W. Ather, A. Attikis, G. Mavromanolakis, J. Mousa, C. Nicolaou, F. Ptochos, P.A. Razis, H. Rykaczewski

Charles University, Prague, Czech Republic

M. Finger6, M. Finger Jr.6

Universidad San Francisco de Quito, Quito, Ecuador

E. Carrera Jarrin

Academy of Scientific Research and Technology of the Arab Republic of Egypt, Egyptian Network of High Energy Physics, Cairo, Egypt

A.A. Abdelalim7,8, Y. Mohammed9, E. Salama10,11

National Institute of Chemical Physics and Biophysics, Tallinn, Estonia

R.K. Dewanjee, M. Kadastik, L. Perrini, M. Raidal, A. Tiko, C. Veelken

Department of Physics, University of Helsinki, Helsinki, Finland

P. Eerola, J. Pekkanen, M. Voutilainen

Helsinki Institute of Physics, Helsinki, Finland

J. H¨ark ¨onen, T. J¨arvinen, V. Karim¨aki, R. Kinnunen, T. Lamp´en, K. Lassila-Perini, S. Lehti, T. Lind´en, P. Luukka, E. Tuominen, J. Tuominiemi, E. Tuovinen

Lappeenranta University of Technology, Lappeenranta, Finland

(27)

A. Abdulsalam, I. Antropov, S. Baffioni, F. Beaudette, P. Busson, L. Cadamuro, C. Charlot, O. Davignon, R. Granier de Cassagnac, M. Jo, S. Lisniak, A. Lobanov, J. Martin Blanco, M. Nguyen, C. Ochando, G. Ortona, P. Paganini, P. Pigard, S. Regnard, R. Salerno, J.B. Sauvan, Y. Sirois, A.G. Stahl Leiton, T. Strebler, Y. Yilmaz, A. Zabi, A. Zghiche

Universit´e de Strasbourg, CNRS, IPHC UMR 7178, F-67000 Strasbourg, France

J.-L. Agram12, J. Andrea, D. Bloch, J.-M. Brom, M. Buttignol, E.C. Chabert, N. Chanon, C. Collard, E. Conte12, X. Coubez, J.-C. Fontaine12, D. Gel´e, U. Goerlach, M. Jansov´a, A.-C. Le Bihan, P. Van Hove

Centre de Calcul de l’Institut National de Physique Nucleaire et de Physique des Particules, CNRS/IN2P3, Villeurbanne, France

S. Gadrat

Universit´e de Lyon, Universit´e Claude Bernard Lyon 1, CNRS-IN2P3, Institut de Physique Nucl´eaire de Lyon, Villeurbanne, France

S. Beauceron, C. Bernet, G. Boudoul, R. Chierici, D. Contardo, P. Depasse, H. El Mamouni, J. Fay, L. Finco, S. Gascon, M. Gouzevitch, G. Grenier, B. Ille, F. Lagarde, I.B. Laktineh, M. Lethuillier, L. Mirabito, A.L. Pequegnot, S. Perries, A. Popov13, V. Sordini, M. Vander Donckt, S. Viret

Georgian Technical University, Tbilisi, Georgia

T. Toriashvili14

Tbilisi State University, Tbilisi, Georgia

Z. Tsamalaidze6

RWTH Aachen University, I. Physikalisches Institut, Aachen, Germany

C. Autermann, S. Beranek, L. Feld, M.K. Kiesel, K. Klein, M. Lipinski, M. Preuten, C. Schomakers, J. Schulz, T. Verlage

RWTH Aachen University, III. Physikalisches Institut A, Aachen, Germany

A. Albert, M. Brodski, E. Dietz-Laursonn, D. Duchardt, M. Endres, M. Erdmann, S. Erdweg, T. Esch, R. Fischer, A. G ¨uth, M. Hamer, T. Hebbeker, C. Heidemann, K. Hoepfner, S. Knutzen, M. Merschmeyer, A. Meyer, P. Millet, S. Mukherjee, M. Olschewski, K. Padeken, T. Pook, M. Radziej, H. Reithler, M. Rieger, F. Scheuch, D. Teyssier, S. Th ¨uer

RWTH Aachen University, III. Physikalisches Institut B, Aachen, Germany

G. Fl ¨ugge, B. Kargoll, T. Kress, A. K ¨unsken, J. Lingemann, T. M ¨uller, A. Nehrkorn, A. Nowack, C. Pistone, O. Pooth, A. Stahl15

Deutsches Elektronen-Synchrotron, Hamburg, Germany

M. Aldaya Martin, T. Arndt, C. Asawatangtrakuldee, K. Beernaert, O. Behnke, U. Behrens, A.A. Bin Anuar, K. Borras16, V. Botta, A. Campbell, P. Connor, C. Contreras-Campana,

F. Costanza, C. Diez Pardos, G. Eckerlin, D. Eckstein, T. Eichhorn, E. Eren, E. Gallo17, J. Garay Garcia, A. Geiser, A. Gizhko, J.M. Grados Luyando, A. Grohsjean, P. Gunnellini, A. Harb, J. Hauk, M. Hempel18, H. Jung, A. Kalogeropoulos, M. Kasemann, J. Keaveney, C. Kleinwort,

(28)

I. Korol, D. Kr ¨ucker, W. Lange, A. Lelek, T. Lenz, J. Leonard, K. Lipka, W. Lohmann18, R. Mankel, I.-A. Melzer-Pellmann, A.B. Meyer, G. Mittag, J. Mnich, A. Mussgiller, E. Ntomari, D. Pitzl, R. Placakyte, A. Raspereza, B. Roland, M. Savitskyi, P. Saxena, R. Shevchenko, S. Spannagel, N. Stefaniuk, G.P. Van Onsem, R. Walsh, Y. Wen, K. Wichmann, C. Wissing, O. Zenaiev

University of Hamburg, Hamburg, Germany

S. Bein, V. Blobel, M. Centis Vignali, A.R. Draeger, T. Dreyer, E. Garutti, D. Gonzalez, J. Haller, M. Hoffmann, A. Junkes, R. Klanner, R. Kogler, N. Kovalchuk, S. Kurz, T. Lapsien, I. Marchesini, D. Marconi, M. Meyer, M. Niedziela, D. Nowatschin, F. Pantaleo15, T. Peiffer, A. Perieanu, C. Scharf, P. Schleper, A. Schmidt, S. Schumann, J. Schwandt, J. Sonneveld, H. Stadie, G. Steinbr ¨uck, F.M. Stober, M. St ¨over, H. Tholen, D. Troendle, E. Usai, L. Vanelderen, A. Vanhoefer, B. Vormwald

Institut f ¨ur Experimentelle Kernphysik, Karlsruhe, Germany

M. Akbiyik, C. Barth, S. Baur, E. Butz, R. Caspart, T. Chwalek, F. Colombo, W. De Boer, A. Dierlamm, B. Freund, R. Friese, M. Giffels, A. Gilbert, D. Haitz, F. Hartmann15, S.M. Heindl, U. Husemann, F. Kassel15, S. Kudella, H. Mildner, M.U. Mozer, Th. M ¨uller, M. Plagge, G. Quast, K. Rabbertz, M. Schr ¨oder, I. Shvetsov, G. Sieber, H.J. Simonis, R. Ulrich, S. Wayand, M. Weber, T. Weiler, S. Williamson, C. W ¨ohrmann, R. Wolf

Institute of Nuclear and Particle Physics (INPP), NCSR Demokritos, Aghia Paraskevi, Greece

G. Anagnostou, G. Daskalakis, T. Geralis, V.A. Giakoumopoulou, A. Kyriakis, D. Loukas, I. Topsis-Giotis

National and Kapodistrian University of Athens, Athens, Greece

S. Kesisoglou, A. Panagiotou, N. Saoulidou

University of Io´annina, Io´annina, Greece

I. Evangelou, C. Foudas, P. Kokkas, N. Manthos, I. Papadopoulos, E. Paradas, J. Strologas, F.A. Triantis

MTA-ELTE Lend ¨ulet CMS Particle and Nuclear Physics Group, E ¨otv ¨os Lor´and University, Budapest, Hungary

M. Csanad, N. Filipovic, G. Pasztor

Wigner Research Centre for Physics, Budapest, Hungary

G. Bencze, C. Hajdu, D. Horvath19, F. Sikler, V. Veszpremi, G. Vesztergombi20, A.J. Zsigmond

Institute of Nuclear Research ATOMKI, Debrecen, Hungary

N. Beni, S. Czellar, J. Karancsi21, A. Makovec, J. Molnar, Z. Szillasi

Institute of Physics, University of Debrecen, Debrecen, Hungary

M. Bart ´ok20, P. Raics, Z.L. Trocsanyi, B. Ujvari

Indian Institute of Science (IISc), Bangalore, India

S. Choudhury, J.R. Komaragiri

National Institute of Science Education and Research, Bhubaneswar, India

S. Bahinipati22, S. Bhowmik, P. Mal, K. Mandal, A. Nayak23, D.K. Sahoo22, N. Sahoo, S.K. Swain

Panjab University, Chandigarh, India

S. Bansal, S.B. Beri, V. Bhatnagar, U. Bhawandeep, R. Chawla, N. Dhingra, A.K. Kalsi, A. Kaur, M. Kaur, R. Kumar, P. Kumari, A. Mehta, M. Mittal, J.B. Singh, G. Walia

(29)

Chowdhury, S. Sarkar, M. Sharan, S. Thakur

Indian Institute of Technology Madras, Madras, India

P.K. Behera

Bhabha Atomic Research Centre, Mumbai, India

R. Chudasama, D. Dutta, V. Jha, V. Kumar, A.K. Mohanty15, P.K. Netrakanti, L.M. Pant, P. Shukla, A. Topkar

Tata Institute of Fundamental Research-A, Mumbai, India

T. Aziz, S. Dugad, B. Mahakud, S. Mitra, G.B. Mohanty, B. Parida, N. Sur, B. Sutar

Tata Institute of Fundamental Research-B, Mumbai, India

S. Banerjee, S. Bhattacharya, S. Chatterjee, P. Das, M. Guchait, Sa. Jain, S. Kumar, M. Maity24, G. Majumder, K. Mazumdar, T. Sarkar24, N. Wickramage25

Indian Institute of Science Education and Research (IISER), Pune, India

S. Chauhan, S. Dube, V. Hegde, A. Kapoor, K. Kothekar, S. Pandey, A. Rane, S. Sharma

Institute for Research in Fundamental Sciences (IPM), Tehran, Iran

S. Chenarani26, E. Eskandari Tadavani, S.M. Etesami26, M. Khakzad, M. Mohammadi Najafabadi, M. Naseri, S. Paktinat Mehdiabadi27, F. Rezaei Hosseinabadi, B. Safarzadeh28, M. Zeinali

University College Dublin, Dublin, Ireland

M. Felcini, M. Grunewald

INFN Sezione di Baria, Universit`a di Barib, Politecnico di Baric, Bari, Italy

M. Abbresciaa,b, C. Calabriaa,b, C. Caputoa,b, A. Colaleoa, D. Creanzaa,c, L. Cristellaa,b, N. De Filippisa,c, M. De Palmaa,b, F. Erricoa,b, L. Fiorea, G. Iasellia,c, G. Maggia,c, M. Maggia, G. Minielloa,b, S. Mya,b, S. Nuzzoa,b, A. Pompilia,b, G. Pugliesea,c, R. Radognaa,b, A. Ranieria, G. Selvaggia,b, A. Sharmaa, L. Silvestrisa,15, R. Vendittia, P. Verwilligena

INFN Sezione di Bolognaa, Universit`a di Bolognab, Bologna, Italy

G. Abbiendia, C. Battilana, D. Bonacorsia,b, S. Braibant-Giacomellia,b, L. Brigliadoria,b,

R. Campaninia,b, P. Capiluppia,b, A. Castroa,b, F.R. Cavalloa, S.S. Chhibraa,b, G. Codispotia,b, M. Cuffiania,b, G.M. Dallavallea, F. Fabbria, A. Fanfania,b, D. Fasanellaa,b, P. Giacomellia, L. Guiduccia,b, S. Marcellinia, G. Masettia, F.L. Navarriaa,b, A. Perrottaa, A.M. Rossia,b, T. Rovellia,b, G.P. Sirolia,b, N. Tosia,b,15

INFN Sezione di Cataniaa, Universit`a di Cataniab, Catania, Italy

S. Albergoa,b, S. Costaa,b, A. Di Mattiaa, F. Giordanoa,b, R. Potenzaa,b, A. Tricomia,b, C. Tuvea,b

INFN Sezione di Firenzea, Universit`a di Firenzeb, Firenze, Italy

G. Barbaglia, K. Chatterjeea,b, V. Ciullia,b, C. Civininia, R. D’Alessandroa,b, E. Focardia,b, P. Lenzia,b, M. Meschinia, S. Paolettia, L. Russoa,29, G. Sguazzonia, D. Stroma, L. Viliania,b,15

INFN Laboratori Nazionali di Frascati, Frascati, Italy

(30)

INFN Sezione di Genovaa, Universit`a di Genovab, Genova, Italy

V. Calvellia,b, F. Ferroa, E. Robuttia, S. Tosia,b

INFN Sezione di Milano-Bicoccaa, Universit`a di Milano-Bicoccab, Milano, Italy

L. Brianzaa,b, F. Brivioa,b, V. Cirioloa,b, M.E. Dinardoa,b, S. Fiorendia,b, S. Gennaia, A. Ghezzia,b, P. Govonia,b, M. Malbertia,b, S. Malvezzia, R.A. Manzonia,b, D. Menascea, L. Moronia, M. Paganonia,b, K. Pauwelsa,b, D. Pedrinia, S. Pigazzinia,b,30, S. Ragazzia,b, T. Tabarelli de Fatisa,b

INFN Sezione di Napolia, Universit`a di Napoli ’Federico II’b, Napoli, Italy, Universit`a della Basilicatac, Potenza, Italy, Universit`a G. Marconid, Roma, Italy

S. Buontempoa, N. Cavalloa,c, S. Di Guidaa,d,15, F. Fabozzia,c, F. Fiengaa,b, A.O.M. Iorioa,b,

W.A. Khana, L. Listaa, S. Meolaa,d,15, P. Paoluccia,15, C. Sciaccaa,b, F. Thyssena

INFN Sezione di Padova a, Universit`a di Padova b, Padova, Italy, Universit`a di Trento c, Trento, Italy

P. Azzia,15, N. Bacchettaa, L. Benatoa,b, D. Biselloa,b, A. Bolettia,b, P. Checchiaa, M. Dall’Ossoa,b, P. De Castro Manzanoa, T. Dorigoa, U. Dossellia, F. Gasparinia,b, A. Gozzelinoa, S. Lacapraraa, M. Margonia,b, A.T. Meneguzzoa,b, M. Michelottoa, F. Montecassianoa, D. Pantanoa, N. Pozzobona,b, P. Ronchesea,b, R. Rossina,b, F. Simonettoa,b, E. Torassaa, M. Zanettia,b,

P. Zottoa,b, G. Zumerlea,b

INFN Sezione di Paviaa, Universit`a di Paviab, Pavia, Italy

A. Braghieria, F. Fallavollitaa,b, A. Magnania,b, P. Montagnaa,b, S.P. Rattia,b, V. Rea, M. Ressegotti, C. Riccardia,b, P. Salvinia, I. Vaia,b, P. Vituloa,b

INFN Sezione di Perugiaa, Universit`a di Perugiab, Perugia, Italy

L. Alunni Solestizia,b, G.M. Bileia, D. Ciangottinia,b, L. Fan `oa,b, P. Laricciaa,b, R. Leonardia,b,

G. Mantovania,b, V. Mariania,b, M. Menichellia, A. Sahaa, A. Santocchiaa,b, D. Spiga

INFN Sezione di Pisaa, Universit`a di Pisab, Scuola Normale Superiore di Pisac, Pisa, Italy

K. Androsova, P. Azzurria,15, G. Bagliesia, J. Bernardinia, T. Boccalia, L. Borrello, R. Castaldia, M.A. Cioccia,b, R. Dell’Orsoa, G. Fedia, A. Giassia, M.T. Grippoa,29, F. Ligabuea,c, T. Lomtadzea, L. Martinia,b, A. Messineoa,b, F. Pallaa, A. Rizzia,b, A. Savoy-Navarroa,31, P. Spagnoloa, R. Tenchinia, G. Tonellia,b, A. Venturia, P.G. Verdinia

INFN Sezione di Romaa, Sapienza Universit`a di Romab, Rome, Italy

L. Baronea,b, F. Cavallaria, M. Cipriania,b, N. Dacia, D. Del Rea,b,15, M. Diemoza, S. Gellia,b, E. Longoa,b, F. Margarolia,b, B. Marzocchia,b, P. Meridiania, G. Organtinia,b, R. Paramattia,b, F. Preiatoa,b, S. Rahatloua,b, C. Rovellia, F. Santanastasioa,b

INFN Sezione di Torino a, Universit`a di Torino b, Torino, Italy, Universit`a del Piemonte Orientalec, Novara, Italy

N. Amapanea,b, R. Arcidiaconoa,c,15, S. Argiroa,b, M. Arneodoa,c, N. Bartosika, R. Bellana,b, C. Biinoa, N. Cartigliaa, F. Cennaa,b, M. Costaa,b, R. Covarellia,b, A. Deganoa,b, N. Demariaa, B. Kiania,b, C. Mariottia, S. Masellia, E. Migliorea,b, V. Monacoa,b, E. Monteila,b, M. Montenoa, M.M. Obertinoa,b, L. Pachera,b, N. Pastronea, M. Pelliccionia, G.L. Pinna Angionia,b, F. Raveraa,b, A. Romeroa,b, M. Ruspaa,c, R. Sacchia,b, K. Shchelinaa,b, V. Solaa, A. Solanoa,b, A. Staianoa, P. Traczyka,b

INFN Sezione di Triestea, Universit`a di Triesteb, Trieste, Italy

(31)

Korea

H. Kim, D.H. Moon, G. Oh

Hanyang University, Seoul, Korea

J.A. Brochero Cifuentes, J. Goh, T.J. Kim

Korea University, Seoul, Korea

S. Cho, S. Choi, Y. Go, D. Gyun, S. Ha, B. Hong, Y. Jo, Y. Kim, K. Lee, K.S. Lee, S. Lee, J. Lim, S.K. Park, Y. Roh

Seoul National University, Seoul, Korea

J. Almond, J. Kim, J.S. Kim, H. Lee, K. Lee, K. Nam, S.B. Oh, B.C. Radburn-Smith, S.h. Seo, U.K. Yang, H.D. Yoo, G.B. Yu

University of Seoul, Seoul, Korea

M. Choi, H. Kim, J.H. Kim, J.S.H. Lee, I.C. Park, G. Ryu

Sungkyunkwan University, Suwon, Korea

Y. Choi, C. Hwang, J. Lee, I. Yu

Vilnius University, Vilnius, Lithuania

V. Dudenas, A. Juodagalvis, J. Vaitkus

National Centre for Particle Physics, Universiti Malaya, Kuala Lumpur, Malaysia

I. Ahmed, Z.A. Ibrahim, M.A.B. Md Ali32, F. Mohamad Idris33, W.A.T. Wan Abdullah, M.N. Yusli, Z. Zolkapli

Centro de Investigacion y de Estudios Avanzados del IPN, Mexico City, Mexico

H. Castilla-Valdez, E. De La Cruz-Burelo, I. Heredia-De La Cruz34, R. Lopez-Fernandez, J. Mejia Guisao, A. Sanchez-Hernandez

Universidad Iberoamericana, Mexico City, Mexico

S. Carrillo Moreno, C. Oropeza Barrera, F. Vazquez Valencia

Benemerita Universidad Autonoma de Puebla, Puebla, Mexico

I. Pedraza, H.A. Salazar Ibarguen, C. Uribe Estrada

Universidad Aut ´onoma de San Luis Potos´ı, San Luis Potos´ı, Mexico

A. Morelos Pineda

University of Auckland, Auckland, New Zealand

D. Krofcheck

University of Canterbury, Christchurch, New Zealand

P.H. Butler

National Centre for Physics, Quaid-I-Azam University, Islamabad, Pakistan

Şekil

Table 1: Summary of estimated systematic uncertainties relative to the given mode for the last p T bin 2.5 &lt; p T &lt; 3.0 GeV/c for PbPb and pPb data.
Figure 1: Leading (α = 1) and subleading (α = 2) modes for n = 2 as a function of p T , measured
Figure 2: Leading (α = 1) and subleading (α = 2) modes for n = 3 as a function of p T , measured
Figure 3: Leading (α = 1) and subleading (α = 2) modes for n = 2 as a function of p T , mea-
+6

Referanslar

Benzer Belgeler

Uçak tipine göre gruplar arası işitme kaybı ortalamalarının karşılaştırıldığı istatistik çalışmamızda sağ kulak 4KHz’de hava ve kemik yolu sonuçları ve

Talebin karşılanma durumunu gösteren yıllık kararlı güç çıkışı oranı, sadece rüzgâr enerjisi santrali için yaklaşık %72,49 olurken, melez sistem için yaklaşık

With advances in medical therapy for elevated pulmonary artery pressure (PAP) and liver transplant surgery, survival of patients with PoPHT and advanced liver disease

Makroskopik Nair adezyon skalası ve mikroskopik Zühlke adezyon skalasına göre propolis grubunda diğer gruplara göre istatistiksel anlamlı olarak daha fazla yapışıklık

According to the analysis of the results, the regression model confirms the analogous findings reported by other authors, growth in economic activity tends to lead to a decrease

Kur’an mealinin tamamında “Yalan” ile ilgili olarak; yalan söylemek, yalanlamak, yalancı, yalan uydurmak, yalan saymak, yalana kulak vermek, yalancılıkla suçlamak, yalan

Cosmology, Ministry of Education; Shanghai Key Laboratory for Particle Physics and Cosmology; Institute of Nuclear and Particle Physics, Shanghai 200240, People ’s Republic of China..

13 Aralık 1914'te Eenebi Anonim ve Sermayesi Eshama Münkasım Şirketler ile Eenebi Sigorta Şirketleri Hakkındaki Kanun-u Muvakkat çıkarıldı ve ülkede faaliyette bulunan