Loading [MathJax]/jax/output/HTML-CSS/jax.js

Study of the CKM angle γ sensitivity using flavor untagged B0s˜D()0ϕ decays

  • A sensitivity study on the measurement of the CKM angle γ from B0s˜D()0ϕ decays is conducted using the D-meson reconstructed in the quasi flavour-specific modes Kπ, K3π, and Kππ0, as well as CP-eigenstate modes KK and ππ, where the notation ˜D0 corresponds to a D0 or ¯D0 meson. The LHCb experiment is presented as a use case. A statistical uncertainty of approximately 819 can be achieved with the pp collision data collected in the LHCb experiment from 2011 to 2018. The sensitivity to γ should be of the order 38 after accumulating 23 fb-1 of pp collision data by 2025, whereas it is expected to improve further by 300 fb-1 by the second half of the 2030 decade. The accuracy is dependent on the strong parameters r()B and δ()B, which together with γdescribe the interference between the leading amplitudes of the B0s˜D()0ϕ decays.
  • The shell correction method proposed by Strutinsky [1, 2] is widely used in macroscopic-microscopic approaches for calculating the properties of atomic nuclei, such as the potential energy surface, ground-state masses and deformations, and fission barriers. At zero temperature, the ground-state masses can be calculated quickly in terms of the macroscopic-microscopic framework [3]. However, calculations of the temperature-dependent shell corrections are quite time consuming [4] owing to a large number of combinations with various shapes of the thousands of potential nuclei.

    Consequently, some empirical or semi-empirical shell correction formulas have been proposed [5-12]. Based on Fermi-gas models without pairing correlation, an exponential dependence of the shell correction of the energy δEshell on the excitation energy E, i.e., δEshell=δEshell(E=0)exp(E/Ed) is proposed in Ref. [6] and has been widely employed in many different models. The damping factor Ed varies substantially from 15 to 60 MeV [4, 13, 14]. Another functional form for a shell correction to free energy δFshell is suggested in Ref. [7] for closed shell nuclei, where the ratio of temperature and hyperbolic sine function τ/sinh(τ), in which τT, is employed. In Ref. [8], a piecewise temperature dependent factor is introduced to a shell correction δEshell, where it stays at one until reaching the excitation energy of 35 MeV and then decreases exponentially. It was recently pointed out that both the shell corrections to energy δEshell and free energy δFshell obtained using the Woods-Saxon potential deviate from the exponential form exp(E/Ed) [12], and the shell correction δEshell at a temperature of 1 MeV, which corresponds to the excitation energy 2030 MeV, is as large as that at zero temperature.

    For open shell nuclei, the pairing correlation cannot be ignored, and a shell correction to the pairing energy at finite temperature should be considered. Consequently, the shell corrections to the energy δEshell, entropy TδSshell, and free energy δFshell are affected by the partial occupation of single particle levels [12].

    A reliable single-particle (s.p.) spectrum is an essential part of the Strutinsky shell correction method used for quantifying the shell effects. The covariant density functional theory (CDFT) [15-18] is a good candidate owing to its success in describing the properties of both spherical and deformed nuclei all throughout the nuclear chart, including superheavy nuclei [19-23], pseudospin symmetry [24-26], single-particle resonances [27, 28], hypernuclei [29-34], and shell correction [35-37].

    The basic thermal theory was developed in a period as early as the 1950s [38]. Later, the finite temperature Hartree-Fock approximation [39-41] and the finite temperature Hartree-Fock-Bogoliubov theory [42] were developed. In 2000, B. K. Agrawal et al. investigated the temperature dependence of shapes and pairing gaps for 166,170Er and rare-earth nuclei using the relativistic Hartree-BCS theory [43, 44]. In recent years, the finite temperature relativistic Hartree-Bogoliubov theory [45] and relativistic Hartree-Fock-Bogoliubov theory [46] for spherical nuclei were developed and employed in studies in which the relations between the critical temperature for the pairing transition and pairing gap at zero temperature are explored. Following the BCS limit of the HFB theory [42], in 2017, we developed a finite-temperature covariant density functional theory for an axial-deformed space and studied the shape evolution of 72,74Kr [47]. The shape evolutions of the octupole deformed nuclei 224Ra and even-even 144154Ba isotopes are studied. Such nuclei first go through an octupole shape transition within the temperature range of 0.50.95 MeV, followed by another quadrupole shape transition from a quadrupole deformed shape to a spherical shape within a higher temperature range of 1.02.2 MeV [48]. Moreover, it should be noted that the transition temperatures are roughly proportional to the corresponding deformations at the ground states [49].

    In this paper, shell corrections to both the internal energy and the free energy are discussed based on the single-particle spectrum extracted from the axial CDFT model. This paper is organized as follows. In Section II, the finite temperature CDFT model along with the shell correction method are briefly introduced. In Section III, numerical details and checks are presented. In Section IV, the results and discussions regarding the shell corrections to the energy, free energy, and entropy, as well as their dependence on the temperature and axial deformation, are explored. Finally, a brief summary and some interesting perspectives are provided in Section V.

    In the nuclear covariant energy density functional with a point-coupling interaction, the starting point is the following effective Lagrangian density [50],

    L=ˉψ(iγμμm)ψ12αS(ˉψψ)(ˉψψ)12αV(ˉψγμψ)(ˉψγμψ)12αTV(ˉψτγμψ)(ˉψτγμψ)13βS(ˉψψ)314γS(ˉψψ)414γV[(ˉψγμψ)(ˉψγμψ)]212δSν(ˉψψ)ν(ˉψψ)12δVν(ˉψγμψ)ν(ˉψγμψ)12δTVν(ˉψτγμψ)ν(ˉψτγμψ)14FμνFμνeˉψγμ1τ32ψAμ,

    (1)

    which is composed of a free nucleon term, four-fermion point-coupling terms, higher-order terms introduced for the effects of medium dependence, gradient terms to simulate the effects of a finite range, and electromagnetic interaction terms.

    For the Lagrangian density L, ψ is the Dirac spinor field of the nucleon with mass m, Aμ and Fμν are respectively the four-vector potential and field strength tensor of the electromagnetic field, e is the charge unit for protons, and τ is an isospin vector with τ3 being its third component. The subscripts S, V, and T in the coupling constants α, β, γ, and δ indicate the scalar, vector, and isovector couplings, respectively. The isovector-scalar (TS) channel is neglected owing to its small contributions to the description of nuclear ground state properties. In the full text, as a convention, we mark the isospin vectors with arrows and the space vectors in bold.

    In the framework of finite-temperature CDFT [48], the Dirac equation for a single nucleon reads

    [γμ(iμVμ(r))(m+S(r))]ψk(r)=0,

    (2)

    where ψk is the Dirac spinor, and

    S(r)=ΣS,

    Vμ(r)=Σμ+τΣμTV,

    are respectively the scalar and vector potentials in terms of the isoscalar-scalar ΣS, isoscalar-vector Σμ, and isovector-vector ΣμTV self-energies,

    ΣS=αSρS+βSρ3S+δSΔρS,

    Σμ=αVjμV+γV(jμV)3+δVΔjμV+eAμ,

    ΣμTV=αTVJμV+δTVΔJμV.

    The isoscalar density ρS, isoscalar current jμV, and isovector current jμTV are represented as follows:

    ρS(r)=kˉψk(r)ψk(r)[v2k(12fk)+fk],

    jμV(r)=kˉψk(r)γμψk(r)[v2k(12fk)+fk],

    jμTV(r)=kˉψk(r)τγμψk(r)[v2k(12fk)+fk],

    where ν2k (μ2k=1ν2k) is the BCS occupancy probability,

    ν2k=12(1εkλEk),

    μ2k=12(1+εkλEk),

    with λ being the Fermi surface and Ek being the quasiparticle energy.

    At finite temperature, the occupation probability ν2k will be altered by the thermal occupation probability of quasiparticle states fk, which is determined by temperature T as follows:

    fk=11+eEk/kBT,

    (7)

    where kB is the Boltzmann constant.

    In the BCS approach, the quasiparticle energy Ek can be calculated as

    Ek=(εkλ)2+Δk,

    (8)

    where εk is the single-particle energy, and the Fermi surface (chemical potential) λ is determined by meeting the conservation condition for particle number Nq,

    Nq=2k>0[v2k(12fk)+fk],

    (9)

    and Δk is the pairing energy gap, which satisfies the gap equation,

    Δk=12k>0VppkˉkkˉkΔkEk(12fk).

    (10)

    At finite temperature, the Dirac equation, mean-field potential, densities and currents, as well as the BCS gap equation in the CDFT, are solved iteratively on a harmonic oscillator basis. After a convergence is achieved, a single-particle spectrum up to 30 MeV is extracted as an input to the following shell correction method.

    The shell corrections to the energy of a nucleus within the mean-field approximation is defined as

    δEshell=ES˜E,

    (11)

    where ES is the sum of the single-particle energy εk of the occupied states calculated with the exact density of states gS(ε) in an axially deformed space,

    ES=occ.2εk=λεgS(ε)dε,

    gS(ε)=k2δ(εεk),

    and ˜E is the average energy calculated with the averaged density of states ˜g(ε),

    ˜E=˜λε˜g(ε)dε,

    ˜g(ε)=1γ+f(εεγ)gS(ε)dε,

    where ˜λ is a smoothed Fermi surface, γ is the smoothing parameter, and f(x) is the Strutinsky smoothing function,

    f(x)=1πex2L1/2M(x2),

    (14)

    with L1/2M(x2) being the M-order generalized Laguerre polynomial.

    At finite temperature T, Eqs. (17)-(21) for the shell corrections can be generalized in a straightforward manner, i.e., [12],

    δEshell(T)=E(T)˜E(T),

    (15)

    where for the energy E(T) of a system of independent particles at finite temperature,

    E(T)=λεk2εknTk,

    nTk=11+e(εkλ)/T.

    For the average energy ˜E(T),

    ˜E(T)=˜λε˜g(ε)nTεdε,

    nTε=11+e(εk˜λ)/T.

    The chemical potentials λ and ˜λ are conserved by the number of neutrons (protons),

    k2nTk=˜λdε˜g(ε)nTε=Nq.

    (18)

    The shell corrections to entropy S and free energy F at finite temperature read

    δSshell(T)=S(T)˜S(T),

    δFshell(T)=F(T)˜F(T),

    and are related to each other as

    δFshell(T)=δEshell(T)TδSshell(T).

    (20)

    For the entropy Sshell(T), the standard definition for the system of independent particles is adopted,

    S(T)=kBk2[nTklnnTk+(1nTk)ln(1nTk)].

    (21)

    The average part of S(T) is defined in an analogous manner by replaying the sum in Eq. (32) by the integral,

    ˜S(T)=kB+˜g(ε)[nTεlnnTε+(1nTε)ln(1nTε)]dε.

    (22)

    Taking the nucleus 144Sm with neutron shell closure as an example, the single-particle spectrum is calculated using the density functional PC-PK1 [50]. For the pairing correlation, the δ pairing force V(r)=Vqδ(r) is adopted, where the pairing strengths Vq are taken as 349.5 and 330.0 MeVfm3 for neutrons and protons, respectively. A smooth energy-dependent cutoff weight is introduced to simulate the effect of the finite range in the evaluation of the local pair density. Further details can be found in Ref. [48].

    At the mean-field level, the internal binding energies E at different axial-symmetric shapes can be obtained by applying constraints with a quadrupole deformation β2,

    H=H+12C(ˆQ2μ2)2,

    (23)

    where C is a spring constant, μ2=3AR24πβ2 is the given quadrupole moment with nuclear mass number A and radius R, and ˆQ2 is the expectation value of quadrupole moment operator ˆQ2=2r2P2(cosθ).

    The free energy is evaluated by F=ETS. For convenience, the temperature used is kBT in units of MeV, and the entropy applied is S/kB, which is unitless.

    First, a numerical check of the binding energy convergence based on size is conducted. In Fig. 1, the average binding energy as a function of the major shell number of the harmonic oscillator basis Nf is plotted. The binding energy is stable against the major shell number beginning from Nf=16 and is thus fixed as a proper number. Further checks at different temperatures T=0.02.0 MeV show that the temperature has a slight effect on the convergence.

    Figure 1

    Figure 1.  (color online) Average binding energy Eb/A as a function of the major shell number of the harmonic oscillator basis Nf obtained by the finite temperature CDFT+BCS calculations using the PC-PK1 density functional at zero temperature.

    Second, the mandatory plateau condition for the shell correction method is checked. The shell correction energy should be insensitive to the smoothing parameter γ and the order of the generalized Laguerre polynomial M, i.e.,

    δEshell(T)γ=0,δEshell(T)M=0.

    (24)

    In Fig. 2, the shell correction energy as a function of the above parameters γ and M for 144Sm is plotted. The unit of the smoothing range γ is ω0=41A1/3(1±13NZA) MeV, where the plus (minus) sign holds for neutrons (protons). It can be seen from Fig. 2 that the optimal values are γ=1.3 ω0 and M=3, which are consistent with previous relativistic calculations [35, 36].

    Figure 2

    Figure 2.  (color online) Neutron shell correction energy δEshell as a function of the smoothing parameter γ and the order of the generalized Laguerre polynomial M for 144Sm obtained by the finite temperature CDFT+BCS calculations using PC-PK1 density functional at zero temperature. The four different curves correspond to the orders M = 1, 2, 3, and 4, respectively.

    The free energy curves at temperatures 0, 0.4, 0.8, 1.2, 1.6 and 2.0 MeV for 144Sm are plotted in Fig. 3. The nucleus 144Sm has spherical minima for all temperatures, which are consistent with the shell closure at neutron number N=82. The energy curve is hard against the deformation near the spherical region. In addition, at low temperatures, a local minimum occurs at approximately β2=0.7 and a flat minimum occurs at approximately β2=0.4. However, it is shown that the fine details on the potential energy curves are washed out with increases in temperatures above T = 1.2 MeV, whereas the relative structures are maintained well at low temperatures.

    Figure 3

    Figure 3.  (color online) The relative free energy curves for 144Sm at different temperatures in the range of 0 to 2 MeV with a step of 0.4 MeV obtained by the constrained CDFT+BCS calculations using the PC-PK1 energy density functional. The ground state free energy at zero temperature is set to zero and is shifted up by 4 MeV for every 0.4 MeV temperature rise.

    Furthermore, the shell corrections to the energy, entropy, and free energy as functions of quadrupole deformation β2 at various temperatures T are shown in Fig. 4. The shell correction to the energy δEshell shows a deep valley at the spherical region demonstrating a strong shell effect. In addition, the valley becomes deeper for T0.8 MeV and then shallower with increasing temperature, whereas the two peaks decrease dramatically after T = 0.4 MeV. The peaks and valleys on the δEshell curve are basically consistent with details of the free energy curve in Fig. 3. In Fig. 4(b), the entropy shell correction curve TδSshell changes slightly. The corresponding amplitudes are generally much smaller compared with those of δEshell. As the difference between δEshell and TδSshell, the curves of shell correction to the free energy δFshell in Fig. 4(c) have similar shapes as δEshell. By contrast, with increasing temperature, both the peaks and valleys of δFshell diminish gradually. Similar to the shell correction at zero temperature, applying a shell correction at finite temperature is a good way to quantify the shell effects, which provide rich information.

    Figure 4

    Figure 4.  (color online) Neutron shell corrections to the energy δEshell, entropy TδSshell, and free energy δFshell as functions of quadrupole deformation β2 for 144Sm at different temperatures from 0 to 2 MeV with steps of 0.4 MeV obtained by the constrained CDFT+BCS calculations using PC-PK1 energy density functional.

    For the minimum states of 144Sm corresponding to increases in temperatures up to 4 MeV, the shell corrections to the energy δEshell, entropy TδSshell, and free energy δEshell are shown in Fig. 5. The non-monotonous behavior of δEshell with respect to temperature is significantly different from the exponential fading. The δEshell first decreases and then increases, monotonously approaching zero at high temperatures. This is consistent with the Woods-Saxon potential calculations carried out in Ref. [12]. In Ref. [8], a piecewise temperature-dependent factor is multiplied by the shell correction δEshell. The factor remains one for low temperatures below 1.65 MeV and then decreases exponentially. Here, the absolute amplitude first enlarges to approximately 120% at a temperature of 0.8 MeV and then bounces back to approximately 90% above 1.65 MeV. For this low temperature range, such behavior is roughly consistent with that in Ref. [8]. The exponential fading holds true for high temperatures for the current case and in Refs. [8] and [12].

    Figure 5

    Figure 5.  (color online) The temperature dependence of the shell corrections to the energy δEshell (black line), entropy TδSshell (red line), and free energy δFshell (blue line) with corresponding fitted empirical Bohr-Mottelson forms [7] (dashed lines) for the states with minimum free energy in 144Sm shown in Fig. 3 obtained using the constrained CDFT+BCS calculations applying the PC-PK1 energy density functional.

    Because δEshell is related to single-particle energy εk, Fermi surface λ, smoothed Fermi surface ˜λ, and temperature T according to Eqs. (23)-(27), the single particle levels near the neutron Fermi surface against the temperature for 144Sm is plotted in Fig. 6. It is shown that the spectrum is almost constant within the region of T<0.8 MeV and only changes slightly at high temperature. Meanwhile, both the original Fermi surface λ and the smoothed surface ˜λ decrease synchronically with increasing temperatures. Thus, excluding εk, λ, and ˜λ, the contribution directly from the temperature may play an important role in the behavior of the obtained shell correction to energy δEshell, as plotted in Fig. 5.

    Figure 6

    Figure 6.  (color online) Neutron single-particle levels as a function of temperature for 144Sm obtained by the constrained CDFT+BCS calculations using PC-PK1 energy density functional. The blue dashed line and red dash-dotted line represent the original and smoothed Fermi surfaces, respectively.

    The shell correction to the free energy δFshell increases monotonously and approaches zero at high temperatures. The shell correction to the entropy TδSshell behaves similar to δEshell. For comparison, the fitted shell corrections to free energy δFshell and entropy TδSshell in the Bohr-Mottelson form [7] are also plotted as dashed lines in Fig. 5. The Bohr-Mottelson [7] form for the shell correction to the free energy δFshell is expressed as

    δFshell(T)/δFshell(0)=ΨBM(T)=τsinh(τ),

    (25)

    where τ=c02π2T/ω0 with the fitting parameter c0=2.08. Similar to δFshell, TδSshell can also be approximated as

    TδSshell(T)/δFshell(0)=TδS0[τcoth(τ)1]sinh(τ),

    (26)

    when introducing the additional parameter δS0=2.15. With these two empirical formula, the shell corrections to the energy δEshell as the sum of δFshell and TδSshell take the following form,

    δEshell(T)=δEshell(0)τ+TδS0[τcoth(τ)1]sinh(τ),

    (27)

    noting that δEshell(0) equals δFshell(0). From Fig. 5, it can be clearly seen that both the shell corrections to the free energy δFshell and the entropy TδSshell can be approximated well using the Bohr-Mottelson forms.

    For more evidence, the same temperature dependence of the shell correction, for both neutrons and protons, is explored in other closed-shell nuclei. In Fig. 7, the shell corrections to the energy δEshell, entropy TδSshell, and free energy δFshell in 100Sn and 208Pb with the corresponding fitted empirical Bohr-Mottelson forms are plotted. In general, the curve shapes for all quantities are extremely similar to those of 144Sm in Fig. 5, proving the same temperature dependence. In addition, the fitting parameters c0 for the neutron and proton shell corrections to the free energy δFshell of 100Sn and 208Pb are 1.90, 2.08, 2.24, and 2.28, respectively, which are close to those of 144Sm 2.08. For the neutron and proton shell corrections to the entropy TδSshell, the values of parameter δS0 are 1.78, 2.00, 2.23, and 2.16, respectively, which are close to those of 144Sm 2.15. It was demonstrated that the Bohr-Mottelson forms describe well the shell corrections for closed-shell nuclei.

    Figure 7

    Figure 7.  (color online) Same as Fig. 5, but for neutrons and protons in 100Sn and 208Pb.

    The temperature dependence of the shell corrections to the energy δEshell, entropy TδSshell, and free energy δEshell was studied by employing the covariant density functional theory with the PC-PK1 density functional for a closed shell nucleus 144Sm. For numerical checks of the harmonic oscillator basis size, the major shell number is set to Nf=16. The plateau condition is satisfied by γ=1.3 ω0 and M=3.

    The fine details of the potential energy curves of free energy F are washed out with increases in temperatures above T = 1.2 MeV, whereas the relative structures are maintained well at low temperatures. Unlike the widely used exponential dependence, δEshell exhibits a non-monotonous behavior. First, it decreases to a certain degree, approaching 0.8 MeV, and then dissipates exponentially, where the direct contribution from the temperature may play an important role. Such a result is consistent with the Woods-Saxon potential calculations carried out in Ref. [12]. In addition, the shell corrections to both free energy δFshell and entropy can be approximated well using the Bohr-Mottelson form τ/sinh(τ) and [τcoth(τ)1]/sinh(τ), where τT. Further studies on shell corrections in other closed-shell nuclei, 100Sn and 208Pb, were conducted, and the same temperature dependencies were obtained.

    It was demonstrated that the shell correction at finite temperatures is a good tool for quantifying the shell effects and provides rich information. Thus, in future, open shell nuclei will also be explored, in which the shell correction to the pairing energy in the BCS framework should be explicitly considered. This is implemented in Ref. [12] with constant pairing strength G. For the δ-force BCS pairing, the development of the shell correction method is under way.

    The theoretical calculation was supported by the nuclear data storage system in Zhengzhou University. The authors appreciate the partial numerical work performed by ChuanXu Zhao.

    [1] M. Kobayashi and T. Maskawa, Prog. Theor. Phys. 49, 652 (1973) doi: 10.1143/PTP.49.652
    [2] I. Dunietz, Phys. Lett. B 270, 75 (1991)
    [3] M. Gronau and D. London, Phys. Lett. B 253, 483 (1991)
    [4] M. Gronau and D. Wyler, Phys. Lett. B 265, 172 (1991)
    [5] D. Atwood, I. Dunietz, and A. Soni, Phys. Rev. Lett. 78, 3257 (1997), arXiv:hep-ph/9612433 doi: 10.1103/PhysRevLett.78.3257
    [6] D. Atwood, I. Dunietz, and A. Soni, Phys. Rev. D 63, 036005 (2001)
    [7] A. Bondar, Proceedings of BINP special analysis meeting on Dalitz analysis, 24-26 Sep. 2002, unpublished;
    [8] A. Giri, Yu. Grossman, A. Soffer et al., Phys. Rev. D 68, 054018 (2003), arXiv:hep-ph/0303187
    [9] A. Poluektov et al. (Belle Collaboration), Phys. Rev. D 70, 072003 (2004), arXiv:hep-ex/0406067
    [10] Yu. Grossman, Z. Ligeti, and A. Soffer, Phys. Rev. D 67, 071301 (2003), arXiv:hep-ph/0210433
    [11] A. Ceccucci, T. Gershon, M. Kenzie et al., arXiv: 2006.12404 [physics.hist-ph]
    [12] M. W. Kenzie and M. P. Withehead (on behalf of the LHCb Collaboration), LHCb-CONF-2018-002, CERN-LHCb-CONF-2018-002
    [13] R. Aaij et al. (LHCb Collaboration), LHCb-PAPER-2020-019, CERN-EP-2020-175, arXiv: 2010.08483 [hep-ex]
    [14] R. Aaij et al (LHCb Collaboration), J. High Energy Phys. 1803, 059 (2018), arXiv: 1712.07428[hep-ex].
    [15] M. Schiller (on behalf of the LHCb Collaboration), CERN LHC seminar, and LHCb-PAPER-2020-030 in preparation.
    [16] M. W. Kenzie and M. P. Withehead (on behalf of the LHCb Collaboration), LHCb-CONF-2020-003, CERN-LHCb-CONF-2020-003
    [17] R. Aaij et al. (LHCb Collaboration), J. High Energy Phys. 08, 137 (2016), arXiv:1605.01082[hep-ex
    [18] R. Aaij et al. (LHCb Collaboration), J. High Energy Phys. 08, 041 (2019), arXiv:1906.08297[hep-ex
    [19] E. Kou et al. (Belle Ⅱ Collaboration), Prog. Theor. Exp. Phys., (2019), arXiv:1808.10567[hep-ex
    [20] I. Bediaga et al. (LHCb Collaboration), CERN-LHCC-2018-027, LHCB-PUB-2018-009, arXiv: 1808.08865 [hep-ex]
    [21] R. Aaij et al. (LHCb Collaboration), Phys. Lett. B 727, 403 (2013), arXiv:1308.4583[hep-ex
    [22] R. Aaij et al. (LHCb Collaboration), Phys. Rev. D 98, 071103 (2018), arXiv:1807.01892[hep-ex
    [23] R. Aaij et al. (LHCb Collaboration), Phys. Lett. B 777, 16 (2017), arXiv: 1708.06370 [hep-ex] And update in LHCb-PAPER-2020-036, in preparation
    [24] M. Gronau, Y. Grossman, N. Shuhmaher et al. , Phys. Rev. D 69, 113003 (2004), arXiv:hep-ph/0402055
    [25] M. Gronau, Y. Grossman, Z. Surujon et al., Phys. Lett. B 649, 61 (2007), arXiv:hep-ph/0702011
    [26] S. Ricciardi, LHCb-PUB-2010-005, CERN-LHCb-PUB-2010-005
    [27] Y. Amhis et al. (Heavy Flavor Averaging Group (HFLAV)), arXiv: 1909.12524 [hep-ex] and online updates at hflav.web.cern.ch
    [28] R. Aaij et al. (LHCb Collaboration), Phys. Lett. B 760, 117 (2016), arXiv:1603.08993[hep-ex
    [29] R. Aaij et al. (LHCb Collaboration), Phys. Rev. D 91, 112014 (2015), arXiv:1504.05442[hep-ex
    [30] Y. Grossman, A. Sofer, and J. Zupan, Phys. Rev. D 72, 031501 (2005), arXiv:hep-ph/0505270
    [31] M. Martone and J. Zupan, Phys. Rev. D 87, 034005 (2013), arXiv:1212.0165[hep-ph
    [32] M. Tanabashi et al. (Particle Data Group), Phys. Rev. D 98, 030001 (2018)
    [33] M. Gronau, Y. Grossman, Z. Surujon et al., Phys. Lett. B 649, 61 (2007)
    [34] R. Aaij et al. (LHCb Collaboration), Phys. Rev. Lett. 108, 101803 (2012), arXiv:1112.3183[hep-ex doi: 10.1103/PhysRevLett.108.101803
    [35] R. Aaij et al. (LHCb Collaboration), J. High Energy Phys. 04, 001 (2013), arXiv: 1301.5286 [hep-ex]; the fs/fd value was updated in the report LHCb-CONF-2013-011, CERN-LHCb-CONF-2013-011
    [36] T. Evans, S. Harnew, J. Libby et al., Phys. Lett. B 757, 520 (2016) and Erratum ibid B 765, 402 (2017), arXiv: 1602.07430 [hep-ex]
    [37] A. Bondar and T. Gershon, Phys. Rev. D 70, 091503 (2004), arXiv:hep-ph/0409281
    [38] The LHCb Collaboration public page
    [39] R. Aaij et al. (LHCb Collaboration), J. High Energy Phys. 1308, 117 (2013), arXiv:1306.3663[hep-ex
    [40] R. Aaij et al. (LHCb Collaboration), Phys. Rev. Lett. 118, 052002 (2017) [Erratum ibid 119, 169901 (2017)], arXiv: 1612.05140 [hep-ex]
    [41] R. Aaij et al. (LHCb Collaboration), J. High Energy Phys. 10, 097 (2014), arXiv:1408.2748[hep-ex
    [42] J. Charles et al. (The CKMfitter Group), CKMfitter Summer 2019 update
    [43] A. Bondar and A. Poluektov, Eur. Phys. J. C 47, 347 (2006), arXiv: hep-ph/0510246 and Eur. Phys. J. C 55, 51 (2008), arXiv: 0801.0840 [hep-ex]
    [44] J. Libby et al. (BES-Ⅲ Collaboration), Phys. Rev. D 82, 112006 (2010), arXiv:1010.2817[hep-ex
    [45] M. Ablikim et al. (BES-Ⅲ Collaboration), Phys. Rev. D 101, 112002 (2020), arXiv: 2003.00091 [hep-ex] and Phys. Rev. D 102, 052008 (2020), arXiv: 2007.07959 [hep-ex]
    [46] I. Adachi et al. (BaBar and Belle Collaboration), Phys. Rev. D 98, 110212 (2018)
    [47] J. Charles et al. (The CKMfitter Group), Eur. Phys. J. C 41, 1 (2005), arXiv: hep-ph/0406184, updated at [ckmfitter. in2p3.fr]
    [48] J. P. Lees et al. (BaBar Collaboration), Phys. Rev. D 84, 112007 (2011) and Erratum ibid D 87, 039901 (2013), arXiv: 1107.5751 [hep-ex]
    [49] V. Tisserand (on behalf of the BaBar and Belle Collaborations), eConf C070512, 009 (2007), arXiv:0706.2786[hep-ex
    [50] B. Sen, M. Walker, and M. Woodroofe, Statistica Sinica 19, 301-314 (2009)
    [51] G. Cowan (for the Particle Data Group), review on statistic in Phys. Rev. D 98, 030001 (2018) and 2019 online update
    [52] T. Evans, J. Libby, S. Malde et al., Phys. Lett. B 802, 135188 (2020), arXiv:1909.10196[hep-ex
    [53] H. Ikeda et al. (Belle Collaboration), Nucl. Instrum. Meth. A 441, 401-426 (2000)
    [54] D. M. Asner et al., Int. J. Mod. Phys. A 24, S1-794 (2009), arXiv: 0809.1869 [hep-ex]
    [55] R. Aaij et al. (LHCb Collaboration), Phys. Rev. D 98, 072006 (2018), arXiv:1807.01891[hep-ex
    [56] S. Nandi and D. London, Phys. Rev. D 85, 114015 (2012), arXiv:1108.5769[hep-ph
    [57] H. Sheng Chen, Nucl. Phys.- Proceedings Supplements B 59(13), 316-323 (1997)
    [58] S. Eidelman, Nucl. and Part. Phys. Proceedings 260, 238-241 (2015) doi: 10.1016/j.nuclphysbps.2015.02.050
    [59] Joint Workshop of future tau-charm factory (2018) Orsay, France and Joint Workshop on future charm-tau Factory (2019) Moscow, Russia.
  • [1] M. Kobayashi and T. Maskawa, Prog. Theor. Phys. 49, 652 (1973) doi: 10.1143/PTP.49.652
    [2] I. Dunietz, Phys. Lett. B 270, 75 (1991)
    [3] M. Gronau and D. London, Phys. Lett. B 253, 483 (1991)
    [4] M. Gronau and D. Wyler, Phys. Lett. B 265, 172 (1991)
    [5] D. Atwood, I. Dunietz, and A. Soni, Phys. Rev. Lett. 78, 3257 (1997), arXiv:hep-ph/9612433 doi: 10.1103/PhysRevLett.78.3257
    [6] D. Atwood, I. Dunietz, and A. Soni, Phys. Rev. D 63, 036005 (2001)
    [7] A. Bondar, Proceedings of BINP special analysis meeting on Dalitz analysis, 24-26 Sep. 2002, unpublished;
    [8] A. Giri, Yu. Grossman, A. Soffer et al., Phys. Rev. D 68, 054018 (2003), arXiv:hep-ph/0303187
    [9] A. Poluektov et al. (Belle Collaboration), Phys. Rev. D 70, 072003 (2004), arXiv:hep-ex/0406067
    [10] Yu. Grossman, Z. Ligeti, and A. Soffer, Phys. Rev. D 67, 071301 (2003), arXiv:hep-ph/0210433
    [11] A. Ceccucci, T. Gershon, M. Kenzie et al., arXiv: 2006.12404 [physics.hist-ph]
    [12] M. W. Kenzie and M. P. Withehead (on behalf of the LHCb Collaboration), LHCb-CONF-2018-002, CERN-LHCb-CONF-2018-002
    [13] R. Aaij et al. (LHCb Collaboration), LHCb-PAPER-2020-019, CERN-EP-2020-175, arXiv: 2010.08483 [hep-ex]
    [14] R. Aaij et al (LHCb Collaboration), J. High Energy Phys. 1803, 059 (2018), arXiv: 1712.07428[hep-ex].
    [15] M. Schiller (on behalf of the LHCb Collaboration), CERN LHC seminar, and LHCb-PAPER-2020-030 in preparation.
    [16] M. W. Kenzie and M. P. Withehead (on behalf of the LHCb Collaboration), LHCb-CONF-2020-003, CERN-LHCb-CONF-2020-003
    [17] R. Aaij et al. (LHCb Collaboration), J. High Energy Phys. 08, 137 (2016), arXiv:1605.01082[hep-ex
    [18] R. Aaij et al. (LHCb Collaboration), J. High Energy Phys. 08, 041 (2019), arXiv:1906.08297[hep-ex
    [19] E. Kou et al. (Belle Ⅱ Collaboration), Prog. Theor. Exp. Phys., (2019), arXiv:1808.10567[hep-ex
    [20] I. Bediaga et al. (LHCb Collaboration), CERN-LHCC-2018-027, LHCB-PUB-2018-009, arXiv: 1808.08865 [hep-ex]
    [21] R. Aaij et al. (LHCb Collaboration), Phys. Lett. B 727, 403 (2013), arXiv:1308.4583[hep-ex
    [22] R. Aaij et al. (LHCb Collaboration), Phys. Rev. D 98, 071103 (2018), arXiv:1807.01892[hep-ex
    [23] R. Aaij et al. (LHCb Collaboration), Phys. Lett. B 777, 16 (2017), arXiv: 1708.06370 [hep-ex] And update in LHCb-PAPER-2020-036, in preparation
    [24] M. Gronau, Y. Grossman, N. Shuhmaher et al. , Phys. Rev. D 69, 113003 (2004), arXiv:hep-ph/0402055
    [25] M. Gronau, Y. Grossman, Z. Surujon et al., Phys. Lett. B 649, 61 (2007), arXiv:hep-ph/0702011
    [26] S. Ricciardi, LHCb-PUB-2010-005, CERN-LHCb-PUB-2010-005
    [27] Y. Amhis et al. (Heavy Flavor Averaging Group (HFLAV)), arXiv: 1909.12524 [hep-ex] and online updates at hflav.web.cern.ch
    [28] R. Aaij et al. (LHCb Collaboration), Phys. Lett. B 760, 117 (2016), arXiv:1603.08993[hep-ex
    [29] R. Aaij et al. (LHCb Collaboration), Phys. Rev. D 91, 112014 (2015), arXiv:1504.05442[hep-ex
    [30] Y. Grossman, A. Sofer, and J. Zupan, Phys. Rev. D 72, 031501 (2005), arXiv:hep-ph/0505270
    [31] M. Martone and J. Zupan, Phys. Rev. D 87, 034005 (2013), arXiv:1212.0165[hep-ph
    [32] M. Tanabashi et al. (Particle Data Group), Phys. Rev. D 98, 030001 (2018)
    [33] M. Gronau, Y. Grossman, Z. Surujon et al., Phys. Lett. B 649, 61 (2007)
    [34] R. Aaij et al. (LHCb Collaboration), Phys. Rev. Lett. 108, 101803 (2012), arXiv:1112.3183[hep-ex doi: 10.1103/PhysRevLett.108.101803
    [35] R. Aaij et al. (LHCb Collaboration), J. High Energy Phys. 04, 001 (2013), arXiv: 1301.5286 [hep-ex]; the fs/fd value was updated in the report LHCb-CONF-2013-011, CERN-LHCb-CONF-2013-011
    [36] T. Evans, S. Harnew, J. Libby et al., Phys. Lett. B 757, 520 (2016) and Erratum ibid B 765, 402 (2017), arXiv: 1602.07430 [hep-ex]
    [37] A. Bondar and T. Gershon, Phys. Rev. D 70, 091503 (2004), arXiv:hep-ph/0409281
    [38] The LHCb Collaboration public page
    [39] R. Aaij et al. (LHCb Collaboration), J. High Energy Phys. 1308, 117 (2013), arXiv:1306.3663[hep-ex
    [40] R. Aaij et al. (LHCb Collaboration), Phys. Rev. Lett. 118, 052002 (2017) [Erratum ibid 119, 169901 (2017)], arXiv: 1612.05140 [hep-ex]
    [41] R. Aaij et al. (LHCb Collaboration), J. High Energy Phys. 10, 097 (2014), arXiv:1408.2748[hep-ex
    [42] J. Charles et al. (The CKMfitter Group), CKMfitter Summer 2019 update
    [43] A. Bondar and A. Poluektov, Eur. Phys. J. C 47, 347 (2006), arXiv: hep-ph/0510246 and Eur. Phys. J. C 55, 51 (2008), arXiv: 0801.0840 [hep-ex]
    [44] J. Libby et al. (BES-Ⅲ Collaboration), Phys. Rev. D 82, 112006 (2010), arXiv:1010.2817[hep-ex
    [45] M. Ablikim et al. (BES-Ⅲ Collaboration), Phys. Rev. D 101, 112002 (2020), arXiv: 2003.00091 [hep-ex] and Phys. Rev. D 102, 052008 (2020), arXiv: 2007.07959 [hep-ex]
    [46] I. Adachi et al. (BaBar and Belle Collaboration), Phys. Rev. D 98, 110212 (2018)
    [47] J. Charles et al. (The CKMfitter Group), Eur. Phys. J. C 41, 1 (2005), arXiv: hep-ph/0406184, updated at [ckmfitter. in2p3.fr]
    [48] J. P. Lees et al. (BaBar Collaboration), Phys. Rev. D 84, 112007 (2011) and Erratum ibid D 87, 039901 (2013), arXiv: 1107.5751 [hep-ex]
    [49] V. Tisserand (on behalf of the BaBar and Belle Collaborations), eConf C070512, 009 (2007), arXiv:0706.2786[hep-ex
    [50] B. Sen, M. Walker, and M. Woodroofe, Statistica Sinica 19, 301-314 (2009)
    [51] G. Cowan (for the Particle Data Group), review on statistic in Phys. Rev. D 98, 030001 (2018) and 2019 online update
    [52] T. Evans, J. Libby, S. Malde et al., Phys. Lett. B 802, 135188 (2020), arXiv:1909.10196[hep-ex
    [53] H. Ikeda et al. (Belle Collaboration), Nucl. Instrum. Meth. A 441, 401-426 (2000)
    [54] D. M. Asner et al., Int. J. Mod. Phys. A 24, S1-794 (2009), arXiv: 0809.1869 [hep-ex]
    [55] R. Aaij et al. (LHCb Collaboration), Phys. Rev. D 98, 072006 (2018), arXiv:1807.01891[hep-ex
    [56] S. Nandi and D. London, Phys. Rev. D 85, 114015 (2012), arXiv:1108.5769[hep-ph
    [57] H. Sheng Chen, Nucl. Phys.- Proceedings Supplements B 59(13), 316-323 (1997)
    [58] S. Eidelman, Nucl. and Part. Phys. Proceedings 260, 238-241 (2015) doi: 10.1016/j.nuclphysbps.2015.02.050
    [59] Joint Workshop of future tau-charm factory (2018) Orsay, France and Joint Workshop on future charm-tau Factory (2019) Moscow, Russia.
  • 加载中

Figures(41) / Tables(8)

Get Citation
D. Ao, D. Decamp, W. B. Qian, S. Ricciardi, H. Sazak, S. T’Jampens, V. Tisserand, Z. R. Wang, Z. W. Yang, S. N. Zhang and X. K. Zhou. Study of the CKM angle γ sensitivity using flavor untagged B0s˜D()0ϕ decays[J]. Chinese Physics C. doi: 10.1088/1674-1137/abd16d
D. Ao, D. Decamp, W. B. Qian, S. Ricciardi, H. Sazak, S. T’Jampens, V. Tisserand, Z. R. Wang, Z. W. Yang, S. N. Zhang and X. K. Zhou. Study of the CKM angle γ sensitivity using flavor untagged B0s˜D()0ϕ decays[J]. Chinese Physics C.  doi: 10.1088/1674-1137/abd16d shu
Milestone
Received: 2020-08-03
Revised: 2020-11-03
Article Metric

Article Views(2104)
PDF Downloads(44)
Cited by(0)
Policy on re-use
To reuse of Open Access content published by CPC, for content published under the terms of the Creative Commons Attribution 3.0 license (“CC CY”), the users don’t need to request permission to copy, distribute and display the final published version of the article and to create derivative works, subject to appropriate attribution.
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Email This Article

Title:
Email:

Study of the CKM angle γ sensitivity using flavor untagged B0s˜D()0ϕ decays

  • 1. University of Chinese Academy of Sciences, Beijing 100049, China
  • 2. Univ. Grenoble Alpes, Univ. Savoie Mont Blanc, CNRS, IN2P3-LAPP, Annecy, France
  • 3. STFC Rutherford Appleton Laboratory, Didcot, United Kingdom
  • 4. Université Clermont Auvergne, CNRS/IN2P3, LPC, Clermont-Ferrand, France
  • 5. Center for High Energy Physics, Tsinghua University, Beijing 100084, China
  • 6. School of Physics State Key Laboratory of Nuclear Physics and Technology, Peking University, Beijing 100871, China

Abstract: A sensitivity study on the measurement of the CKM angle γ from B0s˜D()0ϕ decays is conducted using the D-meson reconstructed in the quasi flavour-specific modes Kπ, K3π, and Kππ0, as well as CP-eigenstate modes KK and ππ, where the notation ˜D0 corresponds to a D0 or ¯D0 meson. The LHCb experiment is presented as a use case. A statistical uncertainty of approximately 819 can be achieved with the pp collision data collected in the LHCb experiment from 2011 to 2018. The sensitivity to γ should be of the order 38 after accumulating 23 fb-1 of pp collision data by 2025, whereas it is expected to improve further by 300 fb-1 by the second half of the 2030 decade. The accuracy is dependent on the strong parameters r()B and δ()B, which together with γdescribe the interference between the leading amplitudes of the B0s˜D()0ϕ decays.

    HTML

    I.   INTRODUCTION
    • Precision measurement of the CKM [1] angle γ (which is defined as arg[VudVub/VcdVcb]) in various B-meson decay modes is one of the main goals of flavor physics. Such measurements can be achieved by exploiting the interference of the decays that proceed via the bcˉus and buˉcs tree-level amplitudes, in which the determination of the relative weak phase γ is not affected by theoretical uncertainties.

      Several methods have been proposed to extract γ [2-11]. In LHCb, the best precision is obtained by combining the measurements of many decay modes, which yields γ=(74.0+5.05.8) [12]. This precision dominates the world average of γ from tree-level decays. LHCb has presented a new measurement based on the BPGGSZ method [7] using the full Run 1 and Run 2 data. The result is γ=(68.7+5.25.1) [13], and it constitutes the single best world measurement of γ. A 2σ difference between B+ and B0s results was observed since the summer of 2018. The B0s measurement is based on a single decay mode with only Run 1 data, i.e., B0sDsK±, and exhibits large uncertainty [14]. In October 2020, LHCb included a similar analysis with the decay B0sDsK±π+π [15] based on Run 1 and Run 2 data, for which γ2βs=(42±10±4±5). The most recent LHCb combination is therefore γ=(67±4) [16], and the new B+ (B0s) result is (64+45) ((82+1720)). Additional B0s decay modes will aid in improving the measurement precision of the B0s modes and the understanding of a possible discrepancy with respect to the B+ modes. The two analyses based on B0sDsK±(π+π) use time-dependent methods and are therefore strongly reliant on the B-tagging capabilities of the LHCb experiment. It should be noted that measurements exist at LHCb for B0 mesons [17, 18], which also exhibit quite good prospects as their current average is (82+89). These are based on the decay B0D0K0, where the B0 is self-tagged from the K0K+π decay. As opposed to measurements with B0s, those with B0 and B+ are also accessible at Belle II [19]. The prospects of these measurements at LHCb are provided in Ref. [20]. For the mode B0sDsK±, one may anticipate a precision on γ of the order of 4 after the end of LHC Run 3 in 2025, and 1 by 2035-2038. In combination with the B0 and B+ modes, the expected sensitivities are 1.5 and 0.35. The anticipated precision provided by Belle II is 1.5.

      In this work, B0s¯D()0ϕ decays, the observations of which were published by the LHCb experiment in 2013 [21] and 2018 [22], are used to determine γ. A novel method presented in Ref. [22] also demonstrated the feasibility of measuring B0s¯D0ϕ decays with a high purity. A partial reconstruction method is used for the ¯D0 meson [23]. A time-integrated method [24-26] was investigated, in which it was shown that information regarding CP violation is preserved in the untagged rate of B0s˜D()0ϕ (or of B0˜D0K0S), and that if a sufficient number of different D-meson final states is included in the analysis, this decay alone can be used to measure γ in principle. The sensitivity to γ is expected to be much better with the case of the B0s˜D()0ϕ decay than for B0˜D0K0S, as it is proportional to the decay width difference y=ΔΓ/2Γ defined in Section II and is equal to (6.3±0.3)% for B0s mesons and (0.05±0.50)% for B0 [27]. The sensitivity to γ from the B0s˜D()0ϕ modes arises from the interference between two colour-suppressed diagrams, as illustrated in Fig. 1. The relatively large expected value of the ratio of the ˉbˉucˉs and ˉbˉcuˉs tree-level amplitudes (20%40%; see Section IVA) serves as additional motivation for measuring γ in B0s˜D()0ϕ decays. In this study, five neutral D-meson decay modes, namely Kπ, K3π, Kππ0, KK, and ππ, are included, the event yields of which are estimated using realistic assumptions based on measurements from LHCb [22, 28, 29]. We justify the choice of these decays and also discuss the case of the two decay modes ˜D0K0Sπ+π and K0SK+K in Section III.

      Figure 1.  (color online) Feynman diagrams for (a) B0s¯D()0ϕ and (b) B0sD()0ϕ decays.

      In Section II, the notations and the selection of the D-meson decay final states are introduced. In Section III, the expected signal yields and their uncertainties are presented. In Section IV, the sensitivity that can be achieved using solely these decays is presented, and further improvements are briefly discussed. In Section V, the future expected precision on γ with B0s¯D()0ϕ at LHCb is discussed for the dataset available after LHC Run 3 by 2025 and after a possible second upgrade of LHCb by 2038. Finally, conclusions are presented in Section VI.

    II.   FORMALISM
    • Following the formalism introduced in Refs. [24-26], we define the amplitudes

      A(B0s¯D()0ϕ)=A()B,

      (1)

      A(B0sD()0ϕ)=A()Br()Bei(δ()B+γ),

      (2)

      where A()B and r()B are the magnitude of the B0s decay amplitude and amplitude magnitude ratio between the suppressed over the favored B0s decay modes, respectively, whereas δ()B and γ are the strong and weak phases, respectively. Neglecting mixing and CP violation in the D decays (see, for example, Refs. [30, 31]), the amplitudes into the final state f (denoted below as [f]D) and its CP conjugate ˉf are defined as

      A(ˉD0f)=A(D0ˉf)=Af,

      (3)

      A(D0f)=A(ˉD0ˉf)=AfrfDeiδfD,

      (4)

      where δfD and rfD are the strong phase difference and relative magnitude, respectively, between the D0f and ˉD0f decay amplitudes.

      The amplitudes of the full decay chains are given by

      ABfA(B0s[f]D()ϕ)=A()BA()f[1+r()BrfDei(δ()B+δfD+γ)],

      (5)

      ABˉfA(B0s[ˉf]D()ϕ)=A()BA()f[r()Bei(δ()B+γ)+rfDeiδfD].

      (6)

      The amplitudes of the CP-conjugate decays are obtained by changing the sign of the weak phase γ

      ˉABfA(ˉB0s[f]D()ϕ)=A()BA()f[r()Bei(δ()Bγ)+rfDeiδfD],

      (7)

      ˉABˉfA(ˉB0s[ˉf]D()ϕ)=A()BA()f[1+r()BrfDei(δ()B+δfDγ)].

      (8)

      Using the standard notations

      τ=Γst,Γs=ΓL+ΓH2,ΔΓs=ΓLΓH,y=ΔΓs2Γs,λf=qp.ˉABfABf,

      and assuming that |q/p|=1 (|q/p|=1.0003±0.0014 [32]), the untagged decay rate for the decay B0s/ˉB0s[f]D()ϕ is obtained by Eq. (10) of Ref. [33]:

      dΓ(B0s(τ)[f]D()ϕ)dτ+dΓ(ˉB0s(τ)[f]D()ϕ)dτeτ|ABf|2×[(1+|λf|2)cosh(yτ)2Re(λf)sinh(yτ)].

      (9)
    • A.   Time acceptance

    • Experimentally, owing to the trigger and selection requirements as well as inefficiencies in the reconstruction, the decay time distribution is affected by the acceptance effects. The acceptance correction has been estimated from pseudoexperiments based on a related publication by the LHCb collaboration [34]. It is described by an empirical acceptance function

      εta(τ)=(ατ)β1+(ατ)β(1ξτ),

      (10)

      with α=1.5, β=2.5, and ξ=0.01.

      Taking into account this effect, the time-integrated untagged decay rate is

      Γ(˜B0s[f]D()ϕ)=0[dΓ(B0s(τ)[f]D()ϕ)dτ+dΓ(ˉB0s(τ)[f]D()ϕ)dτ]εta(τ)dτ.

      (11)

      By defining the function

      g(x)=0exτ(1+ξτ(ατ)β)1+(ατ)βdτ,

      (12)

      and using Eq. (9), one obtains

      Γ(B0s[f]Dϕ)|ABf|2[(1+|λf|2)A2yRe(λf)B],

      (13)

      where A=1[g(1y)+g(1+y)]/2 and B=1[g(1y)g(1+y)]/2y. With y=(0.128±0.009)/2 for the B0s meson [27], one obtains A=0.488±0.005 and B=0.773±0.008. Examples of decay-time acceptance distributions are displayed in Fig. 2.

      Figure 2.  (color online) Examples of decay-time acceptance distributions for three different sets of parameters α, β, and ξ (nominal in green).

    • B.   Observables for D0 decays

    • The D-meson decays are reconstructed in quasi flavor-specific modes f(f)=Kπ+, K3π, and Kπ+π0; their CP-conjugate modes f+(ˉf)=K+π, K+3π, and K+ππ0; and CP-eigenstate modes fCP=K+K and π+π.

      In the following, we introduce the weak phase βsthat is defined as βs=arg(VtsVtbVcsVcb). From Eqs. (5), (7), and (13) and with λf=e2iβsˉABfABf, for a given number of untagged B0s mesons produced in the pp collisions at the LHCb interaction point, N(B0s), we can compute the number of B0sˉD0ϕ decays with the D meson decaying into the final state f. For the reference decay mode fKπ+, we obtain

      N(B0s[Kπ+]D[K+K]ϕ)=CKπ[2ByrBcos(δB+2βsγ)+A(1+rB2+4rBrKπDcosδBcos(δKπD+γ))],

      (14)

      in which the terms proportional to (rKπD)21 and yrKπD1 have been neglected (rKπD=5.90+0.340.25 % [27]). The best approximation for the scale factor CKπ is

      CKπ=N(B0s)×ε(B0s[Kπ+]D[K+K]ϕ)×Br(B0s[Kπ+]D[K+K]ϕ),

      (15)

      where ε(B0s[Kπ+]D[K+K]ϕ) is the global detection efficiency of this decay mode, and Br(B0s[Kπ+]D[K+K]ϕ) is its branching fraction. The value of the scale factor CKπ is estimated from the LHCb Run 1 data [22], the average fs/fd of the b-hadron production fraction ratio measured by LHCb [35], and the different branching fractions [32].

      For better numerical behavior, we use the Cartesian coordinate parameterization

      x()±=r()Bcos(δ()B±γ)  andy()±=r()Bsin(δ()B±γ).

      (16)

      Subsequently, Eq. (14) becomes

      N(B0s[Kπ+]D[K+K]ϕ)=CKπ[2By[xcos(2βs)ysin(2βs)]+A(1+x2+y2+2rKπD[(x++x)cosδKπD(y+y)sinδKπD])].

      (17)

      For three and four body final states K3π and Kππ0, multiple interfering amplitudes exist; therefore, their amplitudes and phases δfD vary across the decay phase space. However, an analysis that is integrated over the phase space can be performed in a very similar manner to two body decays with the inclusion of an additional parameter, namely the so-called coherence factor RfD, which has been measured in previous experiments [36]. The strong phase difference δfD is then treated as an effective phase that is averaged over all amplitudes. For these modes, we obtain an expression similar to (17):

      N(B0s[f]D[K+K]ϕ)=CKπFf[2By[xcos(2βs)ysin(2βs)]+A(1+x2+y2+2rfDRfD[(x++x)cosδfD(y+y)sinδfD])],

      (18)

      where Ff is the scale factor of the f decay relative to the Kπ decay and depends on the ratios of detection efficiencies and branching fractions of the corresponding modes

      Ff=CfCKπ=ε(Df)ε(DKπ)×[Br(D0f)+Br(ˉD0f)][Br(D0Kπ+)+Br(ˉD0Kπ+)].

      (19)

      The value of Ff for the different modes used in this study is determined from the LHCb measurements in the B±DK± and B±Dπ± modes, with two or four-body D decays [28, 29].

      The time-integrated untagged decay rate for B0s[ˉf]Dϕ is given by Eq. (13) by substituting ABfˉABˉf and λfˉλˉf=λ1f=e2iβs(ABˉf/ˉABˉf), which is equivalent to the change βsβs and γγ (i.e., x±x and y±y). Therefore, the observables are

      N(B0s[K+π]D[K+K]ϕ)=CKπ[2By[x+cos(2βs)+y+sin(2βs)]+A(1+x2++y2++2rKπD[(x++x)cosδKπD+(y+y)sinδKπD])],

      (20)

      and for the modes f+K+3π, K+ππ0

      N(B0s[f+]D[K+K]ϕ)=CKπFf[2By[x+cos(2βs)+y+sin(2βs)]+A(1+x2++y2++2rfDRf[(x++x)cosδfD+(y+y)sinδfD])].

      (21)

      Obviously, any significant asymmetries on the yield of the observable corresponding to Eq. (17) with respect to Eq. (20), or Eq. (18) with respect to Eq. (21), are a clear signature for CP violation.

      For the CP-eigenstate modes Dh+h(hK,π), we have rD=1 and δD=0. By following the same approach as that for quasi flavor-specific modes, the observables can be written as

      N(B0s[h+h]D[K+K]ϕ)=4CKπFhh[A(1+x2++y2++x++x)By((1+x++x+x+x+y+y)cos(2βs)+(y+y+y+xx+y)sin(2βs))].

      (22)

      Analogous to Ff, Fhh is defined as

      Fhh=ChhCKπ=ε(Dhh)ε(DKπ)×Br(D0hh)[Br(D0Kπ+)+Br(ˉD0Kπ+)]

      (23)

      and their values are determined in the same manner as Ff.

      For the modes K0Sπ+π and K0SK+K (i.e., K0Shh), we obtain

      N(B0s[K0Shh]D[K+K]ϕ)=2CKπFK0Shh×[By[(x++x)cos(2βs)+(y+y)sin(2βs)]+A(1+x2+y2+2(x++x)×rK0ShhD(m2+,m2)κK0ShhD(m2+,m2)cosδK0ShhD(m2+,m2))],

      (24)

      where FK0Shh is defined as in Eq. (23). The strong parameters rK0ShhD(m2+,m2), κK0ShhD(m2+,m2), and cosδK0ShhD(m2+,m2) vary over the Dalitz plot (m2+,m2)(m2(K0Sπ+), m2(K0Sπ)) and are defined in Section III.

    • C.   Observables for D0 decays

    • For the D0 decays, we consider the two modes D0D0π0 and D0D0γ, where the D0 mesons are reconstructed, as in the above, in quasi flavor-specific modes Kπ, K3π, and Kππ0 as well as CP-eigenstate modes ππ and KK. As demonstrated in Ref. [37], the formalism for the cascade B0sˉD0ϕ,ˉD0ˉD0π0 is similar to that of B0sˉD0ϕ. Therefore, the relevant observables can be written similarly to Eqs. (17), (18), (20), (21), and (22), by substituting CKπCKπ,Dπ0, rBrB, and δBδB (x±x± and y±y±):

      N(B0s[[Kπ+]Dπ0]D[K+K]ϕ)=CKπ,Dπ0[2By[xcos(2βs)ysin(2βs)]+A(1+x2+y2+2rKπD[(x++x)cosδKπD(y+y)sinδKπD])],

      (25)

      N(B0s[[K+π]Dπ0]D[K+K]ϕ)=CKπ,Dπ0[2By[x+cos(2βs)+y+sin(2βs)]+A(1+x2++y2++2rKπD[(x++x)cosδKπD+(y+y)sinδKπD])],

      (26)

      N(B0s[[f]Dπ0]D[K+K]ϕ)=CKπ,Dπ0Ff[2By[xcos(2βs)ysin(2βs)]+A(1+x2+y2+2rfDRf[(x++x)cosδfD(y+y)sinδfD])],

      (27)

      N(B0s[[f+]Dπ0]D[K+K]ϕ)=CKπ,Dπ0Ff[2By[x+cos(2βs)+y+sin(2βs)]+A(1+x2++y2++2rfDRf[(x++x)cosδfD+(y+y)sinδfD])],

      (28)

      N(B0s[[h+h]Dπ0]D[K+K]ϕ)=4CKπ,Dπ0Fhh[A(1+x2++y2++x++x)By((1+x++x+x+x+y+y)cos(2βs)+(y+y+y+xx+y)sin(2βs))].

      (29)

      In the case of D0D0γ, the formalism is very similar, except that there is an effective strong phase shift of π with respect to D0D0π0 [37]. The observables can be derived from the previous ones by substituting CKπ,Dπ0CKπ,Dγ and δBδB+π (i.e. x±x± and y±y±):

      N(B0s[[Kπ+]Dγ]D[K+K]ϕ)=CKπ,Dγ[2By[xcos(2βs)ysin(2βs)]+A(1+x2+y2+2rKπD[(x++x)cosδKπD+(y+y)sinδKπD])],

      (30)

      N(B0s[[K+π]Dγ]D[K+K]ϕ)=CKπ,Dγ[2By[x+cos(2βs)+y+sin(2βs)]+A(1+x2++y2++2rKπD[(x++x)cosδKπD(y+y)sinδKπD])],

      (31)

      N(B0s[[f]Dγ]D[K+K]ϕ)=CKπ,DγFf[2By[xcos(2βs)ysin(2βs)]+A(1+x2+y2+2rfDRf[(x++x)cosδfD+(y+y)sinδfD])],

      (32)

      N(B0s[[f+]Dγ]D[K+K]ϕ)=CKπ,DγFf[2By[x+cos(2βs)+y+sin(2βs)]+A(1+x2++y2++2rfDRf[(x++x)cosδfD(y+y)sinδfD])],

      (33)

      N(B0s[[h+h]Dγ]D[K+K]ϕ)=4CKπ,DγFhh[A(1+x2++y2+x+x)By((1x+x+x+x+y+y)cos(2βs)+(y++y+y+xx+y)sin(2βs))].

      (34)

      CKπ,Dπ0 and CKπ,Dγ are determined in the same manner as CKπ, i.e., from the LHCb Run 1 data [22] and taking into account the fraction of longitudinal polarization in the decay B0sD0ϕ,fL=(73±15±4)% [22], and the branching fractions Br(ˉD0ˉD0π0) and Br(ˉD0ˉD0γ) [32].

    III.   EXPECTED YIELDS
    • The LHCb collaboration has measured the yields of the B0s˜D()0ϕ and ˜D0Kπ modes using Run 1 data, corresponding to an integrated luminosity of 3 fb1 (Ref. [22]). Taking into account the cross-section differences among different center-of-mass energies, the equivalent integrated luminosities in different data over several years from LHCb are summarized in Table 1. The corresponding expected yields of the ˜D0 meson decaying into other modes are also estimated according to Refs. [28], [29], and [41]. The scaled results are listed in Table 2, where the longitudinal polarization fraction fL=(73±15±4)% [22] of B0s˜D0ϕ is considered so that the CP eigenvalue of the final state is well defined and is similar to that of the B0s˜D0ϕ mode.

      Years/Run s /TeV Int. lum./fb−1 Cross section Equiv. 7 TeV data
      2011 7 1.1 σ2011=38.9μb 1.1
      2012 8 2.1 1.17×σ2011 2.4
      Run 1 3.2 3.5
      2015 to 2018 (Run 2) 13 5.9 2.00×σ2011 11.8
      Total 9.1 15.3

      Table 1.  Integrated luminosities and cross-sections of LHCb Run 1 and Run 2 data. The integrated luminosities are obtained from Ref. [38] and the cross-sections are obtained from Refs. [39, 40].

      Expect. yield (Run 1 only)
      B0s˜D0(Kπ)ϕ 577 (132±13 [22])
      B0s˜D0(K3π)ϕ 218
      B0s˜D0(Kππ0)ϕ 58
      B0s˜D0(KK)ϕ 82
      B0s˜D0(ππ)ϕ 24
      B0s˜D0(K0Sππ)ϕ 54
      B0s˜D0(K0SKK)ϕ 8
      B0s˜D0ϕ mode D0π0 D0γ
      B0s˜D0(Kπ)ϕ 337 184
      (119 [22])
      B0s˜D0(K3π)ϕ 127 69
      B0s˜D0(Kππ0)ϕ 34 18
      B0s˜D0(KK)ϕ 48 26
      B0s˜D0(ππ)ϕ 14 8

      Table 2.  Expected yield of each mode for 9.1fb1 (Run 1 and Run 2 data). The expected yields for the B0s˜D0ϕ sub-modes are scaled by the longitudinal fraction of polarization fL=(73±15)%. To be scaled by 6.3 (90) for prospects after 2025 (2038) (see Section V).

      Several additional parameters are used in the sensitivity study, as indicated in Table 3. Most of these originate from D decays, and the scale factors F are calculated by using the data from Refs. [28] and [29], as well as the branching fractions from PDG [32].

      Parameter Value
      −2βS [mrad] 36.86±0.82 [42]
      y=ΔΓs/2Γs (%) 6.40±0.45 [27]
      rKπD (%) 5.90+0.340.25 [27]
      δKπD [deg] 188.9+8.28.9 [27]
      rK3πD (%) 5.49±0.06 [36]
      RK3πD (%) 43+1713 [36]
      δK3πD [deg] 128+2817 [36]
      rKππ0D (%) 4.47±0.12 [36]
      RKππ0D (%) 81±6 [36]
      δKππ0D [deg] 198+1415 [36]
      Scale factor (w.r.t. Kπ) (Stat. uncertainty only)
      FK3π (%) 37.8±0.1 [28]
      FKππ0 (%) 10.0±0.1 [29]
      FKK (%) 14.2±0.1 [28]
      Fππ (%) 4.2±0.1 [28]

      Table 3.  Other external parameters used in sensitivity study. The scale factors F are also listed.

      The expected numbers of signal events are also calculated from the full expressions provided in Sections IIB and IIC by using the detailed branching fraction derivations explained in Ref. [22], as well as scaling by the LHCb Run 1 and Run 2 integrated luminosities, as listed in Table 1. The obtained normalization factors CKπ, CKπ,Dπ0, and CKπ,Dγ are 608±67, 347±56, and 189±31, respectively. To compute the uncertainty on the normalization factors, we assume that it is possible to improve the global uncertainty on the measurement of the branching fraction of the decay modes B0s¯D()0ϕ, and of the polarization of the mode B0s¯D0ϕ, by a factor of 2 when adding the LHCb data from Run 2 [22]. The values of the three normalization factors are in good agreement with the yields listed in Table 2.

      The number of expected event yields and the value of the coherence factor RD listed in Tables 2 and 3 justify a posteriori our choice of performing the sensitivity study on γ with the D-meson decay modes Kπ, K3π, Kππ0, KK and ππ. By definition, the value of RD is 1 for two-body decays and rD=1 for CP-eigenstates, whereas for K3π, RD is approximately 43% and larger, namely 81%, for Kππ0. A larger RD results in stronger sensitivity to γ. According to Eqs. (20), (21), and (22), it is clear that the largest sensitivity to γ is expected to originate from the ordered D0 decay modes: KK, ππ, Kπ, Kππ0, and K3π, for the same number of selected events. Therefore, even with lower yields, the modes Kππ0 and ππ should be of interest; this is discussed in Section IV I.

      Returning to the modes D0K0Sππ and K0SKK, the scale factors are FK0Sππ=(9.3±0.1)% and FK0SKK=(1.4±0.1)% [41]. The strong parameters rK0ShhD(m2+,m2), κK0ShhD(m2+,m2), and cosδK0ShhD(m2+,m2) can be defined according the effective method presented in Ref. [43], by using quantum-correlated ˜D0 decays, in which the phase space (m2+, m2) is divided into N tailored regions or “bins” [44], such that in the bin of index i,

      Ki/Ki=rK0ShhD,i,ci=κK0ShhD,icosδK0ShhD,i, and si=κK0ShhD,isinδK0ShhD,i,

      where δK0ShhD,i is the strong phase difference and κK0ShhD,i is the coherence factor. A recent publication by the BES-III collaboration [45] combines its data with the results of CLEO-c [44], while applying the same technique to obtain the values of the ci, si, and K±i parameters varying with the phase space. The binning schemes are symmetric with respect to the diagonal in the Dalitz plot (m2+, m2) (i.e.±i). These results have also been compared to an amplitude model from the B-factories BaBar and Belle [46]. When porting the result between the BES-III/CLEO-c combination, obtained using quantum correlated ˜D0 decays, and LHCb for B0s˜D()0ϕ measurements, care should be taken with the bin conventions so that there may be a minus sign in the phase (which only affects si). The expected yield listed in Table 2 for the mode D0K0Sππ is 54 events, whereas it is 8 events for the decay D0K0SKK. Although the binning scheme in the latter case is only 2×2, its expected yield is definitely too small to be considered further. For D0K0Sππ, the binned method of Refs. [44, 45] splits the selected ˜D0 events over 2×8 bins, such that with Run 1 and Run 2, only approximately three events may populate each bin. This is the reason that, although the related observable is presented in Eq. (24), we decide not include that mode in the sensitivity study. This choice could eventually be revisited after Run 3, when approximately 340 B0sD0(K0Sππ)ϕ events should be available, at which point approximately 20 events may populate each bin.

    IV.   SENSITIVITY STUDY FOR RUN 1 & 2 LHCB DATASET
    • The sensitivity study consists of testing and measuring the value of the unfolded γ, r()B, and δ()B parameters and their expected resolution, after having computed the values of the observables according to various initial configurations and given the external inputs for the other involved physics parameters or associated experimental observables. To achieve this, a procedure involving global χ2 fit based on the CKMfitter package [47] has been established to generate pseudoexperiments and to fit samples of B0s˜D()0ϕ events.

      This section is organized as follows: In Section IVA, we explain the various configurations that we tested for the nuisance strong parameters r()B, and δ()B, as well as the value of γ. In Section IVB, we explain how the pseudoexperiments have been generated. In Section IVC, the first step of the method is illustrated with one- and two-dimensional (1-D and 2-D) p-value profiles for the γ, r()B, and δ()B parameters. Before showing how the γ, r()B, and δ()B parameters are unfolded from the generated pseudoexperiments in Section IVF, we discuss the stability of the former 1-D p-value profile for γ when changing the time acceptance parameters (Section IVD) and for a newly available binning scheme for the DK3π decay (Section IVE). Thereafter, the unfolded values for γ and precisions (sensitivity) for the Run 1 & 2 LHCb dataset for the various generated configurations of δ()B and r()B are presented in Section IVG. We conclude with Sections IVH and IVI, in which we study the intriguing case where γ=74 (see LHCb 2018 combination [12], recently superseded by [16]), and we test the effect of dropping or not dropping the least abundant expected decays modes B0s˜D()0(ππ)ϕ and B0s˜D()0(Kππ0)ϕ in the Run 1 & 2 LHCb dataset.

    • A.   Various configurations of γ, r()B, and δ()B parameters

    • The sensitivity study was performed with the CKM angle γ true value set to (65.66+0.902.65) (1.146 rad), as obtained by the CKMfitter group, while excluding any measured values of γ in its global fit [42]. As a reminder, the average of the LHCb measurements is γ=(74.0+5.05.8) [12]; therefore, the value γ=74 was also tested (see Subsection IVH).

      The value of the strong phases δ()B is a nuisance parameter that cannot be predicted or guessed by any argument, and therefore, six different values are assigned thereto: 0, 1, 2, 3, 4, and 5 rad (0, 57.3, 114.6, 171.9, 229.2, and 286.5). This corresponds to 36 tested configurations (6×6).

      As both interfering diagrams displayed in Fig. 1 are color-suppressed, the value of the ratio of the ˉbˉucˉs and ˉbˉcuˉs tree-level amplitudes r()B is expected to be |VubVcs|/|VcbVus|0.4. This assumption is well supported by the study performed with B0sDsK± decays by the LHCb collaboration, for which a value of rB=0.37+0.100.09 has been measured [14]. However, as the decay B0sDsK± is color-favored, it is important to test other values obtained from already measured colour-suppressed B-meson decays, as non-factorizing final state interactions can modify the decay dynamics [48]. Among these, the decay B0DK0 plays such a role, for which the LHCb has obtained rB=0.22+0.170.27 [42], which has been confirmed by a more recent and accurate computation: rB=0.265±0.023 [18]. The value of rB is known to impact the precision on γ measurements strongly as 1/rB [49]. Therefore, the two extreme values 0.22 and 0.40 for r()B were tested for the sensitivity study, whereas the values for rB and rB are expected to be similar.

      This leads to a total of 72 tested configurations for the r()B, δB, and δB parameters (2×6×6).

    • B.   Generation of pseudoexperiments for various parameter configurations

    • As the first step, different configurations for the observables corresponding to Sections IIB and IIC are computed. The observables are obtained with the value of the angle γ and with the values of the four nuisance parameters r()B and δ()B fixed to various sets of initial true values (see Section IVA), whereas the external parameters listed in Table 3 and the normalization factors CKπ, CKπ,Dπ0, and CKπ,Dγ are left free to vary within their uncertainties. In the second step, for the obtained observables, including their uncertainties that we assume to be their square root, as well as all the other parameters except for γ, r()B, and δ()B, a global χ2 fit is performed to compute the resulting p-value distributions of the γ, r()B, and δ()B parameters. As the third step, for the obtained observables, including their uncertainties, as well as all the other parameters except for γ, r()B, and δ()B, 4000 pseudoexperiments are generated according to Eqs. (14)-(34) for the various tested configurations. As the fourth step, for each of the generated pseudoexperiments, all of the quantities are varied within their uncertainties. Thereafter, a global χ2 fit is performed to unfold the value of the parameters γ, r()B0s, and δ()B0s for each of the 4000 generated pseudoexperiments. As the fifth step, for the distribution of the 4000 values of the fitted γ, r()B, and δ()B, an extended unbinned maximum likelihood fit is performed to compute the most probable value for each of the former five parameters, together with their dispersions. The resulting values are compared to their injected initial true values. Finally, the sensitivity to γ, r()B, and δ()B is deduced, and any bias correlation is eventually highlighted and studied.

    • C.   1-D and 2-D p-value profiles for γ, r()B, and δ()B parameters

    • Figure 3 displays the 1-D p-value profiles of γ at step 2 of the procedure described in Section IVB. The figure is obtained for an example set of initial parameters: γ=65.66 (1.146 rad), r()B=0.4, δB=3.0 rad, and δB=2.0 rad. The assumed integrated luminosity in this case is that of the LHCb data collected in Run 1 & 2. The corresponding fitted value is γ=(65.7+6.333.8), which is thus in excellent agreement with the initial tested true value. Figure 3 also depicts the corresponding distributions obtained from a full frequentist treatment on the Monte-Carlo simulation basis [50], where γ=(65.7+6.934.9). This is considered as a demonstration that the two estimates on γ are in quite fair agreement, at least at the 68.3% confidence level (CL), such that no obvious under-coverage is experienced with the nominal method, based on the ROOT function TMath::Prob [51]. On the upper part of the distribution, the relative under-coverage of the “Prob” method is approximately 6.3/6.991%. As opposed to the full frequentist treatment on the Monte-Carlo simulation basis, the nominal retained method allows performing computations of a very large number of pseudoexperiments within a reasonable amount of time and with non-prohibitive CPU resources. For the LHCb Run 1 & 2 dataset, 72 configurations of 4000 pseudoexperiments were generated (288000 pseudoexperiments in total). The entire study was repeated another two times for prospective studies with future anticipated LHCb data, such that more than approximately 864000 pseudoexperiments were generated for this publication (see Section V). In the same figure, one can also observe the effect of modifying the value of r()B from 0.4 to 0.22, for which γ=(65.7+12.060.7), where the upper uncertainty scales are roughly as expected:1/r()B (6.3×0.4/0.22=12.6). Compared to the full frequentist treatment on the Monte-Carlo simulation, where γ=(65.7+13.2), the relative under-coverage of the “Prob” method is approximately 12.0/13.291%. Finally, the p-value profile of γ is also displayed when dropping the information provided by the B0s˜D0ϕ mode, thus retaining only that of the B0s˜D0ϕ mode. In this case, γ is equal to (65.7+6.3(12.0)), r()B=0.4 (0.22), such that the CL interval is noticeably enlarged on the lower side of the γ angle distribution (further details can be found in Section VC).

      Figure 3.  (color online) Profile of p-value of global χ2 fit to γ after computing observables (top left) for a set of true initial parameters: γ=1.146 rad, r()B=0.4, δB=3.0 rad, and δB=2.0 rad (the corresponding distribution obtained from the full frequentist treatment on the Monte-Carlo simulation basis [50] is superimposed on the same distribution). The related p-value profile for r()B=0.22 is also presented (top right). The assumed integrated luminosity assumed is that of the LHCb data collected in Run 1 & 2. Profile of p-value of global χ2 fit to γ after computing observables, where only decay mode B0s˜D0ϕ is used, and for a set of true initial parameters: γ=1.146 rad, r()B=0.4 (bottom left) and 0.22 (bottom right), δB=3.0 rad, and δB=2.0 rad. In each figure, the vertical dashed red line indicates the initial γ true value, and the two horizontal dashed black lines refer to 68.3 and 95.4% CLs.

      Figure 4 displays the 1-D p-value profile of the nuisance parameters r()B and δ()B for the same set of initial parameters (γ=65.66 (1.146 rad), r()B=0.4, δB=3.0 rad, and δB=2.0 rad) and the same projected integrated luminosity. It can be observed that the p-value is the maximum at the initial tested value, as expected.

      Figure 4.  (color online) Profiles of p-value distributions of global χ2 fit to r()B (top left (right)) and δ()B (bottom left (right)) after computing observables for set of initial true parameters: γ=65.66 (1.146 rad), r()B=0.4, δB=3.0 rad, and δB=2.0 rad. In each figure, thevertical dashed red line indicates the initial r()B and δ()B true values, and the two horizontal dashed black lines refer to 68.3 and 95.4% CLs.

      The 2-D p-value profiles of the nuisance parameters r()B and δ()B as a function of γ are presented in Figs. 5-8. Figures 5 and 6 correspond to two other example configurations γ=1.146 rad, δB=1.0 rad, and δB=5.0 rad, and r()B=0.4 and r()B=0.22, respectively. Figures 7 and 8 represent the configurations γ=1.146 rad, δB=3.0 rad, and δB=2.0 rad, and r()B=0.4 and r()B=0.22, respectively. These 2-D views allow the correlation between the different parameters to be observed. In general, large correlations between δ()B and γ are observed. In the case of configurations where r()B=0.4, a large fraction of the δ()B vs. γ plane can be excluded at a 95% CL, whereas the fraction is significantly reduced for the corresponding r()B=0.22 configurations. For the δB vs. γ plane, the advantage of our Cartesian coordinate approach (see Section IIB and IIC) can easily be observed, together with the fact that in the case of the mode D0D0γ, there is an effective strong phase shift of π with respect to the D0D0π0 [37], such that additional constraints enable the fold ambiguities with respect to the associated δB vs. γ plane to be removed.

      Figure 5.  (color online) 2-D p-value profile distribution of nuisance parameters r()B and δ()B as a function of γ. In each figure, the dashed black lines indicate the initial true values: γ=65.66 (1.146 rad), δB=57.3 (1.0 rad), and δB=286.5 (5.0 rad), as well as r()B=0.4.

      Figure 6.  (color online) 2-D p-value profile distribution of nuisance parameters r()B and δ()B as a function of γ. In each figure, the dashed black lines indicate the initial true values: γ=65.66 (1.146 rad), δB=57.3 (1.0 rad), and δB=286.5 (5.0 rad), as well as r()B=0.22.

      Figure 7.  (color online) 2-D p-value profile distribution of nuisance parameters r()B and δ()B as a function of γ. In each figure, the dashed black lines indicate the initial true values: γ=65.66 (1.146 rad), δB=171.9 (3.0 rad), and δB=114.6 (2.0 rad), as well asr()B=0.4.

      Figure 8.  (color online) 2-D p-value profile distribution of nuisance parameters r()B and δ()B as a function of γ. In each figure, the dashed black lines indicate the true values: γ=65.66 (1.146 rad), δB=171.9 (3.0 rad), and δB=114.6 (2.0 rad), as well as r()B=0.22.

    • D.   Effect of time acceptance parameters

    • Figure 9 demonstrates that for a tested configuration of γ=1.146 rad, r()B=0.4, and δ()B=1.0 rad, the impact of the time acceptance parameters A and B can eventually be non negligible and the parameters affect the profile distribution of the p-value of the global χ2 fit to γ. For the given example, the fitted value of γ is either (65.3+14.338.4) or (66.5+13.851.0) when the time acceptance is either accounted for or not. The reason due to which the precision improves when the time acceptance is taken into account may not be intuitive. This is because, for B/A1.6, as opposed to the case of B/A1.0, the impact of the first term in Eq. (14), which is directly proportional to cos(δB+2βsγ), is amplified with respect to the second term, for which the sensitivity to γ is more diluted.

      Figure 9.  (color online) Profile of p-value distribution of global χ2 fit to γ for a set of true initial parameters: γ=1.146 rad, r()B=0.4, and δ()B=1.0 rad. The assumed integrated luminosity is that of the LHCb data collected in Run 1 & 2 when the time acceptance values A and B are set to 1 in Eqs. (14) to (34): no time acceptance (top left) or to their nominal values A=0.488±0.005 and B=0.773±0.008 (top right), as computed in Section IIA. The dashed red line indicates the initial γ true value γ=65.66 (1.146 rad).

      Even if the parameters A and B are computed to a precision at the percentage level (Section IIA), we further investigate the effect of changing their values. Note that for this study, the overall efficiency is maintained constant, whereas the shape of the acceptance function is varied. The values α, β, and ξ are changed in Eq. (10), and the results of those changes are listed in Table 4. When α increases, both A and B become larger, but the value of the ratio B/A decreases. When β or ξ decreases, the three values of A, B, and B/A increase. The effect of changing β or ξ alone is small. A modification of α has a much greater impact on A and B. However, all of these changes have a weak impact on the precision of the fitted γ value. This is quite encouraging, as it means that the relative efficiency loss caused by the time acceptance effects will not cause a significant change in the sensitivity to the CKM γ angle. As a result, the time acceptance requirements can be varied without substantial concern to improve the signal purity and statistical significance when analyzing the B0s˜D()0ϕ decays with LHCb data.

      α β ξ A B B/A Fitted γ ()
      1.0 2.5 0.01 0.367 0.671 1.828 66.5+13.840.1
      1.5 2.5 0.01 0.488 0.773 1.584 65.3+14.338.4
      2.0 2.5 0.01 0.570 0.851 1.493 65.3+13.237.8
      1.5 2.0 0.01 0.484 0.751 1.552 65.9+13.239.0
      1.5 3.0 0.01 0.491 0.789 1.607 66.5+13.238.4
      1.5 2.5 0.02 0.480 0.755 1.573 66.5+13.839.5
      1.5 2.5 0.005 0.492 0.783 1.591 65.3+13.836.7

      Table 4.  Expected value of γ as a function of different time acceptance parameters. The second line corresponds to the nominal values. The nominal set of parameters A and B is indicated in bold.

    • E.   Effect of new binning scheme for DK3π decay

    • According to Ref. [46], the averaged values of the K3π input parameters over the phase space, which are defined as

      RK3πDeiδK3πD=AˉD0K3π(x)AD0K3π(x)dxAˉD0K3πAD0K3π,

      (35)

      are used here and correspond to a relatively limited value for the coherence factor: RK3πD=(43+1713)% [36]. A more attractive approach may be to perform the analysis in disjoint bins of the phase space. In this case, the parameters are re-defined within each bin. New values for RK3πD and δK3πD in each bin from Ref. [52] have alternatively been employed. No noticeable changes were observed in the γ and r()B fitted p-value profiles, but it is possible that certain fold-effects on δ()B, e.g., as observed in Figs. 5-7, become less probable. The lack of significant improvement is expected as the ˜D0K3π mode is not the dominant decay, and also because the new measurements of RK3πD and δK3πD in each bin still have large uncertainties.

    • F.   Unfolding γ, r()B, and δ()B parameters from generated pseudoexperiments

    • As explained in Section IVB, for each of the tested γ, r()B, and δ()B configurations, 4000 pseudoexperiments are generated, for which the values of γ, r()B, and δ()B are unfolded from the global χ2 fits (see Section IVC for illustrations). Figure 10 displays the extended unbinned maximum likelihood fits to the nuisance parameters r()B and δ()B. The initial configuration is γ=65.66 (1.146 rad), r()B=0.4, δB=171.9 (3 rad), and δB=114.6 (2 rad), with an integrated luminosity that is equivalent to that of the LHCb Run 1 & 2 data. It can be compared with Fig. 4. All of the distributions are fitted with the Novosibirsk empirical function, the description of which contains a Gaussian core part and a left or right tail, depending on the sign of the tail parameter [53]. The fitted values of r()B are centered at their initial tested values of 0.4, with a resolution of 0.14, and no bias is observed. For δ()B, the fitted value is (176±42) ((104±13)) for an initial true value that is equal to 171.9 (114.6). The fitted value for δB is slightly shifted by approximately 2/3 of a standard deviation, but its measurement is much more precise than that of δB, as it is measured from both the D0D0γ and D0D0π0 observables.

      Figure 10.  (color online) Fit to distributions of nuisance parameters r()B (top left (right)) and δ()B (bottom left (right)) obtained from 4000 pseudoexperiments. The initial configuration is γ=65.66 (1.146 rad), r()B=0.4, δB=171.9 (3 rad), and δB=114.6 (2 rad). The distributions of δ()B are plotted and fitted within ±45of their initial true values. In the distributions, only the candidates with values of γ  [0, 90] are considered.

      Figure 11 presents the corresponding fit to the CKM angle γ, where the value r()B=0.22 is also tested. This figure can be compared to the initial p-value profiles illustrated in Fig. 3. As shown in Figs. 7 and 8, γ is correlated with the nuisance parameters r()B and δ()B. Such correlations may generate long tails in the distributions, as obtained from the 4000 pseudoexperiments. To account for these tails, extended unbinned maximum likelihood fits, constituting two Novosibirsk functions with opposite-side tails, are performed on the γ distributions. With an initial value of 65.66, the fitted value for γ returns a central value equal to μγ=(65.9±0.3) with a resolution of σγ=(8.8±0.2) when r()B=0.4, and μγ=(66.6±0.7) with a resolution of σγ=(14.4±0.5) when r()B=0.22. The worse resolution obtained with r()B=0.22 follows the empirical behavior 1/r()B (8.8×0.4/0.2216.0). Again, no bias is observed.

      Figure 11.  (color online) Fit to distributions of nuisance parameters γ obtained from 4000 pseudoexperiments. The initial configuration is γ=65.66 (1.146 rad), r()B=0.4 (left) and 0.22 (right), δB=171.9 (3 rad), and δB=114.6 (2 rad). In the distributions, only the candidates with a value of γ  [0, 90] are considered. The purple dashed curve represents tails generated by the correlations with the nuisance parameters r()B and δ()B, whereas the blue dashed curve indicates the core part of the distribution, and the plain red line is the sum of the two components of the fit.

      Finally, Fig. 12 displays the 2-D distributions of the nuisance parameters r()B and δ()B as a function of γ obtained from 4000 pseudoexperiments. This figure can be compared with the corresponding p-value profiles presented in Fig. 7.

      Figure 12.  (color online) 2-D distributions of nuisance parameters r()B and δ()B as a function of γ obtained from 4000 pseudoexperiments. In each figure, the horizontal dashed black lines indicate the initial true values: γ=65.66 (1.146 rad), δB=171.9 (3.0 rad), and δB=114.6 (2.0 rad), as well as r()B=0.4.

    • G.   Varying δ()B and r()B

    • According to Section IVA, 72 configurations of nuisance parameters δ()B and r()B have been tested for γ=65.66 (1.146 rad) and 4000 pseudoexperiments have been generated for each set, following the procedure described in Section IVB and illustrated in Section IVF. The assumed integrated luminosity in this section is that of the LHCb data collected in Run 1 & 2.

      The fitted mean values of γ (μγ) for r()B=0.4 and 0.22 as a function of δ()B, for an initial true value of 65.66 (1.146 rad), are displayed in Table 5, whereas the corresponding resolutions (σγ) are listed in Table 6. In general, the fitted means are compatible with the true γ values within less than one standard deviation. For r()B=0.4, the resolution varies from σγ=8.3 to 12.9. For r()B=0.22, the resolution is worse, as expected, and it varies from σγ=13.9 to 18.7. For r()B=0.22, the distribution of γ of the 4000 pseudoexperiments has its maximum above 90 for δ()B=286.5 and it is therefore not considered.

      δB δB
      0 57.3 114.6 171.9 229.2 286.5
      r()B=0.4
      0.0 66.2±0.4 65.9±0.4 67.4±0.4 65.9±0.4 65.9±0.4 67.4±0.4
      57.3 65.4±0.4 69.2±0.4 71.4±0.4 65.5±0.3 67.7±0.4 72.6±0.4
      114.6 65.8±0.4 70.5±0.4 72.9±0.4 65.9±0.3 69.3±0.4 73.8±0.4
      171.9 65.6+0.50.4 65.2±0.3 65.9±0.3 65.5±0.3 65.2±0.4 66.1±0.3
      229.2 64.9±0.4 67.6±0.4 68.9±0.4 65.0±0.4 67.1±0.4 69.0±0.4
      286.5 66.1±0.4 71.0±0.4 75.7±0.4 65.7±0.3 69.8±0.4 78.8±0.4
      δB δB
      0 57.3 114.6 171.9 229.2 286.5
      r()B=0.22
      0.0 67.2+0.91.0 68.1+1.00.9 68.3+0.90.8 67.1±0.8 68.4±0.9 69.0±0.8
      57.3 67.4±0.9 72.2±0.8 74.1+0.80.7 66.6±0.7 71.5±0.8 75.5±0.8
      114.6 65.7±0.9 71.8±0.6 74.9±0.6 68.0±0.6 71.2±0.7 74.9±0.6
      117.9 65.4±0.7 66.9±0.7 66.6±0.7 64.7±0.7 65.2±0.6 68.3±0.6
      229.2 65.9+0.91.0 69.1±0.8 70.1+0.70.8 67.7±0.7 67.4±0.7 71.0+0.60.7
      286.5 67.5±0.9 75.8±0.8 77.5+0.80.7 68.1±0.6 72.8+0.90.8 83.5+2.41.5

      Table 5.  Fitted mean values of γ (μγ) (in [deg]) for r()B=0.4 (top) and 0.22 (bottom), as a function of δ()B, for initial true value of 65.66.

      δB δB
      0 57.3 114.6 171.9 229.2 286.5
      r()B=0.4
      0.0 9.6±0.4 11.2±0.3 11.2±0.3 9.4±0.4 12.2±0.4 11.2±0.3
      57.3 11.2±0.3 12.6±0.3 11.8±0.3 10.2±0.3 12.4±0.3 12.9±0.3
      114.6 11.4±0.3 11.9±0.3 11.1±0.3 10.0±0.3 11.3±0.3 11.9±0.3
      171.9 8.3±0.4 9.8±0.3 8.8±0.2 8.9±0.3 9.5±0.3 9.0±0.2
      229.2 10.8±0.3 11.7±0.3 10.8±0.3 10.3±0.3 12.4±0.3 11.7±0.3
      286.5 11.0±0.3 12.9±0.3 11.6±0.3 9.2±0.2 11.7±0.3 13.2+0.60.5
      δB δB
      0 57.3 114.6 171.9 229.2 286.5
      r()B=0.22
      0.0 16.5±0.7 16.8±0.7 16.0±0.6 16.8+0.80.7 15.9+0.00.6 16.0±0.7
      57.3 16.7±0.6 18.1+0.90.8 17.1+1.00.9 14.3±0.5 16.8±0.7 17.6+1.31.1
      114.6 16.1+0.60.7 15.9±0.6 13.9±0.5 14.1±0.5 15.1±0.6 14.9±0.6
      171.9 15.7+0.70.6 14.5±0.5 14.4±0.5 15.5±0.6 15.7+0.00.5 14.0±0.5
      229.2 15.9±0.6 15.7+0.50.6 15.4±0.6 14.6+0.60.5 15.6±0.5 14.4±0.6
      286.5 16.9+0.70.6 18.0+1.31.1 16.7+1.21.0 14.9+0.50.6 16.1±0.7 18.7+2.82.1

      Table 6.  Fitted resolution of γ (σγ) (in [deg]) for r()B=0.4 (top) and 0.22 (bottom), as a function of δ()B.

      The obtained values for μγ and σγ are also displayed in Figs. 13 and 14. It is clear that the resolution on γ depends to the first order on r()B, and then to the second order on δ()B. The best agreement with respect to the tested initial true value of γ is obtained when δ()B=0 (0 rad) or 180 (π rad), and the best resolutions (the lowest values of σγ) are also obtained in this case. The largest CP violation effects and best sensitivity to γ are observed. In contrast, the worst sensitivity is obtained when δ()B=90 (π/2 rad) or 270 (3π/2 rad). The other best and worst positions for δ()B can easily be deduced from Eq. (14). In most cases, for r()B=0.4 (0.22), the value of the resolution is σγ10 (15) and the fitted mean value μγ65.66 or slightly larger.

      Figure 13.  (color online) Fitted mean values of γ (μγ) for r()B=0.22 (red circles) and 0.4 (blue squares) as a function of δ()B for initial true value of 65.66 (1.146 rad). All of the listed values are in [deg]. In each figure, the horizontal dashed black line indicates the initial γ true value. All of the plotted uncertainties are statistical only.

      Figure 14.  (color online) Fitted resolutions of γ(σγ), for r()B=0.22 (red circles) and 0.4 (blue squares) as a function of δ()Bfor initial true value of 65.66 (1.146 rad). All of the listed values are in [deg]. In each figure, the horizontal dashed black lines are guides for the eye at σγ=5, 10, 15, and 20. All of the plotted uncertainties are statistical only.

      For completeness, the fitted means and resolutions for the nuisances parameters r()B and δ()B are presented inFigs. A1-A4 in Appendix A. It can be observed that the fitted mean values of r()B and δ()B are in good agreement with their initial tested true values, within one standard deviation of their fitted resolutions.

      Figure A1.  (color online) Fitted mean values of rB (μrB), for r()B=0.22 (red circles) and 0.4 (blue squares) as a function of δ()B for initial true value of γ of 65.66 (1.146 rad). In each figure, the horizontal dashed black line indicates the initial rB true value, and the displayed uncertainties are the fitted resolutions on rB (i.e. σrB).

      Figure A2.  (color online) Fitted mean values of rB (μrB), for r()B=0.22 (red circles) and 0.4 (blue squares) as a function of δ()B for initial true value of γ of 65.66 (1.146 rad). In each figure, the horizontal dashed black line indicates the initial rB true value, and the displayed uncertainties are the fitted resolutions on rB (i.e. σrB).

      Figure A3.  (color online) Fitted mean values of δB (μδB), for r()B=0.22 (red circles) and 0.4 (blue squares) as a function of δ()B for initial true value of γ of 65.66 (1.146 rad). In each figure, the horizontal dashed black line indicates the initial δB true value, and the displayed uncertainties are the fitted resolutions on δB (i.e. σδB).

      Figure A4.  (color online) Fitted mean values of δB (μδB), for r()B=0.22 (red circles) and 0.4 (blue squares) as a function of δ()B for initial true value of γ of 65.66 (1.146 rad). In each figure, the dashed black line indicates the initial δB true value, and the displayed uncertainties are the fitted resolutions on δB (i.e. σδB).

    • H.   Case where γ equals 74

    • Configurations in which γ=74 (see Ref. [12]) have also been tested. The potential problem in this case is that, as the true value of γ is closer to the 90 boundary, the unfolding of this parameter may become more difficult for many configurations of the nuisance parameters r()B and δ()B. It is clear from Eq. (14) that the sensitivity to γ is null at 90. This is illustrated in Fig. B1 in Appendix B, which can be compared with Fig. 11. In this case, the initial tested configuration is γ=74, r()B=0.4 and 0.22, δB=171.9 (3 rad), and δB=114.6 (2 rad). For these configurations, μγ=(73.7±0.3) ((74.2±0.7)) and σγ=(7.7±0.3) ((14.7±0.6)) for r()B=0.4 (0.22). There is limited degradation of the resolution compared to the corresponding configuration when the true value of γ is 65.66. The fit to γ for the pseudoexperiments corresponding to the configuration γ=74, r()B=0.4 and 0.22, δB=57.3 (1 rad), and δB=286.5 (5 rad) is presented in Fig. B2 in Appendix B. For r()B=0.22, it can clearly be observed that the fitted γ value approaches the boundary limit of 90, and the corresponding resolution is approximately 18. Such behavior can be clearly understood from the 2-D distribution presented in Fig. 6. This is comparable to the case listed in Tables 5 and 6 when δ()B=286.5 (near 3π/2 rad) and r()B=0.22.

      Figure B1.  (color online) Fit to distributions of nuisance parameters γ obtained from 4000 pseudoexperiments. The initial configuration is γ=74, r()B=0.4 (left) and 0.22 (right), δB=171.9 (3 rad), and δB=114.6 (2 rad). In the distributions, only the candidates with a value of γ  [0, 90] are considered. The purple dashed curve represents the tails generated by the correlations with the nuisance parameters r()B and δ()B, whereas the blue dashed curve indicates the core part of the distribution, and the plain red line is the sum of the two components of the fit.

      Figure B2.  (color online) Fit to distributions of nuisance parameters γ obtained from 4000 pseudoexperiments. The initial configuration is γ=74, r()B=0.4 (left) and 0.22 (right), δB=57.9 (1 rad), and δB=286.5 (5 rad). In the distributions, only the candidates with a value of γ  [0, 90] are considered. The purple dashed curve represents the tails generated by the correlations with the nuisance parameters r()B and δ()B, whereas the blue dashed curve indicates the core part of the distribution, and the plain red line is the sum of the two components of the fit.

    • I.   Effect of using or not the B0s˜D()0(ππ)ϕ and B0s˜D()0(Kππ0)ϕ decays or not

    • As listed in Table 2 the expected yields for the D-meson decays to ππ and Kππ0 are somewhat lower than those for the other modes, down to a few tens of events. Again, these yields have been computed from LHCb studies on B±˜D0(π/K)±, as reported in Refs. [28] and [29], and normalized to Ref. [22] with respect to the mode B0s˜D()0(Kπ)ϕ. Therefore, the selections are not necessarily against the signals B0s˜D()0(ππ)ϕ and B0s˜D()0(Kππ0)ϕ, and the expected yields may be underestimated, as well as all the sub-decays listed in Table 2. Furthermore, it should be noted that the mode ππ is a CP-eigenstate, whereas the Kππ03-body decay also has a large coherence factor value RKππ0D=(81±6)% [36]. Nevertheless, the effect of using the B0s˜D()0(ππ)ϕ and B0s˜D()0(Kππ0)ϕ decays or not has been studied and is reported here, and in Section VC, the effect of including the decays B0s˜D0ϕ or not is also discussed for future more abundant datasets.

      According to Fig. C1 in Appendix C, there is a relative loss on the precision in the unfolded value of γ of approximately 3 to 15%, when the B0s˜D()0(ππ)ϕ decays are not used. Figure C2 in Appendix C indicates that a relative loss in precision of approximately 3 to 22% is observed when the B0s˜D()0(Kππ0)ϕ decays are not used.

      Figure C1.  (color online) Fit to distributions of γ obtained from 4000 pseudoexperiments for Run 1 & 2 LHCb dataset. The initial configuration is γ=65.66, rB=0.4 (left) and 0.22 (right), and (top) δB=171.9 (3 rad) and δB=114.6 (2 rad) or (bottom) δB=57.3 (1 rad) and δB=286.5 (5 rad) (w/o B0s˜D()0(ππ)ϕ). The purple dashed curve represents the tails generated by the correlations with the nuisance parameters r()B and δ()B, whereas the blue dashed curve is the core part of the distribution, and the plain red line is the sum of the two components of the fit.

      Figure C2.  (color online) Fit to distributions of γ obtained from 4000 pseudoexperiments for Run 1 & 2 LHCb dataset. The initial configuration is γ=65.66, rB=0.4 (left) and 0.22 (right), and (top) δB=171.9 (3 rad) and δB=114.6 (2 rad) or (bottom) δB=57.3 (1 rad) and δB=286.5 (5 rad) (w/o B0s˜D()0(Kππ0)ϕ). The purple dashed curve represents the tails generated by the correlations with the nuisance parameters r()B and δ()B, whereas the blue dashed curve is the core part of the distribution, and the plain red line is the sum of the two components of the fit.

    V.   PROSPECTIVE ON SENSITIVITY TO γ FOR RUNS 13 AND FULL HIGH-LUMINOSITY LHC (HL-LHC) LHCb DATASETS
    • The prospectives on the sensitivity to the CKM angle γ with B0s˜D()0ϕ decays have also been studied for the foreseen LHCb integrated luminosities at the end of LHC Run 3 and for the possible full HL-LHC future LHCb program. According to Ref. [20], the LHCb trigger efficiency will be improved by a factor of 2 at the beginning of LHC Run 3. The full expected LHCb dataset of pp collisions at s=13 TeV corresponding to the sum of the Run 1, 2, and 3 LHCb dataset should be equal to 23 fb-1 by 2025, whereas it is expected to be 300 fb-1 by the second half of the 2030 decade. The final integrated LHCb luminosity accounts for an LHCb detector upgrade phase II. In the following, the projected event yields, as listed in Table 2, after 2025 and after 2038 have been scaled by a factor Flater=6.3 and 90, respectively, and with uncertainties on observables of 1/Flater.

    • A.   Projected precision on γ determination with B0s˜D()0ϕ decays

    • For this prospective sensitivity study, we have made the conservative assumption that the precision on the strong parameters of the D-meson decays to Kπ, K3π, and Kππ0 listed in Table 3 should be improved by a factor of 2 at the end of the LHCb program (see the BES-III experiment prospectives [54]). The procedure described for the LHCb Run 1 & 2 data in Section IV has been repeated. The values of the normalization factors CKπ, CKπ,Dπ0, and CKπ,Dγ obtained for Run 1 & 2 (see Section III) have been scaled to their expected equivalent rate for Runs 1 to 3 and full HL-LHC LHCb datasets. The statistical uncertainties of the computed observables (see Secttion IVC) obtained for the Run 1 & 2 LHCb data have been scaled by the square root of a factor two times (trigger improvement) the relative increase in the anticipated collected B0s-meson yield: 2.2 (8.8) for the Run 1 to 3 (full HL-LHC) LHCb dataset. Thereafter, for the Run 1 & 2 sensitivity studies, the same 2×6×6 configurations of the r()B, and δ()B nuisance parameters have been tested (r()B=0.22 or 0.4 and δ()B=0, 1, 2, 3, 4, 5 rad, and γ=65.66 (1.146 rad)).

      The 2-D p-value distribution profiles of the nuisance parameters r()B and δ()B as a function of γ are presented in Figs. 15 and 16 for the expected Run13 LHCb dataset, and in Figs. 17 and 18 for the full HL-LHC LHCb dataset. For the purpose of these illustrations, the initial configuration of true values is γ=65.66 (1.146 rad), δB=171.9 (3.0 rad), and δB=114.6 (2.0 rad), as well as r()B=0.4 (0.22). Therefore, the distributions can be directly compared to those depicted in Figs. 7 and 8. The surface of the excluded regions at a 95.4% CL in r()B vs. γ and δ()B vs. γ clearly increases with the additional data, but even in the semi-asymptotic regime, for the full expected HL-LHC LHCb dataset, one can clearly observe possible strong correlations between γ and the nuisance parameters r()B and δ()B. This is also visible in Figs. D1 and D2 in Appendix D, which are the equivalent version for the full expected HL-LHC LHCb dataset of the Run 1 & 2 LHCb dataset presented in Figs. 5 and 6 for the configurations γ=65.66 (1.146 rad), δB=57.3 (1.0 rad), and δB=286.5 (5.0 rad), as well asr()B=0.4 (0.22).

      Figure 15.  (color online) 2-D p-value profiles of nuisance parameters r()B and δ()B (Runs 13) as a function of γ. In each figure, the dashed black lines indicate the initial true values γ=65.66 (1.146 rad), δB=171.9 (3.0 rad), and δB=114.6 (2.0 rad), as well as r()B=0.4.

      Figure 16.  (color online) 2-D p-value profiles of nuisance parameters r()B and δ()B for Run 13 LHCb dataset as a function of γ. In each figure, the dashed black lines indicate the initial true values γ=65.66 (1.146 rad), δB=171.9 (3.0 rad), and δB=114.6 (2.0 rad), as well as r()B=0.22.

      Figure 17.  (color online) 2-D p-value profiles of nuisance parameters r()B and δ()B for full HL-LHC LHCb dataset as a function of γ. In each figure, the dashed black lines indicate the initial true values γ=65.66 (1.146 rad), δB=171.9 (3.0 rad), and δB=114.6 (2.0 rad), as well as r()B=0.4.

      Figure 18.  (color online) 2-D p-value profiles of nuisance parameters r()B and δ()B for full HL-LHC LHCb dataset as a function of γ. In each figure, the dashed black lines indicate the initial true values γ=65.66 (1.146 rad), δB=171.9 (3.0 rad), and δB=114.6 (2.0 rad), as well as r()B=0.22.

      Figure D1.  (color online) 2-D p-value profiles of nuisance parameters r()B and δ()B for full HL-LHC LHCb dataset as a function of γ. In each figure, the dashed black lines indicate the initial r()B and δ()B (γ) true values: γ=65.66 (1.146 rad), δB=57.3 (1.0 rad), and δB=286.5(5.0 rad), as well as r()B=0.4.

      Figure D2.  (color online) 2-D p-value profiles of nuisance parameters r()B and δ()B for full HL-LHC LHCb dataset as a function of γ. In each figure, the dashed black lines indicate the initial true values γ=65.66 (1.146 rad), δB=57.3 (1.0 rad), and δB=286.5 (5.0 rad), as well as r()B=0.22.

      For the configuration γ=65.66 (1.146 rad), δB= 171.9 (3.0 rad), and δB=114.6 (2.0 rad), as well as r()B=0.4 (0.22), Fig. 19 presents the fitted γ distribution obtained for 4000 pseudoexperiments for the expected Run 13 LHCb dataset. The fitted values are μγ=(67.7±0.1) ((73.5±0.2)) and σγ=(3.5±0.1) ((5.5±0.2)) for r()B=0.4 (0.22). The fitted values presented in Fig. 21 are μγ=(68.1± 0.1) ((71.1±0.1)) and σγ=(2.5±0.1) ((5.3±0.1)) for r()B=0.4 (0.22), respectively, for the expected full HL-LHC LHCb dataset. The fitted values are slightly shifted up with respect to the initial γ true value but are compatible within one standard deviation. When comparing with the numbers listed in Table 6, it can be observed that the resolution improves as 8.8/3.5=2.5 (14.4/5.5=2.6)for r()B=0.4 (0.22) when moving from the Run 1 & 2 to the expected Run 13 LHCb datasets, whereas a factor of 2.2 is expected. However, when moving from the expected Run 13 to the full expected HL-LHC LHCb datasets, the improvement is only 3.5/2.5=1.4 (5.5/5.3=1.040) for r()B=0.4 (0.22), whereas one may naively expect an improvement 8.8/2.2=4.0. This is certainly partially owing to the strong correlations between the nuisance parameters r()B and δ()B. A more sophisticated simultaneous global fit to the nuisance parameters r()B and δ()B, and γ may be useful.

      Figure 19.  (color online) Fit to distributions of γ obtained from 4000 pseudoexperiments for expected Run 13 LHCb dataset. The initial configuration is γ=65.66, r()B=0.4 (left) and 0.22 (right), δB=171.9 (3 rad), and δB=114.6 (2 rad). The purple dashed curve represents the tails generated by the correlations with the nuisance parameters r()B and δ()B, whereas the blue dashed curve indicates the core part of the distribution, and the plain red line is the sum of the two components of the fit.

      Figure 21.  (color online) Fit to distributions of γ obtained from 4000 pseudoexperiments for expected full HL-LHC LHCb dataset. The initial configuration is γ=65.66, r()B=0.4 (left) and 0.22 (right), δB=171.9 (3 rad), and δB=114.6 (2 rad). The purple dashed curve represents the tails generated by the correlations with the nuisance parameters r()B and δ()B, whereas the blue dashed curve is the core part of the distribution, and the plain red line is the sum of the two components of the fit.

      It must also be remembered that TMath::Probstill exhibits some under-coverage, namely 79 (91)% and 94 (89)% for r()B=0.4 (0.22), with respect to the full frequentist treatment on the Monte-Carlo simulation basis [50], as presented in Fig. 23, with the Run 13 and full expected HL-LHC LHCb datasets, respectively. The relative scale factors FK3π, FKππ0, FKK, and Fππ used in this study already have a precision above 2%. The precision on the normalization factors CKπ, CKπ,Dπ0, and CKπ,Dγ may also benefit from another improved precision of the branching fraction of the decay modes B0s¯D()0ϕ and of the longitudinal polarization fraction in the mode B0s¯D0ϕ. However, the normalization factors are the same for all sets of Eqs. (14)-(34) for B0s˜D0ϕ or ˜D0(π0, γ)ϕ, and their improved precision should be a second order effect. All of the above listed improvements are expected to occur to benefit from the total expected HL-LHC LHCb dataset fully.

      Figure 23.  (color online) Profiles of p-value distributions of global χ2 fit to γ for set of true initial parameters γ=1.146 rad, r()B=0.4 (left) 0.22 (right), δB=3.0 rad, and δB=2.0 rad. The assumed integrated luminosity is that of the LHCb data expected to be collected after LHC Run 3 (top) and after the full HL-LHC (bottom) period. The corresponding distribution obtained from the full frequentist treatment on the Monte Carlo simulation basis [50] is superimposed on the profile obtained with TMath::Prob. In each figure, the vertical dashed red line indicates the initial γ true value, and the two horizontal dashed black lines refer to 68.3 and 95.4% CLs.

      The expected resolutions on γ for the other usual configuration (δ()B=0, 1, 2, 3, 4, 5 rad, and γ=65.66 (1.146 rad)) are presented in Fig. 20 for r()B=0.4 and in Fig. 22 for r()B=0.22, for the Run 1 & 2, Run 13, and full HL-LHC LHCb datasets. For r()B=0.4, the resolution mainly ranges from 3.4 to 7.8 for Run 13 and from 2.2 to 7.1, or better, for the full HL-LHC dataset. For r()B=0.22, the resolution ranges from 5.5 to 8.2 for Run 13 and from 3.3 to 7.8, or better, for the full HL-LHC dataset.

      Figure 20.  (color online) Fitted mean values of γ (σγ) for Run 1 & 2 (pink lozenges), Run 13 (blue squares), and full HL-LHC (red circles) LHCb dataset as a function of δ()B, for r()B=0.4 and initial true value of 65.66 (1.146 rad). In each figure, the horizontal dashed black lines are guides for the eye at σγ=5 and 10.

      Figure 22.  (color online) Fitted resolutions of γ (σγ) for Run 1 & 2 (pink lozenges), Run 13 (blue squares), and full HL-LHC (red circles) LHCb dataset as a function of δ()B, for r()B=0.22 and initial true value of 65.66 (1.146 rad). In each figure, the horizontal dashed black lines are guides for the eye at σγ=5, 10, and 15.

      Another expected improvement could arise from a time-dependent CP Dalitz plane analysis of the decay B0s˜D()0CPK+K, as anticipated in Ref. [55]. With the ultimate HL-LHC LHCb dataset, it should be possible to perform such an analysis, thus including the B0s˜D()0ϕ decay, to extract the CKM angle γ, as proposed a few years ago in [56].

      For completeness, an alternate definition of the resolution as half of the 68.3% CL frequentist intervals of the 1-D p-value profiles of a 68.3% CL is provided in Appendix E in Figs. E1 and E2. A better scaling of the performances is observed with the size of the datasets, whereas relatively worse resolutions are obtained with respect to those displayed in Figs. 20 and 22. However, the effects of the nuisance parameters r()B and δ()B are treated in a simplified manner compared to the full treatment by the generated pseudoexperiments.

      Figure E1.  (color online) Half of 68.3% CL intervals of 1-D p-value profiles of γ for Run 1 & 2 (pink lozenges), Run 13 (blue squares), and full HL-LHC (red circles) LHCb dataset as a function of δ()B, for r()B=0.4 and initial true value of 65.66. In each figure, the horizontal dashed black lines are guides for the eye.

      Figure E2.  (color online) Half of 68.3% CL intervals of 1-D p-value profiles of γ for Run 1 & 2 (pink lozenges), Run 13 (blue squares), and full HL-LHC (red circles) LHCb dataset as a function of δ()B, for r()B=0.22 and initial true value of 65.66. In each figure, the horizontal dashed black lines are guides for the eye.

    • B.   Effect of strong parameters from D-mesondecays and of y=ΔΓs/2Γs

    • Most of the strong parameters of the D-meson decays to Kπ, K3π, and Kππ0 are external parameters and are obtained from beauty- or charm-factories, such as BaBar, Belle, CLEO-c, and LHCb [27]. Improvements in their determination are expected soon from the updated BES-III experiment [50] or later, from future super τ-charm factories [57-59]. Several scenarios have been tested to verify the effects of these improvements to the γ sensitivity. With the set of parameters γ=1.146 rad (65.66), r()B=0.4, and δB=3.0 rad, as well as δB=2.0, the uncertainties of the current measurements of the D-meson parameters listed in Table 3 have been scaled down, and their effects on the fitted γ values from pseudoexperiments are listed in Table 7. As the uncertainties of the external parameters are not yet dominant (Run 1 & 2 data), a study has also been performed for the expected full HL-LHC dataset. However, with substantially more data, future improvements in the measurements of the strong parameters from the D-meson decays do not appear to influence the sensitivity to the CKM angle γsignificantly.

      Uncertainties of D-meson params. Now ×1/2 ×1/5 ×1/10
      Run 1 & 2 (r()B=0.4) 8.8±0.2 8.1±0.3 8.0±0.3 7.8±0.2
      Run 1 & 2 (r()B=0.22) 12.9±0.3 13.2±0.5 13.1±0.5 12.8±0.9
      Full HL-LHC (r()B=0.4) 2.6±0.1 2.5±0.1 2.5±0.1 2.5±0.1
      Full HL-LHC (r()B=0.22) 5.4±0.1 5.3±0.1 5.2±0.1 5.1±0.1

      Table 7.  Fitted resolutions of γ (σγ) in [deg] obtained from 4000 pseudoexperiments as a function of decreasing uncertainties of strong D-meson parameters (see Table 3).

      This exercise was repeated using the same initial configuration of the parameters γ, r()B, and δ()B, for the uncertainty on y=ΔΓs/2Γs. The results of this study are listed in Table 8, Again, no obvious sensitivity to these changes is highlighted, neither for Run 1 & 2 nor for the full HL-LHC dataset. To our knowledge, it should be stressed that the tested improvements in y have not been supported by any published prospective studies.

      Uncertainty on y=ΔΓs/2Γs Now ×1/2 ×1/5 ×1/10
      Run 1 & 2 (r()B=0.4) 8.8±0.2 8.3±0.2 8.2±0.2 8.1±0.3
      Run 1 & 2 (r()B=0.22) 12.9±0.3 12.6±0.4 12.5±0.5 12.5±0.5
      Full HL-LHC (r()B=0.4) 2.5±0.1 2.5±0.1 2.5±0.1 2.5±0.1
      Full HL-LHC (r()B=0.22) 5.3±0.1 5.3±0.1 5.2±0.1 5.2±0.1

      Table 8.  Fitted resolutions of γ (σγ) in [deg] obtained from 4000 pseudoexperiments as function of decreasing uncertainties of y=ΔΓs/2Γs. For the full HL-LHC dataset, the uncertainties for the strong D-meson parameters are divided by 2 with respect to the current measurements (see Table 3).

      From the above studies, it can be concluded that the possibly large correlations of γ with respect to the nuisances parameters r()B and δ()B definitely dominate the ultimate precision on γ for the extraction with the B0s˜D()0ϕ modes.

    • C.   Effect of using B0s˜D0ϕ decays or not

    • It has been demonstrated in Ref. [22] that the decays B0s˜D0ϕ can be reconstructed in a clean manner together with B0s˜D0ϕ, with a similar rate and a partial reconstruction method, where the γ or π0 produced in the decay of ˜D0 are omitted. Thus far, these modes have been included in the sensitivity studies. Figures F1-F6 in Appendix F present the 2-D p-value profiles of the nuisance parameters rB and δB as a function of γ as well as the fit to the distributions of γ obtained from 4000 pseudoexperiments for the Run 1 & 2, Run 13, and full HL-LHC LHCb datasets for the initial true values γ=65.66 (1.146 rad), δB=171.9 (3.0 rad), δB=114.6 (2.0 rad), and r()B=0.4 (0.22). The information from B0s˜D0ϕ decays has not been included in these figures. According to Figs. F2, F4, and F6 in Appendix F, there is a relative loss in the precision of the unfolded value of γ of approximately 20 (40%) when the B0s˜D0ϕ decays are not used for r()B=0.4 (0.22). For future datasets, the improvement obtained by including the B0s˜D0ϕ modes is less significant, but not negligible, and aids in improving the measurement of γ.

      Figure F1.  (color online) 2-D p-value profiles of nuisance parameters rB and δB, for the Run 1 & 2 LHCb dataset, as a function of γ. In each figure, the dashed black lines indicate the initial true values γ=65.66 (1.146 rad), δB=171.9 (3.0 rad), and rB=0.4 (left) and 0.22 (right) (w/o B0sD0ϕ).

      Figure F2.  (color online) Fit to distributions of γ obtained from 4000 pseudoexperiments for Run 1 & 2 LHCb dataset. The initial configuration is γ=65.66, rB=0.4 (left) and 0.22 (right), and δB=171.9 (3 rad) (w/o B0sD0ϕ). The purple dashed curve represents the tails generated by the correlations with the nuisance parameters r()B and δ()B, whereas the blue dashed curve indicates the core part of the distribution, and the plain red line is the sum of the two components of the fit.

      Figure F3.  (color online) 2-D p-value profiles of nuisance parameters rB and δB (Run 13 LHCb) as function of γ. In each figure, the dashed black lines indicate the initial true values γ=65.66 (1.146 rad), δB=171.9 (3.0 rad), and δB=114.6 (2.0 rad), as well as rB=0.4 (left) and 0.22 (right) (w/o B0sD0ϕ).

      Figure F4.  (color online) Fit to distributions of γ obtained from 4000 pseudoexperiments (Run 13 LHCb). The initial configuration is γ=65.66, rB=0.4 (left) and 0.22 (right), and δB=171.9 (3 rad) (w/o B0sD0ϕ). The purple dashed curve represents the tails generated by the correlations with the nuisance parameters r()B and δ()B, whereas the blue dashed curve indicates the core part of the distribution, and the plain red line is the sum of the two components of the fit.

      Figure F5.  (color online) 2-D p-value profiles of nuisance parameters rB and δB (full HL-LHC LHCb) as function of γ. In each figure, the dashed black lines indicate the initial true values γ=65.66 (1.146 rad) and δB=171.9 (3.0 rad), as well as rB=0.4 (left) and 0.22 (right) (w/o B0sD0ϕ).

      Figure F6.  (color online) Fit to distributions of γ obtained from 4000 pseudoexperiments (full HL-LHC LHCb). The initial configuration is γ=65.66, rB=0.4 (left) and 0.22 (right), and δB=171.9 (3 rad) (w/o B0sD0ϕ). The purple dashed curve represents the tails generated by the correlations with the nuisance parameters r()B and δ()B, whereas the blue dashed curve indicates the core part of the distribution, and the plain red line is the sum of the two components of the fit.

    VI.   CONCLUSIONS
    • Untagged B0s˜D()0ϕ decays provide another theoretically clean pathway to the measurement of the CKM-angle γ. By using the expected event yields for D decays on Kπ, K3π, Kππ0, KK, and ππ, we have demonstrated that a precision on γ of approximately 8 to 19 can be achieved with LHCb Run 1 & 2 data. With additional data, a precision on γ of 38 can be achieved with the LHCb Run 13 dataset (23 fb-1 in 2025). Ultimately, a precision of the order of 27 is expected with the full HL-LHC LHCb dataset (300 fb-1 in 2038). The asymptotic sensitivity is dominated by the possibly large correlations of γ with respect to the nuisance parameters r()B and δ()B. The use of this method can improve our knowledge of γ from B0s decays and aid in understanding the discrepancy of γ between measurements with B+ and B0s modes.

      We are grateful to all members of the CKMfitter group for their comments and providing us with their private software based on a frequentist approach for computing the many pseudoexperiments performed for this study. In particular, we would like to thank J. Charles for his helpful comments on starting this analysis.

    Appendix A: Fitted nuisance parameters r()B and δ()B
    Appendix B: Case of γ equals 74
    Appendix C: Excluding B0s˜D()0(ππ)ϕ and B0s˜D()0(Kππ0)ϕ decays
    Appendix D: Other examples of 2-D p-value profiles for full HL-LHC LHCb dataset
    Appendix E: Half of 68.3% CL intervals of 1-D p-value profiles of γ
    Appendix F: Excluding B0s˜D0ϕ decays
    Reference (59)

目录

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return