This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Wave-Packet Effects: A Solution for Isospin Anomalies
in Vector-Meson Decay

Kenzo Ishikawa,1,2, Osamu Jinnouchi,3, Kenji Nishiwaki,4, and Kin-ya Oda5,

1Department of Physics, Faculty of Science, Hokkaido University,
Sapporo 060-0810, Japan
2Natural Science Center, Keio University, Yokohama 223-8521, Japan
3Department of Physics, Faculty of Science, Tokyo Institute of Technology,
Tokyo 152-8550, Japan
4Department of Physics, Shiv Nadar Institution of Eminence,
Gautam Buddha Nagar, 201314, India
5Department of Mathematics, Tokyo Woman’s Christian University,
Tokyo 167-8585, Japan
E-mail: ishikawa@particle.sci.hokudai.ac.jpE-mail: jinnouchi@phys.titech.ac.jpE-mail: kenji.nishiwaki@snu.edu.inE-mail: odakin@lab.twcu.ac.jp
Abstract

There is a long-standing anomaly in the ratio of the decay width for ψ(3770)D0D0¯\psi(3770)\to D^{0}\overline{D^{0}} to that for ψ(3770)D+D\psi(3770)\to D^{+}D^{-} at the level of 9.5σ9.5\,\sigma. A similar anomaly exists for the ratio of ϕ(1020)KL0KS0\phi(1020)\to K_{\text{L}}^{0}K_{\text{S}}^{0} to ϕ(1020)K+K\phi(1020)\to K^{+}K^{-} at 2.1σ2.1\,\sigma. In this study, we reassess the anomaly through the lens of Gaussian wave-packet formalism. Our comprehensive calculations include the localization of the overlap of the wave packets near the mass thresholds as well as the composite nature of the initial-state vector mesons. The results align within 1σ\sim 1\sigma confidence level with the Particle Data Group’s central values for a physically reasonable value of the form-factor parameter, indicating a resolution to these anomalies. We also check the deviation of a wave-packet resonance from the Briet-Wigner shape and find that wide ranges of the wave-packet size are consistent with the experimental data.

1 Introduction

There is a long-standing anomaly (discrepancy between experimental and theoretical results) in the ratio of the decay width for ψ(3770)D0D0¯\psi(3770)\to D^{0}\overline{D^{0}} to that for ψ(3770)D+D\psi(3770)\to D^{+}D^{-}. A similar but weaker anomaly exists for the ratio of ϕ(1020)KL0KS0\phi(1020)\to K_{\text{L}}^{0}K_{\text{S}}^{0} to ϕ(1020)K+K\phi(1020)\to K^{+}K^{-}. On the other hand, the ratio of Υ(4S)B+B\Upsilon(4S)\to B^{+}B^{-} to Υ(4S)B0B0¯\Upsilon(4S)\to B^{0}\overline{B^{0}} is consistent with the standard theoretical predictions.

At the quark level, these processes are111 Here and hereafter, we omit (3770)(3770), (1020)(1020), and (4S)(4S). Υ(4S)\Upsilon(4S) is sometimes written as Υ(10580)\Upsilon(10580). We do not distinguish the weak-interaction eigenstates K0K0¯K^{0}\overline{K^{0}} and the mass eigenstates KL0KS0K^{0}_{\text{L}}K^{0}_{\text{S}}, neglecting the small CPCP violating effects. Other processes have even smaller CPCP violating effects and we neglect them too.

ϕ(ss¯)\displaystyle\phi\!\left(s\overline{s}\right) K+(us¯)K(su¯),\displaystyle\to K^{+}\!\left(u\bar{s}\right)K^{-}\!\left(s\bar{u}\right), ψ(cc¯)\displaystyle\psi\!\left(c\overline{c}\right) D+(cd¯)D(dc¯),\displaystyle\to D^{+}\!\left(c\overline{d}\right)D^{-}\!\left(d\overline{c}\right), Υ(bb¯)\displaystyle\Upsilon\!\left(b\overline{b}\right) B+(ub¯)B(bu¯),\displaystyle\to B^{+}\!\left(u\overline{b}\right)B^{-}\!\left(b\bar{u}\right),
ϕ(ss¯)\displaystyle\phi\!\left(s\overline{s}\right) K0(ds¯)K0¯(sd¯)KL0KS0,\displaystyle\to K^{0}\!\left(d\bar{s}\right)\overline{K^{0}}\!\left(s\bar{d}\right)\to K^{0}_{\text{L}}K^{0}_{\text{S}}, ψ(cc¯)\displaystyle\psi\!\left(c\overline{c}\right) D0(cu¯)D0¯(uc¯),\displaystyle\to D^{0}\!\left(c\overline{u}\right)\overline{D^{0}}\!\left(u\overline{c}\right), Υ(bb¯)\displaystyle\Upsilon\!\left(b\overline{b}\right) B0(db¯)B0¯(bd¯),\displaystyle\to B^{0}\!\left(d\overline{b}\right)\overline{B^{0}}\!\left(b\overline{d}\right),

which can be summarized as V(QQ¯)P(Qq¯)+P¯(qQ¯)V\!\left(Q\overline{Q}\right)\to P\!\left(Q\overline{q}\right)+\overline{P}\!\left(q\overline{Q}\right), where VV and PP are vector and pseudo-scalar mesons, respectively, and QQ and qq are heavy (ss, cc, bb) and light (uu, dd) quarks, respectively. This ratio of decay widths is theoretically clean because most of the quantum-chromodynamics (QCD) corrections cancel out between the numerator and the denominator. These decay processes are via strong interaction, and hence in the limit of exact isospin symmetry udu\leftrightarrow d, the ratio becomes unity. The isospin violation makes a deviation from unity.

We name the ratio of the widths as222 In the original Ref. [1], the first two of Eq. (2) are given in its inverse (Rϕ1)PDG\displaystyle\left(R_{\phi}^{-1}\right){}^{\text{PDG}} =0.690±0.015,\displaystyle=0.690\pm 0.015, (Rψ1)PDG\displaystyle\left(R_{\psi}^{-1}\right){}^{\text{PDG}} =1.253±0.016,\displaystyle=1.253\pm 0.016, and we have inverted them in Eq. (2). In the theoretical literature, the ratio (1) of charged to neutral modes is mainly used, and we follow it for ease of comparison.

Rϕ\displaystyle R_{\phi} :=Γ(ϕK+K)Γ(ϕKL0KS0),\displaystyle:={\Gamma\!\left(\phi\to K^{+}K^{-}\right)\over\Gamma\!\left(\phi\to K_{\text{L}}^{0}K_{\text{S}}^{0}\right)}, Rψ\displaystyle R_{\psi} :=Γ(ψD+D)Γ(ψD0D0¯),\displaystyle:={\Gamma\!\left(\psi\to D^{+}D^{-}\right)\over\Gamma\!\left(\psi\to D^{0}\overline{D^{0}}\right)}, RΥ\displaystyle R_{\Upsilon} :=Γ(ΥB+B)Γ(ΥB0B0¯).\displaystyle:={\Gamma\!\left(\Upsilon\to B^{+}B^{-}\right)\over\Gamma\!\left(\Upsilon\to B^{0}\overline{B^{0}}\right)}. (1)

The experimental results are combined by the Particle Data Group (PDG) [1]:333 In the evaluation of RΥ=1.058±0.024R_{\Upsilon}=1.058\pm 0.024 [1], the following isospin symmetry among, e.g., B0J/ψKSB^{0}\to J/\psi\,K_{\text{S}} and B+J/ψK+B^{+}\to J/\psi\,K^{+} is assumed. The extent to which the isospin symmetry is valid in hadronic decays is debatable [Private communication with Dr. Akimasa Ishikawa].

RϕPDG\displaystyle R_{\phi}^{\text{PDG}} =1.45±0.03,\displaystyle=1.45\pm 0.03, RψPDG\displaystyle R_{\psi}^{\text{PDG}} =0.798±0.010,\displaystyle=0.798\pm 0.010, RΥPDG\displaystyle R_{\Upsilon}^{\text{PDG}} =1.058±0.024.\displaystyle=1.058\pm 0.024. (2)

The theoretical prediction of the decay rates for VP+PV\to P^{+}P^{-} and VP0P0¯V\to P^{0}\overline{P^{0}} is based on the plane-wave formalism so far. The tree-level result of the chiral perturbation theory reads

RVplane=gV+2gV02(mV24mP+2mV24mP02)3/2,\displaystyle R_{V}^{\text{plane}}={g_{V+}^{2}\over g_{V0}^{2}}\left(m_{V}^{2}-4m_{P^{+}}^{2}\over m_{V}^{2}-4m_{P^{0}}^{2}\right)^{3/2}, (3)

where gV+g_{V+} (gV0g_{V0}) is the coupling between VV and P+PP^{+}P^{-} (P0P0¯P^{0}\overline{P^{0}}) and mVm_{V}, mP+m_{P^{+}}, and mP0m_{P^{0}} are the masses of VV, P+P^{+}, and P0P^{0}, respectively. Even if we assume an isospin-symmetric coupling gV+=gV0g_{V+}=g_{V0}, the difference in the pseudo-scalar-meson masses mP+mP0m_{P^{+}}\neq m_{P^{0}} results in a deviation in RVR_{V} from unity: Putting the mass values in Ref. [1],444 Concretely, mϕ\displaystyle m_{\phi} =(1019.461±0.016)MeV,\displaystyle=\left(1019.461\pm 0.016\right)\text{MeV}, mψ\displaystyle m_{\psi} =(3773.7±0.4)MeV,\displaystyle=\left(3773.7\pm 0.4\right)\text{MeV}, mΥ\displaystyle m_{\Upsilon} =(10579.4±1.2)MeV,\displaystyle=\left(10579.4\pm 1.2\right)\text{MeV}, 2mK+\displaystyle 2m_{K^{+}} =(987.354±0.032)MeV,\displaystyle=\left(987.354\pm 0.032\right)\text{MeV}, 2mD+\displaystyle 2m_{D^{+}} =(3739.32±0.10)MeV,\displaystyle=\left(3739.32\pm 0.10\right)\text{MeV}, 2mB+\displaystyle 2m_{B^{+}} =(10558.7±0.24)MeV,\displaystyle=\left(10558.7\pm 0.24\right)\text{MeV}, 2mK0\displaystyle 2m_{K^{0}} =(995.222±0.026)MeV,\displaystyle=\left(995.222\pm 0.026\right)\text{MeV}, 2mD0\displaystyle 2m_{D^{0}} =(3729.68±0.10)MeV,\displaystyle=\left(3729.68\pm 0.10\right)\text{MeV}, 2mB0\displaystyle 2m_{B^{0}} =(10559.3±0.24)MeV,\displaystyle=\left(10559.3\pm 0.24\right)\text{MeV}, assuming the standard error propagation for the twice pseudo-scalar mass. The total decay widths are Γϕ\displaystyle\Gamma_{\phi} =(4.249±0.013)MeV,\displaystyle=\left(4.249\pm 0.013\right)\text{MeV}, Γψ\displaystyle\Gamma_{\psi} =(27.2±1.0)MeV,\displaystyle=\left(27.2\pm 1.0\right)\text{MeV}, ΓΥ\displaystyle\Gamma_{\Upsilon} =(20.5±2.5)MeV.\displaystyle=\left(20.5\pm 2.5\right)\text{MeV}. we obtain

Rϕplane\displaystyle R_{\phi}^{\text{plane}} =gϕ+2gϕ02(1.5156±0.0033),\displaystyle={g_{\phi+}^{2}\over g_{\phi 0}^{2}}\left(1.5156\pm 0.0033\right), Rψplane\displaystyle R_{\psi}^{\text{plane}} =gψ+2gψ02(0.6915±0.0046),\displaystyle={g_{\psi+}^{2}\over g_{\psi 0}^{2}}\left(0.6915\pm 0.0046\right), RΥplane\displaystyle R_{\Upsilon}^{\text{plane}} =gΥ+2gΥ02(1.047±0.026).\displaystyle={g_{\Upsilon+}^{2}\over g_{\Upsilon 0}^{2}}\left(1.047\pm 0.026\right). (4)

Comparing Eqs. (2) and (4), we see that the isospin-symmetric limit for the coupling gV+=gV0g_{V+}=g_{V0} results in the anomaly at the level of 2.1σ2.1\,\sigma, 9.5σ9.5\,\sigma, and 0.32σ0.32\,\sigma for ϕ\phi, ψ\psi, and Υ\Upsilon, respectively.

We briefly review the theoretical accounts for the anomaly within plane-wave formalism. For RϕR_{\phi}, it turned out that radiative corrections make the anomaly more significant [2]: The standard quantum-electrodynamics (QED) corrections make the theoretical prediction of the ratio 4 % larger, and isospin-breaking corrections to the ratio gϕ+2/gϕ02g^{2}_{\phi+}/g^{2}_{\phi 0} further make it “some 2 %” [2] larger, leading to a larger anomaly of roughly 5.2σ5.2\,\sigma assuming that the error is dominated by that in Eq. (2). In Ref. [3] the authors introduce a smeared decay rate that is a function of the energy difference between the initial and final plane-wave states; this smearing is by the Lorentzian distribution due to the inclusion of the width as well as by a phenomenological form factor put by hand to regularize an ultraviolet (UV) divergence; the anomaly for ϕ\phi can be explained with a mass parameter M1.5GeVM\simeq 1.5\,\text{GeV} in the phenomenological form factor. In Ref. [4], the authors have estimated the effects of the electromagnetic structure of kaons and other model-dependent contributions to the radiative corrections, and the resultant corrections have turned out to be tiny. In Ref. [5], two (a Breit-Wigner and a non-relativistic Lorentzian) types of averaged decay widths over the initial-state energy are introduced with two phenomenologically chosen energy intervals 1.010–1.060 GeV and 1.000–1.100 GeV, to relax the anomaly.

For RψR_{\psi}, another type of averaged decay width is introduced in Ref. [6], and the resultant anomaly has become even more significant. There is no explanation for this 9.5σ9.5\,\sigma anomaly so far.

The above smearing/averaging over the energy provides significant effects because the decay VPP¯V\to P\overline{P} is near the threshold mV2mPm_{V}\simeq 2m_{P}. In situations near the threshold, it is desirable to treat the decay more rigorously by using wave packets for the initial and final states. Recall that the SS-matrix in the plane-wave formalism contains the energy-momentum-conserving delta function and is theoretically ill-defined when computing the probability rather than the rate. A well-defined decay probability can be calculated only as a transition from a wave packet to a pair of wave packets. This is theoretically more reliable.

In the previous analyses [3, 4, 5], it has been assumed that the transition processes are described by the (plane-wave) rates alone.555 See also Refs. [7, 8] for non-standard approaches within the plane-wave formalisms. In this paper, we present an analysis based on the transition probability of the normalized states, wave packets, without the divergence of the delta-function squared. Concretely, we compute the decay VPP¯V\to P\overline{P} in the Gaussian wave-packet formalism [9, 10, 11, 12, 13]; see also Refs. [14, 15, 16].666 There is an ongoing experimental project directly to confirm this wave-packet effect [17, 18]; see also Ref. [19]. In particular, we include a wave packet effect, called the in-time-boundary effect for the decay, by simply limiting the time-integral of the decay-interaction point to t>Tint>T_{\text{in}} [10]. Here, TinT_{\text{in}} is the time from which the interaction is switched on. This procedure is proven to provide approximate modeling of the full production process of VV in the corresponding two-to-two wave-packet scattering, say, e+eVPP¯e^{+}e^{-}\to V\to P\overline{P} [13]; see also Refs. [20, 21, 11, 12, 22, 23, 24] for related discussions.

The organization of this paper is as follows: In Section 2, we will introduce the minimum basics of calculating the (generalized) SS-matrix that describes wave-packet-to-wave-packet transitions considering the initial state’s decaying nature when wave packets take the Gaussian form. In Section 3, we will review significant properties of the Gaussian wave-packet SS-matrix. In Section 4, we will compare the theoretical predictions for the ratios of RϕR_{\phi}, RψR_{\psi}, and RΥR_{\Upsilon} in the wave-packet and the plane-wave formalisms taking into account the form factor of the vector mesons. In Section 5, we will discuss the constraint from the resonant shape in the electron-positron-collider experiments for ϕ\phi and ψ\psi. In Section 6, we will provide a summary and further discussions. In Appendix A, we will review the form-factor details for vector mesons used for our analysis. In Appendix B, we will provide the details on how to derive the total probability of VPP¯V\to P\overline{P} under non-relativistic approximations. In Appendix C, a brief review on how to derive the plane-wave decay rate for VPP¯V\to P\overline{P} will be provided. In Appendix D, we will comment on a specific formal limit where the wave-packet decay rate coincides with the plane-wave decay rate. In Appendix E, we will briefly consider the isospin violation on the ρ\rho system.

2 Basics of Gaussian wave-packet formalism

For the near-threshold decay, the velocities in the final state are small, and the overlap of the wave packets becomes more significant in general. Therefore, it is important to take them into account.

Here, we spell out how to compute the probability for the VPP¯V\to P\overline{P} decay in the Gaussian wave-packet formalism. Throughout this paper, we work in the natural units =c=1\hbar=c=1. Readers who are more interested in analyses of experimental results rather than detailed theoretical formulation may skim through this section.

2.1 Wave-packet SS-matrix

In the Gaussian wave-packet formalism, a transition from an initial wave-packet state |𝒲𝒫0\ket{\mathcal{WP}_{0}} to a two-body final wave-packet state |𝒲𝒫1,𝒲𝒫2\ket{\mathcal{WP}_{1},\mathcal{WP}_{2}} is characterized by the following generalized SS-matrix [9]:

S𝒲𝒫0𝒲𝒫1𝒲𝒫2\displaystyle S_{\mathcal{WP}_{0}\to\mathcal{WP}_{1}\mathcal{WP}_{2}} =𝒲𝒫1,𝒲𝒫2|U^(Tout,Tin)|𝒲𝒫0,\displaystyle=\braket{\mathcal{WP}_{1},\mathcal{WP}_{2}}{\,\widehat{U}\!\left(T_{\text{out}},T_{\text{in}}\right)}{\mathcal{WP}_{0}}, (5)

where U^\widehat{U} describes the unitary time evolution from the initial time TinT_{\text{in}} to the final time ToutT_{\text{out}}

U^(Tout,Tin):=Texp(iTinToutdtd3𝒙^int(I)(t,𝒙)),\displaystyle\widehat{U}\!\left(T_{\text{out}},T_{\text{in}}\right):=\text{T}\,\text{exp}\!\left(-i\int_{T_{\text{in}}}^{T_{\text{out}}}\text{d}t\,\int\text{d}^{3}\bm{x}\,\widehat{{\cal H}}_{\text{int}}^{\text{(I)}}\!\left(t,\bm{x}\right)\right), (6)

in which T denotes the time-ordering and ^int(I)\widehat{{\cal H}}_{\text{int}}^{\text{(I)}} is the interaction Hamiltonian density in the interaction picture. The local interaction point (t,𝒙)\left(t,\bm{x}\right) is integrated in the four-dimensional spacetime. It is noteworthy that the wave-packet states |𝒲𝒫0\ket{\mathcal{WP}_{0}} and |𝒲𝒫1,𝒲𝒫2\ket{\mathcal{WP}_{1},\mathcal{WP}_{2}} are normalizable and hence the transition amplitude (5) is finite, unlike in the ordinary plane-wave formalism.777 See “Discussion” subsection in Ref. [13] for further discussion. Through the Dyson series expansion of U^(Tout,Tin)\widehat{U}\!\left(T_{\text{out}},T_{\text{in}}\right), a perturbative SS-matrix can be systematically constructed at any order of perturbation using Wick’s theorem, as in the plane wave case [11]. Throughout this paper, the subscripts 0, 1, and 2 denote VV, PP, and P¯\overline{P}, respectively.

A free Gaussian wave packet is characterized by a set of parameters {m,σ,X0,𝑿,𝑷}\Set{m,\sigma,X^{0},\bm{X},\bm{P}}, where mm is the mass; σ\sigma is the width-squared; and X0X^{0} is a reference time at which the wave packet takes the Gaussian form with the central values of the peak position 𝑿\bm{X} and momentum 𝑷\bm{P}.

Within the chiral perturbation theory, the effective-interaction-Hamiltonian density is

^int,eff(I)=igV+𝒱μ[𝒫+μ𝒫𝒫μ𝒫+]+igV0𝒱μ[𝒫0μ𝒫0¯𝒫0¯μ𝒫0],\displaystyle\widehat{{\cal H}}^{\text{(I)}}_{\text{int,eff}}=ig_{V+}{\cal V}^{\mu}\left[{\cal P}^{+}\partial_{\mu}{\cal P}^{-}-{\cal P}^{-}\partial_{\mu}{\cal P}^{+}\right]+ig_{V0}{\cal V}^{\mu}\left[{\cal P}^{0}\partial_{\mu}\overline{{\cal P}^{0}}-\overline{{\cal P}^{0}}\partial_{\mu}{\cal P}^{0}\right], (7)

where 𝒱\cal V, 𝒫±\cal P^{\pm}, 𝒫0{\cal P}^{0}, and 𝒫0¯\overline{{\cal P}^{0}} are the fields representing the vector meson, the charged pseudo-scalar mesons, the neutral pseudo-scalar meson, and its antiparticle, respectively, and gV+g_{V+} and gV0g_{V0} are the vector-meson effective couplings to the charged pseudo-scalars and to the neutral pseudo-scalars, respectively. In this paper, we take the isospin-symmetric limit,

gV+=gV0(=:gV),\displaystyle g_{V+}=g_{V0}\quad(=:g_{V}), (8)

with which the effective coupling geffg_{\text{eff}} takes the form in the momentum space

geff(λ0,P0,P1,P2):=gVεμ(P0,λ0)(P1μP2μ),\displaystyle g_{\text{eff}}\!\left(\lambda_{0},P_{0},P_{1},P_{2}\right):=g_{V}\,\varepsilon_{\mu}\!\left(P_{0},\lambda_{0}\right)\left(P_{1}^{\mu}-P_{2}^{\mu}\right), (9)

where P0P_{0}, P1P_{1}, and P2P_{2} are the four-momenta of the vector meson VV, the pseudo-scalar meson PP, and its antiparticle P¯\overline{P}, respectively, and εμ\varepsilon_{\mu} is the polarization vector of the vector meson with λ0\lambda_{0} being its helicity.888 In Eq. (9), P0=(P0μ)μ=0,,3=(P00,𝑷0)P_{0}=\left(P_{0}^{\mu}\right)_{\mu=0,\dots,3}=\left(P_{0}^{0},\bm{P}_{0}\right) stands for the four-momentum of VV, with its subscript denoting the initial particle. We use the same letter P0P^{0} for the particle label of the neutral pseudo-scalar, with its zero denoting its charge. The distinction should be apparent from the context.

In this paper, we investigate the transition from an off-shell initial state for VV to an on-shell final state for PP¯P\overline{P}, having an off-shell energy E~0\widetilde{E}_{0} and on-shell ones E1E_{1}, E2E_{2}, respectively:999 This procedure of introducing E~0\widetilde{E}_{0} is equivalent to the Weisskopf-Wigner approximation [25, 26]. See Ref. [27] for its inclusion in the Gaussian wave-packet formalism.

E~0\displaystyle\widetilde{E}_{0} :=mV2+𝑷02iΓVmV=E02iΓVmVE0imV2E0ΓV,\displaystyle:=\sqrt{m_{V}^{2}+\bm{P}_{0}^{2}-i\,\Gamma_{V}m_{V}}=\sqrt{E_{0}^{2}-i\,\Gamma_{V}m_{V}}\simeq E_{0}-i\frac{m_{V}}{2E_{0}}\Gamma_{V},
E1\displaystyle E_{1} :=mP2+𝑷12,\displaystyle:=\sqrt{m_{P}^{2}+\bm{P}_{1}^{2}},
E2\displaystyle E_{2} :=mP2+𝑷22,\displaystyle:=\sqrt{m_{P}^{2}+\bm{P}_{2}^{2}}, (10)

where mVm_{V} and mPm_{P} are the masses of VV and PP, respectively; E0:=mV2+𝑷02E_{0}:=\sqrt{m_{V}^{2}+\bm{P}_{0}^{2}} is the on-shell energy of VV; and ΓV\Gamma_{V} is the “decay width” of VV, or more precisely, the imaginary part of its plane-wave propagator divided by mVm_{V}; see Ref. [13] for detailed discussion; see also footnote 13. Throughout this paper, we take the narrow-width approximation for ΓV\Gamma_{V} as in Eq. (10).101010 Theoretically, ΓV\Gamma_{V} is obtained as the imaginary part of the plane-wave VV propagator at the loop level. See footnote 13. Here, the off-shell VV should eventually be regarded as an intermediate state for a scattering process that includes the production of VV, which necessarily introduces the in-time-boundary effect appearing below.

Their wave functions take the form

fV(x)\displaystyle f_{V}\!\left(x\right) :=NV(σ0π)3/4(πσ0)3/212P00(2π)3/2eiP0(xX0)(𝒙𝚵0(t))22σ0|P00=E~0,\displaystyle:=N_{V}\left(\frac{\sigma_{0}}{\pi}\right)^{3/4}\left(\frac{\pi}{\sigma_{0}}\right)^{3/2}\frac{1}{\sqrt{2P^{0}_{0}}\left(2\pi\right)^{3/2}}e^{iP_{0}\cdot(x-X_{0})-\frac{\left(\bm{x}-\bm{\Xi}_{0}\!\left(t\right)\right)^{2}}{2\sigma_{0}}}\Bigg{|}_{P_{0}^{0}=\widetilde{E}_{0}},
fP(x)\displaystyle f_{P}\!\left(x\right) :=(σ1π)3/4(πσ1)3/212P10(2π)3/2eiP1(xX1)(𝒙𝚵1(t))22σ1|P10=E1,\displaystyle:=\phantom{N_{V}}\left(\frac{\sigma_{1}}{\pi}\right)^{3/4}\left(\frac{\pi}{\sigma_{1}}\right)^{3/2}\frac{1}{\sqrt{2P^{0}_{1}}\left(2\pi\right)^{3/2}}e^{iP_{1}\cdot(x-X_{1})-\frac{\left(\bm{x}-\bm{\Xi}_{1}\!\left(t\right)\right)^{2}}{2\sigma_{1}}}\Bigg{|}_{P^{0}_{1}=E_{1}},
fP¯(x)\displaystyle f_{\overline{P}}\!\left(x\right) :=(σ2π)3/4(πσ2)3/212P20(2π)3/2eiP2(xX2)(𝒙𝚵2(t))22σ2|P20=E2,\displaystyle:=\phantom{N_{V}}\left(\frac{\sigma_{2}}{\pi}\right)^{3/4}\left(\frac{\pi}{\sigma_{2}}\right)^{3/2}\frac{1}{\sqrt{2P^{0}_{2}}\left(2\pi\right)^{3/2}}e^{iP_{2}\cdot(x-X_{2})-\frac{\left(\bm{x}-\bm{\Xi}_{2}\!\left(t\right)\right)^{2}}{2\sigma_{2}}}\Bigg{|}_{P^{0}_{2}=E_{2}}, (11)

where NVN_{V} is a wave-function (field) renormalization factor for VV due to its offshellness111111 NVN_{V} shows a factor that accounts for the possible extra decrease of the norm of the initial state due to the off-shellness ΓV>0\Gamma_{V}>0. Anyway, NVN_{V} will drop out of the final ratio of the decay probabilities. and

𝚵A(t):=𝑿A+𝑽A(tXA0)(A=0,1,2),\displaystyle\bm{\Xi}_{A}\!\left(t\right):=\bm{X}_{A}+\bm{V}_{A}\left(t-X^{0}_{A}\right)\qquad(A=0,1,2), (12)

describes the location of the center of the wave packet at a time tt with the central velocity

𝑽A:=𝑷AEA.\displaystyle\bm{V}_{A}:=\frac{\bm{P}_{A}}{E_{A}}. (13)

Now, it is straightforward to compute the SS-matrix from Eq. (5) at the leading order in the Dyson series (6) with the effective Hamiltonian (7) from the wave functions (2.1[11]:

SVPP¯\displaystyle S_{V\to P\overline{P}} =igeff(λ0,P0,P1,P2)(A=0212EA(1πσA)3/4)eσt2(δω)2σs2(δ𝑷)22\displaystyle=ig_{\text{eff}}\!\left(\lambda_{0},P_{0},P_{1},P_{2}\right)\left(\prod_{A=0}^{2}\frac{1}{\sqrt{2E_{A}}}\left(\frac{1}{\pi\sigma_{A}}\right)^{3/4}\right)e^{-\frac{\sigma_{t}}{2}(\delta\omega)^{2}-\frac{\sigma_{s}}{2}(\delta\bm{P})^{2}-\frac{\cal R}{2}}
×TinToutdte12σt[t(𝔗+iσtδω)]2d3𝒙e12σs[𝒙(𝖃¯+𝑽¯tiσsδ𝑷)]2\displaystyle\quad\times\int_{T_{\text{in}}}^{T_{\text{out}}}\text{d}t\,e^{-\frac{1}{2\sigma_{t}}\left[t-\left(\mathfrak{T}+i\sigma_{t}\delta\omega\right)\right]^{2}}\int\text{d}^{3}\bm{x}\,e^{-\frac{1}{2\sigma_{s}}\left[\bm{x}-\left(\overline{\bm{\mathfrak{X}}}+\overline{\bm{V}}t-i\sigma_{s}\delta\bm{P}\right)\right]^{2}}
×NVeΓV2(tT0)F~(|𝑽1𝑽2|),\displaystyle\quad\times N_{V}\,e^{-{\Gamma_{V}\over 2}\left(t-{T_{0}}\right)}\widetilde{F}\!\left(\left|\bm{V}_{1}-\bm{V}_{2}\right|\right), (14)

where the notation follows Eq. (27)(27) of Ref. [11] (see also below for a short summary).121212 In Eq. (14), we have dropped the overall phase factor which is irrelevant to the calculation of the probability, while properly taking into account the real damping factor eΓV2(tT0)e^{-{\Gamma_{V}\over 2}\left(t-T_{0}\right)} coming from the imaginary part of E~0\widetilde{E}_{0}; see Eq. (2.1). Differences from the previous calculation [11] are the following four points: First, the coupling is changed to κ/2geff\kappa/\sqrt{2}\to g_{\text{eff}}. Second, the “decay width” of VV is included as the phenomenological factor eΓV2(tT0)e^{-{\Gamma_{V}\over 2}\left(t-{T_{0}}\right)},131313 When one includes the production process in the amplitude, e.g., as e+eVPP¯e^{+}e^{-}\to V\to P\overline{P}, the result does not change whether we expand the complete set of intermediate states of VV by the Gaussian-wave-packet or plane-wave bases; see Sec. 2.3 in Ref. [12]. The imaginary part mVΓVm_{V}\Gamma_{V} of the plane-wave propagator of VV appears through loop corrections, and when translated to the decay process VPP¯V\to P\overline{P}, its effect can be expressed as the phenomenological factor eΓV(tT0)/2e^{-\Gamma_{V}\left(t-T_{0}\right)/2} in the plane-wave formalism. Here, we also phenomenologically take into account the exponentially decaying nature of the initial wave packet of VV through the channel that is common to the plane-wave decay, namely, through the bulk effect that appears below. See also footnote 9. where T0:=X00T_{0}:=X^{0}_{0} is the initial time from which VV starts to exist.141414 T0T_{0} is indeed irrelevant in the sense that the time-translational invariance results in the dependence of the final result only on the difference TinT0T_{\text{in}}-T_{0}. Furthermore, this dependence on TinT0T_{\text{in}}-T_{0} cancels out between the numerator and denominator of the final ratio of the decay probabilities, as we will see. (Physically, we would expect TinT0T_{\text{in}}\simeq T_{0}.) Third, NVN_{V} in Eq. (2.1) is introduced. Fourth, we have included a phenomenological form factor F~\widetilde{F} due to the composite nature of VV:

F~(|𝑽1𝑽2|):=11+(R0mP|𝑽1𝑽2|2)2,\displaystyle\widetilde{F}\!\bigl{(}\left|\bm{V}_{1}-\bm{V}_{2}\right|\bigr{)}:=\frac{1}{1+\left(\frac{R_{0}m_{P}\left|\bm{V}_{1}-\bm{V}_{2}\right|}{2}\right)^{2}}, (15)

where R0R_{0} describes a typical length scale of the compositeness of VV; see Appendix A. The normalization is such that F~\widetilde{F} becomes unity for 𝑽1=𝑽2\bm{V}_{1}=\bm{V}_{2}.

Refer to caption
Refer to caption
Figure 1: Schematic figure for the finite wave-packet process (left) and the infinite plane-wave process (right), without taking into account the decay width ΓV\Gamma_{V}. In the left, we have shown the time of intersection 𝔗\mathfrak{T}; the spatial and temporal sizes of the overlap σs\sqrt{\sigma_{s}} and σt\sqrt{\sigma_{t}}, respectively; the center of wave packets 𝚵A\bm{\Xi}_{A} (A=0,1,2A=0,1,2); and the initial and final times of the scattering TinT_{\text{in}} and ToutT_{\text{out}}, respectively. Also, the bulk TintToutT_{\text{in}}\ll t\ll T_{\text{out}}, in-time-boundary (|tTin|σt\left|t-T_{\text{in}}\right|\lesssim\sqrt{\sigma_{t}}), and out-time-boundary (|Toutt|σt\left|T_{\text{out}}-t\right|\lesssim\sqrt{\sigma_{t}}) regions are shown. (This panel corresponds to the bulk-like case |𝔗Tin|σt\left|\mathfrak{T}-T_{\text{in}}\right|\gg\sqrt{\sigma_{t}}; see Fig. 2.) In the right, the spatial overlap of the plane waves never decreases in time, and hence the interaction would be never switched off, and the scattering would be never completed; therefore the extra damping factor eϵtd3𝒙^int(I)(t,𝒙)e^{\mp\epsilon t\int\text{d}^{3}\bm{x}\,\widehat{{\cal H}}_{\text{int}}^{\text{(I)}}\!\left(t,\bm{x}\right)} with an infinitesimal ϵ>0\epsilon>0 is conventionally put by hand for the future and past infinite times t±t\to\pm\infty, which is depicted by the damping of the opacity of the orange region. This factor eventually results in the propagator (p2+m2iϵ)1\propto\left(p^{2}+m^{2}-i\epsilon\right)^{-1} in the conventional Feynman diagram calculation.

Now we provide a brief introduction to other variables in the first two lines of (14) (see Section 3.1 of [11] for more details):

  • σs\sqrt{\sigma_{s}} is a typical spacial size of the region of interaction:

    σs1:=A=021σA.\displaystyle\sigma_{s}^{-1}:=\sum_{A=0}^{2}\frac{1}{\sigma_{A}}. (16)
  • σt\sqrt{\sigma_{t}} is a typical temporal size of the interaction region:151515 We adopt the following notation for arbitrary scalar and vector variables CC and 𝑪\bm{C}, respectively: C¯\displaystyle\overline{C} :=σsA=02CAσA,\displaystyle:=\sigma_{s}\sum_{A=0}^{2}\frac{C_{A}}{\sigma_{A}}, 𝑪¯\displaystyle\overline{\bm{C}} :=σsA=02𝑪AσA,\displaystyle:=\sigma_{s}\sum_{A=0}^{2}\frac{\bm{C}_{A}}{\sigma_{A}}, Δ𝑪2\displaystyle\Delta\bm{C}^{2} :=𝑪2¯𝑪¯2.\displaystyle:=\overline{\bm{C}^{2}}-{\overline{\bm{C}}}^{2}.

    σt:=σsΔ𝑽2.\displaystyle\sigma_{t}:=\frac{\sigma_{s}}{\Delta\bm{V}^{2}}. (17)
  • 𝔗\mathfrak{T} is the time of intersection of the three wave packets:

    𝔗:=σt𝑽¯𝖃¯𝑽𝖃¯σs,\displaystyle\mathfrak{T}:=\sigma_{t}\frac{\overline{\bm{V}}\cdot\overline{\bm{\mathfrak{X}}}-\overline{\bm{V}\cdot\bm{\mathfrak{X}}}}{\sigma_{s}}, (18)

    where

    𝖃A:=𝚵A(0)(=𝑿A𝑽AXA0)\displaystyle\bm{\mathfrak{X}}_{A}:={\bm{\Xi}}_{A}\!\left(0\right)\qquad\left(\,=\bm{X}_{A}-\bm{V}_{A}X_{A}^{0}\,\right) (19)

    is the location of the center of each wave packet at our reference time t=0t=0. As mentioned above, each wave packet takes the Gaussian form centered at 𝑿A\bm{X}_{A} at its reference time XA0X_{A}^{0}.

  • {\cal R} is called the overlap exponent, which provides the exponential suppression when wave packets are separated from each other:

    :=Δ𝖃2σs𝔗2σt.\displaystyle{\cal R}:=\frac{\Delta\bm{\mathfrak{X}}^{2}}{\sigma_{s}}-\frac{\mathfrak{T}^{2}}{\sigma_{t}}. (20)
  • We write the deviation of energy-momentum from the conserved values (for their central values of wave packets)

    δ𝑷\displaystyle\delta\bm{P} :=𝑷1+𝑷2𝑷0,\displaystyle:=\bm{P}_{1}+\bm{P}_{2}-\bm{P}_{0},
    δE\displaystyle\delta E :=E1+E2E0,\displaystyle:=E_{1}+E_{2}-E_{0}, δω\displaystyle\delta\omega :=δE𝑽¯δ𝑷,\displaystyle:=\delta E-\overline{\bm{V}}\cdot\delta\bm{P}, (21)

    where ωA:=EA𝑽¯𝑷A\omega_{A}:=E_{A}-\overline{\bm{V}}\cdot\bm{P}_{A} is the “shifted energy” of each packet.

A schematic figure is shown in the left panel of Fig. 1, compared with the plane-wave counterpart in the right.

After the square completion of tt and the analytic Gaussian integration over 𝒙\bm{x} in (14), as made in [11], we represent the SS-matrix as follows:

SVPP¯\displaystyle S_{V\to P\overline{P}} =igeffNV(A=0212EA(1πσA)3/4)eσt2(δω)2σs2(δ𝑷)22(2πσs)3/22πσtG(𝔗)\displaystyle=ig_{\text{eff}}N_{V}\left(\prod_{A=0}^{2}\frac{1}{\sqrt{2E_{A}}}\left(\frac{1}{\pi\sigma_{A}}\right)^{3/4}\right)e^{-\frac{\sigma_{t}}{2}(\delta\omega)^{2}-\frac{\sigma_{s}}{2}(\delta\bm{P})^{2}-\frac{\cal R}{2}}\left(2\pi\sigma_{s}\right)^{3/2}\sqrt{2\pi\sigma_{t}}\,G(\mathfrak{T})
×eΓV2(𝔗T0+iσtδω)+ΓV2σt8F~(|𝑽1𝑽2|),\displaystyle\quad\times e^{-{\Gamma_{V}\over 2}\left(\mathfrak{T}-{T_{0}}+i\sigma_{t}\delta\omega\right)+\frac{\Gamma_{V}^{2}\sigma_{t}}{8}}\,\widetilde{F}\!\left(\left|\bm{V}_{1}-\bm{V}_{2}\right|\right), (22)

where the window function G(𝔗)G(\mathfrak{T}) is defined as161616 In Eq. (23), 𝔗\mathfrak{T} on the right-hand side is replaced from the original definition of G(𝔗)G(\mathfrak{T}) in [11] as 𝔗𝔗ΓVσt2\mathfrak{T}\to\mathfrak{T}-{\Gamma_{V}\sigma_{t}\over 2}.

G(𝔗)\displaystyle G(\mathfrak{T}) :=TinToutdt2πσte12σt[t(𝔗ΓVσt2+iσtδω)]2\displaystyle:=\int_{T_{\text{in}}}^{T_{\text{out}}}\frac{\text{d}t}{\sqrt{2\pi\sigma_{t}}}e^{-\frac{1}{2\sigma_{t}}\left[t-\left(\mathfrak{T}-{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega\right)\right]^{2}}
=12[erf(𝔗TinΓVσt2+iσtδω2σt)erf(𝔗ToutΓVσt2+iσtδω2σt)],\displaystyle=\frac{1}{2}\left[\text{erf}\left(\frac{\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega}{\sqrt{2\sigma_{t}}}\right)-\text{erf}\left(\frac{\mathfrak{T}-T_{\text{out}}-{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega}{\sqrt{2\sigma_{t}}}\right)\right], (23)

with

erf(z):=2π0zex2dx\displaystyle\text{erf}(z):=\frac{2}{\sqrt{\pi}}\int_{0}^{z}e^{-x^{2}}\text{d}x (24)

being the Gauss error function. The window function G(𝔗)G\!\left(\mathfrak{T}\right) becomes unity for Tin𝔗ToutT_{\text{in}}\ll\mathfrak{T}\ll T_{\text{out}} and zero for 𝔗Tin\mathfrak{T}\ll T_{\text{in}} and for Tout𝔗T_{\text{out}}\ll\mathfrak{T}. For a given configuration of in and out states, which fixes the value of σt\sigma_{t}, the time regions TintToutT_{\text{in}}\ll t\ll T_{\text{out}},171717 More precisely, the bulk region is the one satisfying |tTin|σt\left|t-T_{\text{in}}\right|\gg\sqrt{\sigma_{t}} and |Toutt|σt\left|T_{\text{out}}-t\right|\gg\sqrt{\sigma_{t}}. |tTin|σt\left|t-T_{\text{in}}\right|\lesssim\sqrt{\sigma_{t}}, and |tTout|σt\left|t-T_{\text{out}}\right|\lesssim\sqrt{\sigma_{t}} are called the bulk, in-time-boundary, and out-time-boundary regions, respectively. In the phenomenological analysis below, we will neglect the out-time-boundary contributions as we will discuss.

2.2 Differential decay probability

From the SS-matrix (22), the differential decay probability can be derived as

dPVPP¯\displaystyle\text{d}P_{V\to P\overline{P}} =d3𝑿1d3𝑷1(2π)3d3𝑿2d3𝑷2(2π)3|SVPP¯|2\displaystyle=\frac{\text{d}^{3}\bm{X}_{1}\text{d}^{3}\bm{P}_{1}}{(2\pi)^{3}}\frac{\text{d}^{3}\bm{X}_{2}\text{d}^{3}\bm{P}_{2}}{(2\pi)^{3}}\left|S_{V\to P\overline{P}}\right|^{2}
=|geff|2¯NV212E0d3𝑷1(2π)32E1d3𝑷2(2π)32E2(2π)4(σtπeσt(δω)2)((σsπ)3/2eσs(δ𝑷)2)\displaystyle=\overline{|g_{\text{eff}}|^{2}}N_{V}^{2}\frac{1}{2E_{0}}\frac{\text{d}^{3}\bm{P}_{1}}{(2\pi)^{3}2E_{1}}\frac{\text{d}^{3}\bm{P}_{2}}{(2\pi)^{3}2E_{2}}(2\pi)^{4}\left(\sqrt{\frac{\sigma_{t}}{\pi}}e^{-\sigma_{t}\left(\delta\omega\right)^{2}}\right)\left(\left(\frac{\sigma_{s}}{\pi}\right)^{3/2}e^{-\sigma_{s}\left(\delta\bm{P}\right)^{2}}\right)
×σtπ5(σsσ0σ1σ2)3d3𝑿1d3𝑿2e|G(𝔗)|2eΓV(𝔗T0)+ΓV2σt4|F~(|𝑽1𝑽2|)|2,\displaystyle\quad\times\sqrt{\frac{\sigma_{t}}{\pi^{5}}\left(\frac{\sigma_{s}}{\sigma_{0}\sigma_{1}\sigma_{2}}\right)^{3}}\text{d}^{3}\bm{X}_{1}\text{d}^{3}\bm{X}_{2}e^{-\mathcal{R}}\,|G(\mathfrak{T})|^{2}\,e^{-\Gamma_{V}\left(\mathfrak{T}-{T_{0}}\right)+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}}\,\left|\widetilde{F}\!\left(\left|\bm{V}_{1}-\bm{V}_{2}\right|\right)\right|^{2}, (25)

where we have taken the average over the helicity λ0\lambda_{0}, which results in the helicity-averaged effective coupling:

|geff|2¯\displaystyle\overline{|g_{\text{eff}}|^{2}} :=gV23λ0|εμ(P0,λ0)(P1μP2μ)|2=gV23(𝑷1𝑷2)2.\displaystyle:=\frac{g_{V}^{2}}{3}\sum_{\lambda_{0}}\bigl{|}\varepsilon_{\mu}(P_{0},\lambda_{0})\left(P_{1}^{\mu}-P_{2}^{\mu}\right)\bigr{|}^{2}={\frac{g_{V}^{2}}{3}}(\bm{P}_{1}-\bm{P}_{2})^{2}. (26)

Here, the last equality further assumes the vanishing initial momentum 𝑷0=0\bm{P}_{0}=0. We will compute the integrated decay probability under this assumption in Sec. 2.3 and in Appendix B.

Refer to caption
Figure 2: Schematic figures for two limiting cases |𝔗Tin|σt\left|\mathfrak{T}-T_{\text{in}}\right|\gg\sqrt{\sigma_{t}} (left) and σtΓV1\sqrt{\sigma_{t}}\gg\Gamma_{V}^{-1} (right). The case δωσt1/2\delta\omega\gg\sigma_{t}^{-1/2} is hard to draw in the position space and is not shown here. The overlap region is determined by both the initial and final states as in Fig. 1.

Hereafter, we assume both the following conditions

|𝔗TinΓVσt2+iσtδω|2σt\displaystyle\frac{\left|\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega\right|}{\sqrt{2\sigma_{t}}} 1,\displaystyle\gg 1, |𝔗ToutΓVσt2+iσtδω|2σt\displaystyle\frac{\left|\mathfrak{T}-T_{\text{out}}-{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega\right|}{\sqrt{2\sigma_{t}}} 1.\displaystyle\gg 1. (27)

Physically, each of these conditions is satisfied when at least one of the following three conditions is met:

  • |𝔗Tin|σt\left|\mathfrak{T}-T_{\text{in}}\right|\gg\sqrt{\sigma_{t}} (or |𝔗Tout|σt\left|\mathfrak{T}-T_{\text{out}}\right|\gg\sqrt{\sigma_{t}}) when the interaction time 𝔗\mathfrak{T} is apart enough from TinT_{\text{in}} (or ToutT_{\text{out}}) compared to the temporal width of the overlap σt\sqrt{\sigma_{t}}, which typically corresponds to the “bulk-like” case (Fig. 2, left);

  • σtΓV1\sqrt{\sigma_{t}}\gg\Gamma_{V}^{-1} when the “mean lifetime through bulk effect” ΓV1\Gamma_{V}^{-1} is much shorter than the temporal width of the overlap region σt\sqrt{\sigma_{t}}, which typically corresponds to the so-to-say “decay within wave-packet overlap” case (Fig. 2, right);

  • δωσt1/2\delta\omega\gg\sigma_{t}^{-1/2} when the deviation from the conservation of the shifted-energy, δω\delta\omega, is much larger than the inverse of the temporal width of the overlap 1/σt1/\sqrt{\sigma_{t}}, namely the “violation of shifted-energy” case.

This assumption (27) is made for simplicity, and there is no obstacle to using the full form (23) in the numerical computation in principle, but the result would remain the same approximately because this is anyway satisfied in the ordinary bulk-like case as well as when anything interesting happens around the (in-)time-boundary.

Under the assumption (27), the following asymptotic form is obtained [11]:

G(𝔗)\displaystyle G(\mathfrak{T}) W(𝔗)12e(𝔗TinΓVσt2)22σt+σt2(δω)2iδω(𝔗TinΓVσt2)2σtπ1𝔗TinΓVσt2+iσtδω\displaystyle\simeq W(\mathfrak{T})-{\frac{1}{2}}e^{-\frac{\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)^{2}}{2\sigma_{t}}+\frac{\sigma_{t}}{2}(\delta\omega)^{2}-i\delta\omega\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)}\sqrt{\frac{2\sigma_{t}}{\pi}}\frac{1}{\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega}
+12e(𝔗ΓVσt2Tout)22σt+σt2(δω)2iδω(𝔗ΓVσt2Tout)2σtπ1𝔗ToutΓVσt2+iσtδω,\displaystyle\quad+{\frac{1}{2}}e^{-\frac{\left(\mathfrak{T}-{\Gamma_{V}\sigma_{t}\over 2}-T_{\text{out}}\right)^{2}}{2\sigma_{t}}+\frac{\sigma_{t}}{2}(\delta\omega)^{2}-i\delta\omega\left(\mathfrak{T}-{\Gamma_{V}\sigma_{t}\over 2}-T_{\text{out}}\right)}\sqrt{\frac{2\sigma_{t}}{\pi}}\frac{1}{\mathfrak{T}-T_{\text{out}}-{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega}, (28)

where we have defined the “bulk window function”

W(𝔗):=12[sgn(𝔗TinΓVσt2+iσtδω2σt)sgn(𝔗ToutΓVσt2+iσtδω2σt)],\displaystyle W(\mathfrak{T}):=\frac{1}{2}\left[\text{sgn}\left(\frac{\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega}{\sqrt{2\sigma_{t}}}\right)-\text{sgn}\left(\frac{\mathfrak{T}-T_{\text{out}}-{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega}{\sqrt{2\sigma_{t}}}\right)\right], (29)

in which the sign function for a complex variable is

sgn(z)\displaystyle\text{sgn}(z) :={+1for z>0 or (z=0 and z>0),1for z<0 or (z=0 and z<0),0for z=0.\displaystyle:=\begin{cases}+1&\text{for }\Re z>0\text{ or }(\Re z=0\text{ and }\Im z>0),\\ -1&\text{for }\Re z<0\text{ or }(\Re z=0\text{ and }\Im z<0),\\ 0&\text{for }z=0.\end{cases} (30)

Here and hereafter, \Re and \Im denote the real and imaginary parts, respectively. Equation (29) describes the ordinary “bulk contribution” for the quantum transition from the in to out states in the time period [Tin,Tout][T_{\text{in}},T_{\text{out}}]. The second and third terms of Eq. (28) show the contributions near the in and out time boundaries TinT_{\text{in}} and ToutT_{\text{out}}, respectively. More explicitly,181818 Precisely speaking, Eq. (31) is given except right at the boundary 𝔗=Tin+ΓVσt2\mathfrak{T}=T_{\text{in}}+{\Gamma_{V}\sigma_{t}\over 2} or 𝔗=Tout+ΓVσt2\mathfrak{T}=T_{\text{out}}+{\Gamma_{V}\sigma_{t}\over 2}, which is rather a peculiarity of how to define a boundary value and is out of our current interest.

W(𝔗)\displaystyle W(\mathfrak{T}) ={1(Tin+ΓVσt2<𝔗<Tout+ΓVσt2),0otherwise.\displaystyle=\begin{cases}1&(T_{\text{in}}+{\Gamma_{V}\sigma_{t}\over 2}<\mathfrak{T}<T_{\text{out}}+{\Gamma_{V}\sigma_{t}\over 2}),\\ 0&\text{otherwise}.\end{cases} (31)

Because the contribution from the out-time-boundary 𝔗Tout\mathfrak{T}\simeq T_{\text{out}} is suppressed by the extra dumping factor eΓV(𝔗T0)e^{-\Gamma_{V}\left(\mathfrak{T}-T_{0}\right)} in the differential probability (25),191919 Here, we physically assume ToutTinΓV1T_{\text{out}}-T_{\text{in}}\gg\Gamma_{V}^{-1} with T0TinT_{0}\simeq T_{\text{in}}; see also footnote 14. it is safe to neglect the out-time-boundary contribution, and we can take

Tout+,\displaystyle T_{\text{out}}\to+\infty, (32)

with which the second line in Eq. (28) goes down to zero. With this limit, |G(𝔗)|2\left|G(\mathfrak{T})\right|^{2} reads

|G(𝔗)|2\displaystyle\left|G(\mathfrak{T})\right|^{2} [𝒢𝒢]bulk(𝔗)+[𝒢𝒢]bdry(𝔗)+[𝒢𝒢]intf(𝔗),\displaystyle\to\left[\cal GG\right]_{\text{bulk}}(\mathfrak{T})+\left[\cal GG\right]_{\text{bdry}}(\mathfrak{T})+\left[\cal GG\right]_{\text{intf}}(\mathfrak{T}), (33)

where

[𝒢𝒢]bulk(𝔗)\displaystyle\left[\cal GG\right]_{\text{bulk}}(\mathfrak{T}) :=|W(𝔗)|2,\displaystyle:=\bigl{|}{W(\mathfrak{T})}\bigr{|}^{2}, (34)
[𝒢𝒢]bdry(𝔗)\displaystyle\left[\cal GG\right]_{\text{bdry}}(\mathfrak{T}) :=14e(𝔗TinΓVσt2)2σt+σt(δω)22σtπ1(𝔗TinΓVσt2)2+(σtδω)2,\displaystyle:=\frac{1}{4}e^{-\frac{\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)^{2}}{\sigma_{t}}+{\sigma_{t}}(\delta\omega)^{2}}\frac{2\sigma_{t}}{\pi}\frac{1}{\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)^{2}+\left(\sigma_{t}\delta\omega\right)^{2}}, (35)
[𝒢𝒢]intf(𝔗)\displaystyle\left[\cal GG\right]_{\text{intf}}(\mathfrak{T}) :=W(𝔗)2e(𝔗TinΓVσt2)22σt+σt2(δω)22σtπ\displaystyle:=-\frac{W\!\left(\mathfrak{T}\right)}{2}e^{-\frac{\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)^{2}}{2\sigma_{t}}+\frac{\sigma_{t}}{2}(\delta\omega)^{2}}\sqrt{\frac{2\sigma_{t}}{\pi}}
×{eiδω(𝔗TinΓVσt2)𝔗TinΓVσt2+iσtδω+e+iδω(𝔗TinΓVσt2)𝔗TinΓVσt2iσtδω}.\displaystyle\ \qquad\times\left\{\frac{e^{-i\delta\omega\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)}}{\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega}+\frac{e^{+i\delta\omega\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)}}{\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}-i\sigma_{t}\delta\omega}\right\}. (36)

The three functions [𝒢𝒢]bulk\left[\cal GG\right]_{\text{bulk}}, [𝒢𝒢]bdry\left[\cal GG\right]_{\text{bdry}} and [𝒢𝒢]intf\left[\cal GG\right]_{\text{intf}} describe the square of the bulk term, the square of the in-time-boundary term, and the interference between the bulk and in-boundary terms, respectively.

With the approximation (27) and the limit (32), the differential probability (25) takes the simpler form and can be classified into the following three parts:

dPVPP¯=dPVPP¯bulk+dPVPP¯bdry+dPVPP¯intf,\displaystyle\text{d}P_{V\to P\overline{P}}=\text{d}P_{V\to P\overline{P}}^{\text{bulk}}+\text{d}P_{V\to P\overline{P}}^{\text{bdry}}+\text{d}P_{V\to P\overline{P}}^{\text{intf}}, (37)

with

dPVPP¯“type”\displaystyle\text{d}P_{V\to P\overline{P}}^{\text{``type"}} :=|geff|2NV212E0d3𝑷1(2π)32E1d3𝑷2(2π)32E2(2π)4(σtπeσt(δω)2)((σsπ)3/2eσs(δ𝑷)2)\displaystyle:=|g_{\text{eff}}|^{2}N_{V}^{2}\frac{1}{2E_{0}}\frac{\text{d}^{3}\bm{P}_{1}}{(2\pi)^{3}2E_{1}}\frac{\text{d}^{3}\bm{P}_{2}}{(2\pi)^{3}2E_{2}}(2\pi)^{4}\left(\sqrt{\frac{\sigma_{t}}{\pi}}e^{-\sigma_{t}\left(\delta\omega\right)^{2}}\right)\left(\left(\frac{\sigma_{s}}{\pi}\right)^{3/2}e^{-\sigma_{s}\left(\delta\bm{P}\right)^{2}}\right)
×σtπ5(σsσ0σ1σ2)3d3𝑿1d3𝑿2e[𝒢𝒢]“type”(𝔗)eΓV(𝔗T0)+ΓV2σt4|F~(|𝑽1𝑽2|)|2,\displaystyle\quad\times\sqrt{\frac{\sigma_{t}}{\pi^{5}}\left(\frac{\sigma_{s}}{\sigma_{0}\sigma_{1}\sigma_{2}}\right)^{3}}\,\text{d}^{3}\bm{X}_{1}\text{d}^{3}\bm{X}_{2}\,e^{-\mathcal{R}}\,\left[\cal GG\right]_{\text{``type"}}\!\left(\mathfrak{T}\right)\,e^{-\Gamma_{V}\left(\mathfrak{T}-{T_{0}}\right)+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}}\left|\widetilde{F}\!\bigl{(}\left|\bm{V}_{1}-\bm{V}_{2}\right|\bigr{)}\right|^{2}, (38)

where the argument “type” discriminates the three types of contributions.

2.3 Integrated decay probability

To compare the theoretical predictions in the Gaussian wave-packet formalism with the experimental results in Eq. (2), we integrate the differential decay probability (37) over the whole position-momentum phase space of the final-state pseudoscalar mesons, namely, over 𝑿1\bm{X}_{1}, 𝑿2\bm{X}_{2}, 𝑷1\bm{P}_{1}, and 𝑷2\bm{P}_{2}:

PVPP¯=PVPP¯bulk+PVPP¯bdry+PVPP¯intf.\displaystyle P_{V\to P\overline{P}}=P_{V\to P\overline{P}}^{\text{bulk}}+P_{V\to P\overline{P}}^{\text{bdry}}+P_{V\to P\overline{P}}^{\text{intf}}. (39)

We focus on the situation where these integrals can be performed analytically using the saddle-point approximation; see e.g. Ref. [11]. In the current setup, we can safely take non-relativistic approximations in the kinematics of the system because the mass difference mV2mPm_{V}-2m_{P} is small.

Here, we only list the final form of the three types of contributions to the integrated decay probability: the bulk, boundary, and interference contributions. These calculations’ details are provided in Appendix B. For later convenience, we define a common dimensionless factor 𝒞VPP¯{\cal C}_{V\to P\overline{P}} for all of PVPP¯bulkP_{V\to P\overline{P}}^{\text{bulk}}, PVPP¯bdryP_{V\to P\overline{P}}^{\text{bdry}}, and PVPP¯intfP_{V\to P\overline{P}}^{\text{intf}}:

𝒞VPP¯:=gV2mPNV2eΓV(TinT0)12πmV.\displaystyle{\cal C}_{V\to P\overline{P}}:=\frac{g^{2}_{V}m_{P}N_{V}^{2}e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right)}}{12\pi m_{V}}. (40)

2.3.1 Bulk contribution

Integrating the bulk contribution in Eq. (37), we obtain

PVPP¯bulk\displaystyle P_{V\to P\overline{P}}^{\text{bulk}} 𝒞VPP¯mPΓV((mV2mP)2mP2+ΓV24mP2)3/4\displaystyle\simeq{{\cal C}_{V\to P\overline{P}}\,m_{P}\over\Gamma_{V}}\left(\frac{\left(m_{V}-2m_{P}\right)^{2}}{m_{P}^{2}}+{\Gamma_{V}^{2}\over 4m_{P}^{2}}\right)^{3/4}
×12[1+erf(mPσPVB2)]eFbulk0Abulk3/2|F~(VB)|2,\displaystyle\quad\times\frac{1}{2}\left[1+\text{erf}\left(\frac{m_{P}\sqrt{\sigma_{P}}{V_{-}^{\text{B}}}}{\sqrt{2}}\right)\right]{\frac{e^{-F^{0}_{\text{bulk}}}}{A_{\text{bulk}}^{3/2}}\left|\widetilde{F}\!\left(V_{-}^{\text{B}}\right)\right|^{2}}, (41)

where

VB\displaystyle V_{-}^{\text{B}} :=2[(mV2mP)2mP2+ΓV24mP2]1/4,\displaystyle:=2\left[\frac{\left(m_{V}-2m_{P}\right)^{2}}{m_{P}^{2}}+{\Gamma_{V}^{2}\over 4m_{P}^{2}}\right]^{1/4}, (42)
Fbulk0\displaystyle F_{\text{bulk}}^{0} :=mPσP((mV2mP)+(mV2mP)2+ΓV24),\displaystyle:=m_{P}\sigma_{P}\left(-\left(m_{V}-2m_{P}\right)+\sqrt{\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}}\right), (43)
Abulk\displaystyle A_{\text{bulk}} :=12+(mV2mP)2(mV2mP)2+ΓV24.\displaystyle:=\frac{1}{2}+\frac{\left(m_{V}-2m_{P}\right)}{2\sqrt{\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}}}. (44)

We note that the wave packet size of the decaying particle σV\sigma_{V} drops out of this expression at this order of the saddle-point approximation.

2.3.2 Boundary contribution

Integrating the boundary contribution in Eq. (37), we obtain

PVPP¯bdry\displaystyle P_{V\to P\overline{P}}^{\text{bdry}} 𝒞VPP¯Ibdry2π,\displaystyle\simeq{\cal C}_{V\to P\overline{P}}\frac{I_{\text{bdry}}}{2\pi}, (45)

where IbdryI_{\text{bdry}} is written as an integration of a function of VV_{-}:

Ibdry\displaystyle I_{\text{bdry}} :=0dVf~bdry(V),\displaystyle:=\int_{0}^{\infty}\text{d}V_{-}\widetilde{f}_{\text{bdry}}\!\left(V_{-}\right), (46)

in which

f~bdry(V)\displaystyle\widetilde{f}_{\text{bdry}}\!\left(V_{-}\right) =V4(V24mV2mPmP)2+42mP2ΓV24|F~(V)|2\displaystyle=\frac{V_{-}^{4}}{\left(V_{-}^{2}-4\frac{m_{V}-2m_{P}}{m_{P}}\right)^{2}+\frac{4^{2}}{m_{P}^{2}}{{\Gamma_{V}^{2}\over 4}}}\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}
V4[(V24mV2mPmP)234ΓV2mP2]22σPV2(mP4)2[(V24mV2mPmP)2+4ΓV2mP2]3|F~(V)|2.\displaystyle\quad-\frac{V_{-}^{4}\left[\left(V_{-}^{2}-4\frac{m_{V}-2m_{P}}{m_{P}}\right)^{2}-3\frac{4\Gamma_{V}^{2}}{m_{P}^{2}}\right]}{2{\frac{2\sigma_{P}}{V_{-}^{2}}}\left(\frac{m_{P}}{4}\right)^{2}\left[\left(V_{-}^{2}-4\frac{m_{V}-2m_{P}}{m_{P}}\right)^{2}+\frac{4\Gamma_{V}^{2}}{m_{P}^{2}}\right]^{3}}\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}. (47)

The integral (46) will be evaluated numerically.

2.3.3 Interference contribution

Integrating the interference contribution in Eq. (37), we obtain

PVPP¯intf\displaystyle P^{\text{intf}}_{V\to P\overline{P}} 𝒞VPP¯mPσP22π(VI)212[1+erf(mPσPVI2)]eFintf0Aintf3/2|F~(VI)|2\displaystyle\sim-\,{\cal C}_{V\to P\overline{P}}\frac{m_{P}\sqrt{\sigma_{P}}}{2\sqrt{2}\sqrt{\pi}}\left(V_{-}^{\text{I}}\right)^{2}\frac{1}{2}\left[1+\text{erf}\left(\frac{m_{P}\sqrt{\sigma_{P}}{V_{-}^{\text{I}}}}{2}\right)\right]{\frac{e^{-F^{0}_{\text{intf}}}}{A_{\text{intf}}^{3/2}}}\,\left|\widetilde{F}\!\left(V^{\text{I}}_{-}\right)\right|^{2}
×(ΓV2(δω~)2)cos(2ΓVσt~δω~)2ΓVδω~sin(2ΓVσt~δω~)σt~(ΓV2+(δω~)2)2,\displaystyle\quad\times\frac{\left(\Gamma_{V}^{2}-\left(\widetilde{\delta\omega}\right)^{2}\right)\cos\!\left(2\Gamma_{V}\widetilde{\sigma_{t}}\widetilde{\delta\omega}\right)-2\Gamma_{V}\widetilde{\delta\omega}\sin\!\left(2\Gamma_{V}\widetilde{\sigma_{t}}\widetilde{\delta\omega}\right)}{\widetilde{\sigma_{t}}\left(\Gamma_{V}^{2}+\left(\widetilde{\delta\omega}\right)^{2}\right)^{2}}, (48)

where the definition of new parameters is as follows:

VI\displaystyle V_{-}^{\text{I}} :=2[(mV2mP)2+ΓV22]1/4mP,\displaystyle:=\frac{2\left[\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 2}\right]^{1/4}}{\sqrt{m_{P}}}, (49)
Fintf0\displaystyle{F_{\text{intf}}^{0}} :=mPσP2[(mV2mP)+(mV2mP)2+ΓV22],\displaystyle:=\frac{m_{P}\sigma_{P}}{2}\left[-{\left(m_{V}-2m_{P}\right)}+\sqrt{\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 2}}\right], (50)
Aintf\displaystyle A_{\text{intf}} :=14[3+mV2mP(mV2mP)2+ΓV22],\displaystyle:=\frac{1}{4}\left[3+\frac{m_{V}-2m_{P}}{\sqrt{\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 2}}}\right], (51)
σt~\displaystyle\widetilde{\sigma_{t}} :=2σP(VI)2,\displaystyle:=\frac{2\sigma_{P}}{\left(V_{-}^{\text{I}}\right)^{2}}, (52)
δω~\displaystyle\widetilde{\delta\omega} :=14mP(VI)2(mV2mP).\displaystyle:=\frac{1}{4}m_{P}\left(V_{-}^{\text{I}}\right)^{2}-\left(m_{V}-2m_{P}\right). (53)

The tilde denotes that the values are evaluated at the saddle point for the interference contribution.

3 Magnitudes of three kinds of contributions

We have seen the magnitudes of the three kinds of contributions to the integrated probability from the bulk part PVPP¯bulkP_{V\to P\overline{P}}^{\text{bulk}} in Eq. (41), from the boundary part PVPP¯bdryP_{V\to P\overline{P}}^{\text{bdry}} in Eq. (45), and from the bulk-boundary interference PVPP¯intfP_{V\to P\overline{P}}^{\text{intf}} in Eq. (48). Hereafter, we call the following three ratios the P-factors:

𝒫bulk\displaystyle\mathscr{P}_{\text{bulk}} :=PVPP¯bulk𝒞VPP¯,\displaystyle:={P_{V\to P\overline{P}}^{\text{bulk}}\over{\cal C}_{V\to P\overline{P}}}, 𝒫bdry\displaystyle\mathscr{P}_{\text{bdry}} :=PVPP¯bdry𝒞VPP¯,\displaystyle:={P_{V\to P\overline{P}}^{\text{bdry}}\over{\cal C}_{V\to P\overline{P}}}, 𝒫intf\displaystyle\mathscr{P}_{\text{intf}} :=PVPP¯intf𝒞VPP¯,\displaystyle:={P_{V\to P\overline{P}}^{\text{intf}}\over{\cal C}_{V\to P\overline{P}}}, (54)

where the common 𝒞VPP¯{\cal C}_{V\to P\overline{P}} given in Eq. (40) is factored out.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: The P-factors (54) are shown for ϕK+K\phi\to K^{+}K^{-} and ϕK0K0¯\phi\to K^{0}\overline{K^{0}} (in the top row), ψD+D\psi\to D^{+}D^{-} and ψD0D0¯\psi\to D^{0}\overline{D^{0}} (in the middle row), and ΥB+B\Upsilon\to B^{+}B^{-} and ΥB0B0¯\Upsilon\to B^{0}\overline{B^{0}} (in the bottom row), where we take a typical value 0.0015MeV1=(0.67GeV)10.0015\,\text{MeV}^{-1}=\left(0.67\,\text{GeV}\right)^{-1} for R0R_{0}; see Appendix A. The wave-packet treatment breaks down when the wave-packet size of the decay product σP\sqrt{\sigma_{P}} is smaller than the de-Broglie wavelength of PP, which is depicted by the hatched region.

In Fig. 3, the P-factors (54) are shown for the following cases, with a typical value 0.0015MeV1=(0.67GeV)10.0015\,\text{MeV}^{-1}=\left(0.67\,\text{GeV}\right)^{-1} for R0R_{0} (see Appendix A):

  • ϕK+K\phi\to K^{+}K^{-} and ϕK0K0¯\phi\to K^{0}\overline{K^{0}} (1st row),

  • ψD+D\psi\to D^{+}D^{-} and ψD0D0¯\psi\to D^{0}\overline{D^{0}} (2nd row),

  • ΥB+B\Upsilon\to B^{+}B^{-} and ΥB0B0¯\Upsilon\to B^{0}\overline{B^{0}} (3rd row).

We remind that all of PVPP¯bulkP_{V\to P\overline{P}}^{\text{bulk}}, PVPP¯bdryP_{V\to P\overline{P}}^{\text{bdry}}, and PVPP¯intfP_{V\to P\overline{P}}^{\text{intf}} do not depend on σV\sigma_{V} within the non-relativistic saddle-point approximation; see Appendix B for details.

From Fig. 3, we can read the following properties:

  • If the magnitude of σP\sqrt{\sigma_{P}} is relatively low, the three kinds of the P-factors are of the same order.

  • When σP\sqrt{\sigma_{P}} is a certain magnitude, the bulk contribution is exponentially suppressed, while the boundary contribution takes almost the same value, where the interference part is negligible. For such a σP\sqrt{\sigma_{P}} and other higher choices of it, the boundary part dominates.

  • Physically, the wave-packet treatment of the decay product PP breaks down when the wave-packet size σP\sqrt{\sigma_{P}} is shorter than the de-Broglie wavelength of PP:

    λde-Broglie:=2πmPVB,\displaystyle\lambda_{\text{de-Broglie}}:=\frac{2\pi}{m_{P}V_{-}^{\text{B}}}, (55)

    where VBV_{-}^{\text{B}} is the expectation value in the bulk part in Eq. (42). It is straightforward to estimate it for each decay process:

    λde-Broglie|ϕK+K\displaystyle\lambda_{\text{de-Broglie}}|_{\phi\to K^{+}K^{-}} =0.025MeV1=140MeV,\displaystyle=0.025\,\text{MeV}^{-1}={1\over 40\,\text{MeV}}, λde-Broglie|ϕK0K0¯\displaystyle\lambda_{\text{de-Broglie}}|_{\phi\to K^{0}\overline{K^{0}}} =0.029MeV1=134MeV,\displaystyle=0.029\,\text{MeV}^{-1}={1\over 34\,\text{MeV}},
    λde-Broglie|ψD+D\displaystyle\lambda_{\text{de-Broglie}}|_{\psi\to D^{+}D^{-}} =0.012MeV1=183MeV,\displaystyle=0.012\,\text{MeV}^{-1}={1\over 83\,\text{MeV}}, λde-Broglie|ψD0D0¯\displaystyle\lambda_{\text{de-Broglie}}|_{\psi\to D^{0}\overline{D^{0}}} =0.011MeV1=191MeV,\displaystyle=0.011\,\text{MeV}^{-1}={1\over 91\,\text{MeV}},
    λde-Broglie|ΥB+B\displaystyle\lambda_{\text{de-Broglie}}|_{\Upsilon\to B^{+}B^{-}} =0.0090MeV1=10.11GeV,\displaystyle=0.0090\,\text{MeV}^{-1}={1\over 0.11\,\text{GeV}}, λde-Broglie|ΥB0B0¯\displaystyle\lambda_{\text{de-Broglie}}|_{\Upsilon\to B^{0}\overline{B^{0}}} =0.0091MeV1=10.11GeV.\displaystyle=0.0091\,\text{MeV}^{-1}={1\over 0.11\,\text{GeV}}. (56)

    The theoretically excluded region σP<λde-Broglie\sqrt{\sigma_{P}}<\lambda_{\text{de-Broglie}} is depicted by the hatched region.

  • The differences in the P-factors between the charged-meson and neutral-meson final states are sizable for higher-σP\sqrt{\sigma_{P}} regions.

Refer to caption
Figure 4: The distributions of the P-factors (54) representing ψD+D\psi\to D^{+}D^{-} are shown as functions of (σD)1/2\left(\sigma_{D}\right)^{1/2} and R0R_{0}, where the bulk, boundary, and interference ones are shown by the orange, magenta, and green color, respectively.
Refer to caption
Refer to caption
Figure 5: We show the P-factors (54) for ρ0π+π\rho^{0}\to\pi^{+}\pi^{-} (left panel) and the difference between the P-factors for ρ0π+π\rho^{0}\to\pi^{+}\pi^{-} and ρ+π+π0\rho^{+}\to\pi^{+}\pi^{0} (right panel), with a typical value 0.0015MeV1=(0.67GeV)10.0015\,\text{MeV}^{-1}=\left(0.67\,\text{GeV}\right)^{-1} for R0R_{0}; see Appendix A. We set the lowest σπ\sqrt{\sigma_{\pi}} near the bound from the de-Broglie wavelength 0.012MeV1\simeq 0.012\,\text{MeV}^{-1}.

To clarify the dependence on R0R_{0}, we prepare the surface plots for the bulk, boundary, and interference parts of the P-factors for ΨD+D\Psi\to D^{+}D^{-} as a typical example as Fig. 4. Here, the following properties are observed: (i) The magnitude of each part becomes larger for a smaller σD\sqrt{\sigma_{D}} and a smaller R0R_{0}; (ii) In the entire domain of σD\sqrt{\sigma_{D}} and R0R_{0}, the boundary part exceeds the bulk part in magnitude.

In Fig. 5, we have also plotted the P-factors for ρ0π+π\rho^{0}\to\pi^{+}\pi^{-} and ρ+π+π0\rho^{+}\to\pi^{+}\pi^{0} by adopting the same formulas (41), (45) and (48) for a purpose of qualitative comparison between the decays with narrow phase spaces (ϕ\phi, ψ\psi, and Υ\Upsilon) and that with broad phase spaces (ρ\rho), knowing that it is speculative whether we can still use the non-relativistic approximation.202020 We set mρ+mρ0770MeVm_{\rho^{+}}\simeq m_{\rho^{0}}\simeq 770\,\text{MeV} and remind mπ0135MeVm_{\pi^{0}}\simeq 135\,\text{MeV}, and mπ+140MeVm_{\pi^{+}}\simeq 140\,\text{MeV}. We also remind the property that the decay channel ρ0π0π0\rho^{0}\to\pi^{0}\pi^{0} is prohibited by the conservation of the isospin. As expected, the difference between ρ0π+π\rho^{0}\to\pi^{+}\pi^{-} and ρ+π+π0\rho^{+}\to\pi^{+}\pi^{0} is small since the magnitude of the isospin breaking is much smaller even for a smaller σπ\sqrt{\sigma_{\pi}}.

4 Analysis of ratio of decay probabilities RVR_{V}

In this section, we discuss the ratio RVR_{V} of decay probabilities for three vector mesons ϕ\phi, ψ\psi, and Υ\Upsilon in the wave-packet formalism. We compare each with the PDG result and find an agreement around a reasonable value of R0R_{0} of the form factor (for the compositeness of VV). In particular, the 9.5 σ\sigma discrepancy for ψ\psi is dramatically ameliorated. We find that the effect of the form factor is significant in both the wave-packet and plane-wave formalisms.

4.1 Wavepacket analysis

We estimate the wave-packet counterparts of the ratio of the decay rates defined in Eq. (1) for the three vector mesons:

RϕWP\displaystyle R_{\phi}^{\text{WP}} :=PϕK+KPϕK0K0¯,\displaystyle:={P_{\phi\to K^{+}K^{-}}\over P_{\phi\to K^{0}\overline{K^{0}}}}, RψWP\displaystyle R_{\psi}^{\text{WP}} :=PψD+DPψD0D0¯,\displaystyle:={P_{\psi\to D^{+}D^{-}}\over P_{\psi\to D^{0}\overline{D^{0}}}}, RΥWP\displaystyle R_{\Upsilon}^{\text{WP}} :=PΥB+BPΥB0B0¯,\displaystyle:={P_{\Upsilon\to B^{+}B^{-}}\over P_{\Upsilon\to B^{0}\overline{B^{0}}}}, (57)

where we ignored the tiny CP-violation effect in ϕKL0KS0\phi\to K^{0}_{\text{L}}K^{0}_{\text{S}}. We note that the ratio does not depend on the wave-function (field) renormalization factor NVN_{V} (accounting for the offshellness of the vector meson VV), nor on the decay factor eΓV(TinT0)e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right)}. Also, we take the isospin-symmetric limit in the couplings as introduced in Eq. (8), and the dependence on the coupling is dropped off. For further comparison, we also define a “bulk” ratio:

RVbulk:=\displaystyle R_{V}^{\text{bulk}}:= PVP+PbulkPVP0P0¯bulk,\displaystyle\ \frac{P^{\text{bulk}}_{V\to P^{+}P^{-}}}{P^{\text{bulk}}_{V\to P^{0}\overline{P^{0}}}}, (58)

which contains the wave-packet contribution only from the bulk part. Also, we introduce the ratio without interference:

RVwithout interference:=\displaystyle R_{V}^{\text{without interference}}:= PVP+Pbulk+PVP+PbdryPVP0P0¯bulk+PVP0P0¯bdry.\displaystyle\ \frac{P^{\text{bulk}}_{V\to P^{+}P^{-}}+P^{\text{bdry}}_{V\to P^{+}P^{-}}}{P^{\text{bulk}}_{V\to P^{0}\overline{P^{0}}}+P^{\text{bdry}}_{V\to P^{0}\overline{P^{0}}}}. (59)
Refer to caption
Refer to caption
Figure 6: The ratio comparing the decay rates of ϕK+K\phi\to K^{+}K^{-} to ϕK0K0¯\phi\to K^{0}\overline{K^{0}} is drawn as a function of R0R_{0} for two fixed wave-packet sizes of the Kaons of σK=1MeV1\sqrt{\sigma_{K}}=1\,\text{MeV}^{-1} (left panel) and σK=0.1MeV1\sqrt{\sigma_{K}}=0.1\,\text{MeV}^{-1} (right panel). The experimental result is provided by the PDG [1] [shown in Eq. (2)].
Refer to caption
Refer to caption
Figure 7: The ratio comparing the decay rates of ψD+D\psi\to D^{+}D^{-} to ψD0D0¯\psi\to D^{0}\overline{D^{0}} is drawn as a function of R0R_{0} for two fixed wave-packet sizes of the D-mesons of σD=1MeV1\sqrt{\sigma_{D}}=1\,\text{MeV}^{-1} (left panel) and σD=0.01MeV1\sqrt{\sigma_{D}}=0.01\,\text{MeV}^{-1} (right panel). The experimental result is provided by the PDG [1] [shown in Eq. (2)].
Refer to caption
Refer to caption
Figure 8: The ratio comparing the decay rates of ΥB+B\Upsilon\to B^{+}B^{-} to ΥB0B0¯\Upsilon\to B^{0}\overline{B^{0}} is drawn as a function of R0R_{0} for two fixed wave-packet sizes of the B-mesons of σB=1MeV1\sqrt{\sigma_{B}}=1\,\text{MeV}^{-1} (left panel) and σB=0.01MeV1\sqrt{\sigma_{B}}=0.01\,\text{MeV}^{-1} (right panel). The experimental result is provided by the PDG [1] [shown in Eq. (2)].

4.2 Wavepacket results

In Figs. 6, 7, and 8, we show the wave-packet results for the decay probability ratios of ϕ\phi, ψ\psi, and Υ\Upsilon, respectively. Several comments are in order:

  • For the ϕ\phi decay in Fig. 6, the full wave-packet result in the red solid line can fit the PDG result around the form-factor size R02×103MeV1R_{0}\simeq 2\times 10^{-3}\,\text{MeV}^{-1} and 6×103MeV16\times 10^{-3}\,\text{MeV}^{-1} for the wave-packet size of the decay product σK=1MeV1{\sqrt{\sigma_{K}}}=1\,\text{MeV}^{-1} and 0.1MeV10.1\,\text{MeV}^{-1}, respectively.

  • For the ψ\psi decay in Fig. 7, the full wave-packet result in the red solid line can fit the PDG result around the form-factor size R02×103MeV1R_{0}\simeq 2\times 10^{-3}\,\text{MeV}^{-1} and 3×103MeV13\times 10^{-3}\,\text{MeV}^{-1} for the wave-packet size of the decay product σD=1MeV1{\sqrt{\sigma_{D}}}=1\,\text{MeV}^{-1} and 0.01MeV10.01\,\text{MeV}^{-1}, respectively.

  • For the Υ\Upsilon decay in Fig. 8, the full wave-packet result in the red solid line can fit the PDG result around the form-factor size R02×103MeV1R_{0}\simeq 2\times 10^{-3}\,\text{MeV}^{-1} and 2×103MeV12\times 10^{-3}\,\text{MeV}^{-1} for the wave-packet size of the decay product σB=1MeV1{\sqrt{\sigma_{B}}}=1\,\text{MeV}^{-1} and 0.01MeV10.01\,\text{MeV}^{-1}, respectively.

  • As discussed in Fig. 3, if σP\sqrt{\sigma_{P}} is sufficiently large, the bulk contribution becomes exponentially suppressed compared to the boundary one. In this regime, we may still formally evaluate the ratio between the (exponentially small) bulk contributions of P0P^{0} and PP^{-}:

    RVbulkeFbulk0|for P+eFbulk0|for P0{e12MeV2σKfor V=ϕ,e1.0×103MeV2σDfor V=ψ,e3.5×102MeV2σBfor V=Υ,\displaystyle R_{V}^{\text{bulk}}\sim\frac{e^{-F^{0}_{\text{bulk}}|_{\text{for }P^{+}}}}{e^{-F^{0}_{\text{bulk}}|_{\text{for }P^{0}}}}\sim\begin{cases}e^{12\,\text{MeV}^{2}\,\sigma_{K}}&\text{for }V=\phi,\\ e^{-1.0\times 10^{3}\,\text{MeV}^{2}\,\sigma_{D}}&\text{for }V=\psi,\\ e^{3.5\times 10^{2}\,\text{MeV}^{2}\,\sigma_{B}}&\text{for }V=\Upsilon,\end{cases} (60)

    where the exponent is from Eq. (43). This ratio becomes either exponentially large or small due to the mass difference between P+P^{+} and P0P^{0}, where the magnitude of the exponents is much greater than 𝒪(1){\cal O}\!\left(1\right). For example, we obtain σP10MeV2\sigma_{P}\gtrsim 10\,\text{MeV}^{-2} and 100MeV2100\,\text{MeV}^{-2} if we estimate σP\sqrt{\sigma_{P}} to be larger than the smallest radius of an electron in atoms that interact with decay products of PP, namely, the Bohr radius divided by a typical atomic number of the detector atoms, say, aB/Z3MeV1a_{\text{B}}/Z\simeq 3\,\text{MeV}^{-1} and 10MeV110\,\text{MeV}^{-1} for lead and iron with Z=82Z=82 and 26, respectively. As introduced, the experimental results of RϕR_{\phi}, RψR_{\psi}, and RΥR_{\Upsilon} are around unity, and they disagree with RVbulkR_{V}^{\text{bulk}}, both for ϕ\phi and ψ\psi. So, the RVbulkR_{V}^{\text{bulk}} curves are completely out of the depicted ranges of the left panels of Figs. 67, and 8.

4.3 Planewave analysis

For a comparison with the wave-packet results, we also show results with the plane-wave decay rate Γplane\Gamma^{\text{plane}} (see Eq. (168) in Appendix C), taking into account the relativistic form factor (103). The resultant plane-wave ratio becomes

RVplane\displaystyle R_{V}^{\text{plane}} :=ΓVP+PplaneΓVP0P0¯plane=RVparton|(R0(mV24mP02)1/22)2+1(R0(mV24mP+2)1/22)2+1|2,\displaystyle:=\frac{\Gamma^{\text{plane}}_{V\to P^{+}P^{-}}}{\Gamma^{\text{plane}}_{V\to P^{0}\overline{P^{0}}}}=R_{V}^{\text{parton}}\left|\frac{\left(\frac{R_{0}\left(m_{V}^{2}-4m_{P^{0}}^{2}\right)^{1/2}}{2}\right)^{2}+1}{\left(\frac{R_{0}\left(m_{V}^{2}-4m_{P^{+}}^{2}\right)^{1/2}}{2}\right)^{2}+1}\right|^{2}, (61)

where the parton-level contribution to the ratio is

RVparton:=(mV24mP+2mV24mP02)3/2,\displaystyle R_{V}^{\text{parton}}:=\left(\frac{m_{V}^{2}-4m_{P^{+}}^{2}}{m_{V}^{2}-4m_{P^{0}}^{2}}\right)^{3/2}, (62)

and the other factor is from the relativistic form factor (103) written in terms of the masses and R0R_{0}.

For another comparison, we will also show analyses using its non-relativistic approximated form:

RVplaneRVplane, non-rel:=mP+1/2(mV2mP+)3/2mP01/2(mV2mP0)3/2|(R0mP0|𝑽1𝑽2|P02)2+1(R0mP+|𝑽1𝑽2|P+2)2+1|2,\displaystyle R_{V}^{\text{plane}}\Rightarrow R_{V}^{\text{plane, non-rel}}:=\frac{m_{P^{+}}^{1/2}\left(m_{V}-2m_{P^{+}}\right)^{3/2}}{m_{P^{0}}^{1/2}\left(m_{V}-2m_{P^{0}}\right)^{3/2}}\left|\frac{\left(\frac{R_{0}m_{P^{0}}\left|\bm{V}_{1}-\bm{V}_{2}\right|_{P^{0}}}{2}\right)^{2}+1}{\left(\frac{R_{0}m_{P^{+}}\left|\bm{V}_{1}-\bm{V}_{2}\right|_{P^{+}}}{2}\right)^{2}+1}\right|^{2}, (63)

with

|𝑽1𝑽2|P+\displaystyle\left|\bm{V}_{1}-\bm{V}_{2}\right|_{P^{+}} 2(mV2mP+)1/2mP+1/2,\displaystyle\approx\frac{2\left(m_{V}-2m_{P^{+}}\right)^{1/2}}{m_{P^{+}}^{1/2}}, |𝑽1𝑽2|P0\displaystyle\left|\bm{V}_{1}-\bm{V}_{2}\right|_{P^{0}} 2(mV2mP0)1/2mP01/2,\displaystyle\approx\frac{2\left(m_{V}-2m_{P^{0}}\right)^{1/2}}{m_{P^{0}}^{1/2}}, (64)

where “\Rightarrow” represents the operation of taking the non-relativistic approximation and \approx denotes equality under the non-relativistic approximation. The contributions from the form factor are not canceled out in RVplaneR_{V}^{\text{plane}}.212121 A similar factor is taken into account in Ref. [3] as a purely phenomenological cutoff factor of a divergent integral within the plane-wave formalism. Note that the ratio (63) can be obtained from the wave-packet counterpart by taking the limits ΓV0\Gamma_{V}\to 0 and σP\sigma_{P}\to\infty in ΓVPVPP¯\Gamma_{V}P_{V\to P\overline{P}}; see Appendix D for details.

Refer to caption
Refer to caption
Figure 9: The plane-wave ratio comparing the decay rates of ϕK+K\phi\to K^{+}K^{-} to ϕK0K0¯\phi\to K^{0}\overline{K^{0}} is drawn as a function of R0R_{0}, where the captions “rel” and “Non-rel” mean the relativistic and non-relativistic results shown in Eqs. (61) and (63), respectively. For comparison, we also show the wave-packet results for two fixed wave-packet sizes of the Kaons of σK=1MeV1\sqrt{\sigma_{K}}=1\,\text{MeV}^{-1} (left panel) and σK=0.1MeV1\sqrt{\sigma_{K}}=0.1\,\text{MeV}^{-1} (right panel). In both panels, the plane-wave results are the same since they are independent of σK\sigma_{K}. The experimental result, shown in Eq. (2), is provided by the PDG [1].
Refer to caption
Refer to caption
Figure 10: The plane-wave ratio comparing the decay rates of ψD+D\psi\to D^{+}D^{-} to ψD0D0¯\psi\to D^{0}\overline{D^{0}} is drawn as a function of R0R_{0}. For comparison, we also show the wave-packet results for two fixed wave-packet sizes of the D-mesons of σD=1MeV1\sqrt{\sigma_{D}}=1\,\text{MeV}^{-1} (left panel) and σD=0.01MeV1\sqrt{\sigma_{D}}=0.01\,\text{MeV}^{-1} (right panel). The other conventions are the same as in Fig. 9.
Refer to caption
Refer to caption
Figure 11: The plane-wave ratio comparing the decay rates of ΥB+B\Upsilon\to B^{+}B^{-} to ΥB0B0¯\Upsilon\to B^{0}\overline{B^{0}} is drawn as a function of R0R_{0}. For comparison, we also show the wave-packet results for two fixed wave-packet sizes of the B-mesons of σB=1MeV1\sqrt{\sigma_{B}}=1\,\text{MeV}^{-1} (left panel) and σB=0.01MeV1\sqrt{\sigma_{B}}=0.01\,\text{MeV}^{-1} (right panel). The other conventions are the same as in Fig. 9.

4.4 Plane-wave results

We provide comments on the plane-wave results shown in Figs. 9, 10, and 11 below:

  • For all of the vector mesons, ϕ\phi, ψ\psi, and Υ\Upsilon, the parton-level ratios (4) (under the isospin-symmetric limit for the couplings,222222 Here, we are assuming the isospin-symmetric limit in the sense of Eq. (8) for both the wave-packet and plane-wave calculations. without taking into account the form factor) are disfavored with the PDG’s central values at the level of 2.1σ2.1\,\sigma, 9.5σ9.5\,\sigma, and 0.32σ0.32\,\sigma, respectively.

  • On the other hand, when the form-factor effect is included, which is compulsory since the vector mesons are composite particles, we can see agreements with the PDG’s results. It suggests the importance of the form factor in addressing the ratio, where its effect is not fully canceled. Also, we can confirm that the non-relativistic results approximate their relativistic counterparts well for the current system.

  • We can find appropriate ranges of the form-factor parameter R0R_{0}, where theoretical predictions agree with the PDG’s results for both the full wave-packet and plane-wave curves. For each vector meson, the favored regions of R0R_{0} for the wave packet and the plane wave are close to each other but different.

The calculation based on the plane wave works successfully, even though the presumption of free plane waves characterizing initial and final states is, at most, a viable approximation. The wave packet-based calculation provides a comprehensive approach, accounting for all aspects of the quantum nature inherent in the initial and final states, thereby enhancing its reliability. It would be important to precisely discuss the theoretically valid region of R0R_{0}, which depends on many details on the strong interaction. We leave this point for future research.

Note that we also briefly consider the isospin violation on the ρ\rho system, where the result is separately available in Appendix E since it might be out of our main interest.

5 Constraint from the shape of vector-meson resonances

In general, it is expected that the resonance shape of VV is modified by the inclusion of the wave packet effects. That is, vector mesons, produced as resonances in electron-positron colliders, are subjected to a shape-fitting process. This section addresses the constraint on the wave-packet size of pseudoscalar mesons through the resonance shape of the process ee+VPP¯e^{-}e^{+}\to V\to P\overline{P}. Sufficiently precise resonance data from experiments is available for ϕ\phi and ψ\psi, facilitating this purpose. However, detailed resonance data for Υ\Upsilon is currently unavailable. Consequently, our focus is maintained on the instances of ϕ\phi and ψ\psi. Our analysis in this section is meant to be a brief consistency check, assuming the factorization of the production and decay processes of VV in both the wave-packet and plane-wave formalisms, and hence is confined to data around the peak.

5.1 Invariant mass distribution of decaying vector-meson wave packet

First, we summarize the invariant mass distribution of the decaying vector-meson when the Gaussian wave packet describes the decaying state.

We define a Lorentz-invariant mass squared M2M^{2} for the pair of pseudo-scalars in the final state:

M2\displaystyle M^{2} :=(E1+E2)2(𝑷1+𝑷2)2mP2(4+V2),\displaystyle:=\left(E_{1}+E_{2}\right)^{2}-\left({\bm{P}_{1}+\bm{P}_{2}}\right)^{2}\Rightarrow m_{P}^{2}\left(4+V_{-}^{2}\right), (65)

where VV_{-} is the magnitude of 𝑽:=𝑽1𝑽2\bm{V}_{-}:=\bm{V}_{1}-\bm{V}_{2} with 𝑽a=𝑷a/Ea\bm{V}_{a}=\bm{P}_{a}/E_{a} for a=1,2a=1,2 (see Eq. (108) in Appendix B). We will use

V21mP2(M24mP2),\displaystyle V_{-}^{2}\approx\frac{1}{m_{P}^{2}}\left(M^{2}-4m_{P}^{2}\right), (66)

which results in

VdV\displaystyle V_{-}\text{d}V_{-} MdMmP2mVdMmP2,\displaystyle\approx{M\text{d}M\over m_{P}^{2}}\simeq{m_{V}\text{d}M\over m_{P}^{2}}, (67)

where we have approximated that VV decays at rest in the last step.

It is straightforward to derive the following forms after integrating Eq. (38) over the final state phase space, except for VV_{-}, under the current non-relativistic approximation, which is easily rewritten as the invariant mass distribution by use of Eq. (67):

dPVPP¯dM=dPVPP¯bulkdM+dPVPP¯bdrydM+dPVPP¯intfdM,\displaystyle\frac{\text{d}P_{V\to P\overline{P}}}{\text{d}M}=\frac{\text{d}P_{V\to P\overline{P}}^{\text{bulk}}}{\text{d}M}+\frac{\text{d}P_{V\to P\overline{P}}^{\text{bdry}}}{\text{d}M}+\frac{\text{d}P_{V\to P\overline{P}}^{\text{intf}}}{\text{d}M}, (68)

where

dPVPP¯bulkdM\displaystyle\frac{\text{d}P_{V\to P\overline{P}}^{\text{bulk}}}{\text{d}M} 2mVmP2(1162π𝒞VPP¯)mP2σPΓVV2eFbulk0mP2σP2(VVB)2Abulk3/2|F~(V)|2,\displaystyle\simeq\frac{2m_{V}}{m_{P}^{2}}\left(\frac{1}{16\sqrt{2\pi}}{\cal C}_{V\to P\overline{P}}\right){m_{P}^{2}\sqrt{\sigma_{P}}\over\Gamma_{V}}V_{-}^{2}{\frac{e^{-F^{0}_{\text{bulk}}-\frac{m_{P}^{2}\sigma_{P}}{2}\left(V_{-}-V_{-}^{\text{B}}\right)^{2}}}{A_{\text{bulk}}^{3/2}}}\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}, (69)
dPVPP¯bdrydM\displaystyle\frac{\text{d}P_{V\to P\overline{P}}^{\text{bdry}}}{\text{d}M} 2mVmP2(14π𝒞VPP¯)f~bdry(V)V|F~(V)|2,\displaystyle\simeq\frac{2m_{V}}{m_{P}^{2}}\left(\frac{1}{4\pi}{\cal C}_{V\to P\overline{P}}\right)\frac{\widetilde{f}_{\text{bdry}}\!\left(V_{-}\right)}{V_{-}}\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}, (70)
dPVPP¯intfdM\displaystyle\frac{\text{d}P_{V\to P\overline{P}}^{\text{intf}}}{\text{d}M} 2mVmP2(182π𝒞VPP¯)(σPmP2)VeFintf012(mP2σP2)(VVI)2Aintf3/2\displaystyle\simeq-\frac{2m_{V}}{m_{P}^{2}}\left(\frac{1}{8\sqrt{2}\pi}{\cal C}_{V\to P\overline{P}}\right)\left(\sigma_{P}m_{P}^{2}\right)V_{-}{\frac{e^{-F^{0}_{\text{intf}}-\frac{1}{2}\left(\frac{m_{P}^{2}\sigma_{P}}{2}\right)\left(V_{-}-V_{-}^{\text{I}}\right)^{2}}}{A_{\text{intf}}^{3/2}}}
×[(ΓV2(δω)2)cos(2ΓVσtδω)2ΓVδωsin(2ΓVσtδω)(ΓV2+(δω)2)2σt]V+0|F~(V)|2.\displaystyle\quad\times\left[\frac{\left(\Gamma_{V}^{2}-\left(\delta\omega\right)^{2}\right)\cos\!\left(2\Gamma_{V}\sigma_{t}\delta\omega\right)-2\Gamma_{V}\delta\omega\sin\!\left(2\Gamma_{V}\sigma_{t}\delta\omega\right)}{\left(\Gamma_{V}^{2}+\left({\delta\omega}\right)^{2}\right)^{2}\sigma_{t}}\right]_{V_{+}\to 0}\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}. (71)

Here, we consider the distribution of MM instead of M2M^{2} due to the convenience of comparing the wave-packet shape with the non-relativistic Breit-Wigner (BW) shape, which is the well-known resonant shape for the decaying plane wave with the decay rate ΓV\Gamma_{V}; see the next subsection.232323 Experimentally, one may perform a precision experiment by measuring the ratio per each bin ΔM\Delta M near the resonance, in principle: dPVP+PdMΔMdPVP0P0¯dMΔM\displaystyle{{\text{d}P_{V\to P^{+}P^{-}}\over\text{d}M}\Delta M\over{\text{d}P_{V\to P^{0}\overline{P^{0}}}\over\text{d}M}\Delta M} =dPVP+PdMdPVP0P0¯dM.\displaystyle={{\text{d}P_{V\to P^{+}P^{-}}\over\text{d}M}\over{\text{d}P_{V\to P^{0}\overline{P^{0}}}\over\text{d}M}}.

We note that, under the current setup ToutT_{\text{out}}\to\infty, the factor NV2N_{V}^{2} in 𝒞VPP¯{\cal C}_{V\to P\overline{P}} (recall Eq. (40)) can be determined by the normalization

PVPP¯bulk+PVPP¯bdry+PVPP¯intf=(the corresponding branching ratio)\displaystyle P^{\text{bulk}}_{V\to P\overline{P}}+P^{\text{bdry}}_{V\to P\overline{P}}+P^{\text{intf}}_{V\to P\overline{P}}=\left(\text{the corresponding branching ratio}\right) (72)

that is obtained after integrating over MM; see Eqs. (41), (45), and (48).

5.2 Breit-Wigner shape

For the plane-wave calculation, it is well-known that the non-relativistic Breit-Wigner distribution nicely describes the shape of a narrow resonance,242424 The relativistic Breit-Wigner distribution takes fR-BW(s)\displaystyle f_{\text{R-BW}}\!\left(s\right) =(mresΓ)/π(smres2)2+(mresΓ)2\displaystyle=\frac{\left(m_{\text{res}}\,\Gamma\right)/\pi}{\left(s-m_{\text{res}}^{2}\right)^{2}+\left(m_{\text{res}}\,\Gamma\right)^{2}} (+dsfR-BW(s)\displaystyle\bigg{(}\int_{-\infty}^{+\infty}\text{d}s\,f_{\text{R-BW}}\!\left(s\right) =1),\displaystyle=1\bigg{)}, which is not used for the calculation. Mandelstam’s variable ss is equal to E2E^{2}.

fNR-BW(E)\displaystyle f_{\text{NR-BW}}\!\left(E\right) =Γ/(2π)(Emres)2+Γ2/4\displaystyle=\frac{\Gamma/\left(2\pi\right)}{\left(E-m_{\text{res}}\right)^{2}+{\Gamma}^{2}/4} (+dEfNR-BW(E)\displaystyle\bigg{(}\int_{-\infty}^{+\infty}\text{d}E\,f_{\text{NR-BW}}\!\left(E\right) =1),\displaystyle=1\bigg{)}, (73)

where mresm_{\text{res}}, EE, and Γ\Gamma are the resonant mass, the total energy in the center-of-the-mass frame, and the total width of an intermediate resonant particle, respectively. Note that

E=M.\displaystyle E=M. (74)

Since we used the non-relativistic approximation, we are adopting the non-relativistic Breit-Wigner shape (73) for comparison.

5.3 Method of analyzing resonant shape

We assume the following factorization for the resonant production, where the cross-section of the resonant production of VV and its subsequent decay into PP and P¯\overline{P}, σee+VPP¯\sigma_{e^{-}e^{+}\to V\to P\overline{P}}, can be factorized in the wave-packet (WP) and plane-wave (PW) formalisms, respectively as

σee+VPP¯WP(M)\displaystyle\sigma_{e^{-}e^{+}\to V\to P\overline{P}}^{\text{WP}}\!\left(M\right) =𝒩ee+VWPdPVPP¯dM,\displaystyle={\cal N}^{\text{WP}}_{e^{-}e^{+}\to V}\frac{\text{d}P_{V\to P\overline{P}}}{\text{d}M}, (75)
σee+VPP¯PW-Parton(M)\displaystyle\sigma_{e^{-}e^{+}\to V\to P\overline{P}}^{\text{PW-Parton}}\!\left(M\right) =𝒩ee+VPW-PartonfNR-BW(M),\displaystyle={\cal N}^{\text{PW-Parton}}_{e^{-}e^{+}\to V}f_{\text{NR-BW}}\!\left(M\right), (76)
σee+VPP¯PW-FF(M)\displaystyle\sigma_{e^{-}e^{+}\to V\to P\overline{P}}^{\text{PW-FF}}\!\left(M\right) =𝒩ee+VPW-FFfNR-BW(M)|11+R02(M24mP2)4|2,\displaystyle={\cal N}^{\text{PW-FF}}_{e^{-}e^{+}\to V}f_{\text{NR-BW}}\!\left(M\right)\left|\frac{1}{1+{\frac{R_{0}^{2}\left(M^{2}-4m_{P}^{2}\right)}{4}}}\right|^{2}, (77)

where we consider the two cases for PW with and without the form factor (FF); the two cases are discriminated by the short-hand notations “PW-FF” and “PW-Parton”. For the form-factor part of (77), we used the relation in Eq. (66) to convert VV_{-} to MM.

𝒩ee+VWP{\cal N}^{\text{WP}}_{e^{-}e^{+}\to V}, 𝒩ee+VPW-Parton{\cal N}^{\text{PW-Parton}}_{e^{-}e^{+}\to V}, and 𝒩ee+VPW-FF{\cal N}^{\text{PW-FF}}_{e^{-}e^{+}\to V} possess the mass dimension of minus one and describe the factorized production part ee+Ve^{-}e^{+}\to V via the ee+e^{-}e^{+} collision at the center-of-the-mass energy MM. Here, we take these three factors to be independent of MM since the primal structure of the resonance is in dPVPP¯/dM{\text{d}P_{V\to P\overline{P}}}/{\text{d}M} or fNR-BWf_{\text{NR-BW}}, and we use only the data points near the peak of a resonance.252525 As is widely known, under the narrow-width approximation in the plane-wave calculation at the resonant peak M=mresM=m_{\text{res}}, we can derive the factorized form explicitly: σee+VPP¯PW(M)σee+VPWBr(VPP¯);\displaystyle\sigma_{e^{-}e^{+}\to V\to P\overline{P}}^{\text{PW}}\!\left(M\right)\simeq\sigma^{\text{PW}}_{e^{-}e^{+}\to V}\,\text{Br}\!\left(V\to P\overline{P}\right); refer to, e.g., Chapter 16 of [28]. And at this point, 𝒩ee+VPW{\cal N}^{\text{PW}}_{e^{-}e^{+}\to V} is determined as 𝒩ee+VPWσee+VPWπ2ΓVPP¯.\displaystyle{\cal N}^{\text{PW}}_{e^{-}e^{+}\to V}\simeq\sigma^{\text{PW}}_{e^{-}e^{+}\to V}\,\frac{\pi}{2}\Gamma_{V\to P\overline{P}}. Also, we note that the width-to-mass ratios of the vector mesons take Γϕ/mϕ0.42%\Gamma_{\phi}/m_{\phi}\simeq 0.42\% and Γψ/mψ0.72%\Gamma_{\psi}/m_{\psi}\simeq 0.72\%, where the adaptation of the narrow-width approximation is justified. We will take mresm_{\text{res}} and Γ\Gamma for fNR-BW(M)f_{\text{NR-BW}}\!\left(M\right) in Eq. (73) as mVm_{V} and ΓV\Gamma_{V}, respectively; see also Eq. (74). The actual analysis for ee+VPP¯e^{-}e^{+}\to V\to P\overline{P} will be done in the following manner:

  • We focus on the values of the experimentally-given cross sections only around the resonant peak, namely in [mVΓV/2,mV+ΓV/2][m_{V}-\Gamma_{V}/2,m_{V}+\Gamma_{V}/2] since the factorized forms in Eqs. (75), (76), and (77) may work only around the peak. Here, we will adopt the PDG values for mVm_{V} and ΓV\Gamma_{V} [1].

  • In the analysis, we fix the values of ΓV\Gamma_{V} and mPm_{P} as confirmed by the PDG group [1], while we treat mVm_{V} as an unfixed parameter and will determine it through our statistical fit. The isospin-symmetric coupling gVg_{V} and the wave-function renormalization factor for VV, NVN_{V} are taken as unity since it can be absorbed into the factor 𝒩ee+V{\cal N}_{e^{-}e^{+}\to V}. Furthermore, for simplicity, we focus on Tin=T0T_{\text{in}}=T_{0}, where the exponential decay factor in 𝒞VPP¯{\cal C}_{V\to P\overline{P}} in Eq. (40) does not work.

  • Under the current scheme, σee+VPP¯WP\sigma_{e^{-}e^{+}\to V\to P\overline{P}}^{\text{WP}} has six parameters {𝒩ee+VWP,mV,ΓV,mP,R0,σP}\{{\cal N}^{\text{WP}}_{e^{-}e^{+}\to V},\,m_{V},\,\Gamma_{V},\,m_{P},\,R_{0},\,\sigma_{P}\}, σee+VPP¯PW-FF\sigma_{e^{-}e^{+}\to V\to P\overline{P}}^{\text{PW-FF}} has five parameters {𝒩ee+VPW-FF,mV,ΓV,mP,R0}\{{\cal N}^{\text{PW-FF}}_{e^{-}e^{+}\to V},\,m_{V},\,\Gamma_{V},\,m_{P},\,R_{0}\}, and σee+VPP¯PW-Parton\sigma_{e^{-}e^{+}\to V\to P\overline{P}}^{\text{PW-Parton}} has three parameters {𝒩ee+VPW-Parton,mV,ΓV}\{{\cal N}^{\text{PW-Parton}}_{e^{-}e^{+}\to V},\,m_{V},\,\Gamma_{V}\}, respectively. We will determine them through statistical analysis. We remind ourselves that the vector-meson wavepacket size σV\sigma_{V} does not appear in σee+VPP¯WP\sigma_{e^{-}e^{+}\to V\to P\overline{P}}^{\text{WP}} under the saddle-point approximation.

5.4 Result of ϕ\phi

In Ref. [29], the latest result of the resonant shape of ϕ\phi through ee+ϕK+Ke^{-}e^{+}\to\phi\to K^{+}K^{-} measured with the CMD-3 detector in the center-of-mass energy range 101010101060MeV1060\,\text{MeV} was reported, where the Born cross sections of ee+ϕK+Ke^{-}e^{+}\to\phi\to K^{+}K^{-} around the resonance are available in Table I of [29]. According to our guideline, we adopt the seven data points from 1018.0MeV1018.0\,\text{MeV} to 1021.3MeV1021.3\,\text{MeV} and adopt the χ2\chi^{2} functions:

χϕ,WP2\displaystyle\chi^{2}_{\phi,\,\text{WP}} :=i=17(σiWPσiexp)2(δσiexp)2,\displaystyle:=\sum_{i=1}^{7}\frac{\left(\sigma_{i}^{\text{WP}}-\sigma_{i}^{\text{exp}}\right)^{2}}{\left(\delta\sigma_{i}^{\text{exp}}\right)^{2}},
χϕ,PW-Parton2\displaystyle\chi^{2}_{\phi,\,\text{PW-Parton}} :=i=17(σiPW-Partonσiexp)2(δσiexp)2,\displaystyle:=\sum_{i=1}^{7}\frac{\left(\sigma_{i}^{\text{PW-Parton}}-\sigma_{i}^{\text{exp}}\right)^{2}}{\left(\delta\sigma_{i}^{\text{exp}}\right)^{2}}, χϕ,PW-FF2\displaystyle\chi^{2}_{\phi,\,\text{PW-FF}} :=i=17(σiPW-FFσiexp)2(δσiexp)2,\displaystyle:=\sum_{i=1}^{7}\frac{\left(\sigma_{i}^{\text{PW-FF}}-\sigma_{i}^{\text{exp}}\right)^{2}}{\left(\delta\sigma_{i}^{\text{exp}}\right)^{2}}, (78)

where ii discriminates the seven points of MM where experimental data is available; σiexp\sigma_{i}^{\text{exp}} and δσiexp\delta\sigma_{i}^{\text{exp}} are the central and error of the experimentally-determined cross section at the point ii, respectively. σiWP\sigma_{i}^{\text{WP}}, σiPW-Parton\sigma_{i}^{\text{PW-Parton}} and σiPW-FF\sigma_{i}^{\text{PW-FF}} represent the theoretical values of the corresponding cross sections at the energy point identified by ii.

Refer to caption
Refer to caption
Figure 12: The fitted resonance distributions of ϕ\phi in ee+ϕK+Ke^{-}e^{+}\to\phi\to K^{+}K^{-} are drawn for the two sets of the fixed parameters σK=10MeV1\sqrt{\sigma_{K}}=10\,\text{MeV}^{-1} and R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1} [Left panel] and σK=10MeV1\sqrt{\sigma_{K}}=10\,\text{MeV}^{-1} and R0=0.01MeV1R_{0}=0.01\,\text{MeV}^{-1} [Right panel], where the mass of ϕ\phi and the normalization factor 𝒩ee+ϕ{\cal N}_{e^{-}e^{+}\to\phi} are determined through our statistical analysis based on the χ2\chi^{2} function defined in Eq. (78). The best-fit parameters and the χ2\chi^{2} functions for the left/right panel are shown in Eqs. (79) and (80) / in Eqs. (81) and (82), respectively, for the wave-packet (WP) and plane-wave without/with form factor (BW/BW with FF).

As examples, we show the fitted distributions for the two sets of the fixed parameters σK=10MeV1\sqrt{\sigma_{K}}=10\,\text{MeV}^{-1} and R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1} for the left panel of Fig. 12,262626 Note that the value of R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1} is a typical value in the current scheme of the form factor (see Appendix A), and its magnitude is favored by the analysis on RVR_{V} (see Section 4). σK=10MeV1\sqrt{\sigma_{K}}={10}\,\text{MeV}^{-1} and R0=0.01MeV1R_{0}=0.01\,\text{MeV}^{-1} for the right panel of Fig. 12, where the two remaining parameters {𝒩ee+ϕ,mϕ}\{{\cal N}_{e^{-}e^{+}\to\phi},\,m_{\phi}\} take the best-fit values, and the values of the χ2\chi^{2} over the degrees of freedoms (DOFs), which is currently five, at the best-fit points are calculated as

  • for σK=10MeV1\sqrt{\sigma_{K}}=10\,\text{MeV}^{-1} and R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1} [Left panel of Fig. 12]:

    mϕWP|best-fit\displaystyle\left.m_{\phi}^{\text{WP}}\right|_{\text{best-fit}} =1019.8MeV,\displaystyle=1019.8\,\text{MeV}, 𝒩ee+ϕWP|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\phi}^{\text{WP}}\right|_{\text{best-fit}} =6.23×105MeV1,\displaystyle=6.23\times 10^{5}\,\text{MeV}^{-1},
    mϕPW-Parton|best-fit\displaystyle\left.m_{\phi}^{\text{PW-Parton}}\right|_{\text{best-fit}} =1019.4MeV,\displaystyle=1019.4\,\text{MeV}, 𝒩ee+ϕPW-Parton|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\phi}^{\text{PW-Parton}}\right|_{\text{best-fit}} =1.50×104MeV1,\displaystyle=1.50\times 10^{4}\,\text{MeV}^{-1},
    mϕPW-FF|best-fit\displaystyle\left.m_{\phi}^{\text{PW-FF}}\right|_{\text{best-fit}} =1019.4MeV,\displaystyle=1019.4\,\text{MeV}, 𝒩ee+ϕPW-FF|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\phi}^{\text{PW-FF}}\right|_{\text{best-fit}} =1.62×104MeV1,\displaystyle=1.62\times 10^{4}\,\text{MeV}^{-1}, (79)
    χϕ,WP2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{WP}}}{(\text{DOFs})}\right|_{\text{best-fit}} 5.5,\displaystyle\simeq 5.5, χϕ,PW-Parton2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{PW-Parton}}}{(\text{DOFs})}\right|_{\text{best-fit}} 6.3,\displaystyle\simeq 6.3, χϕ,PW-FF2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{PW-FF}}}{(\text{DOFs})}\right|_{\text{best-fit}} 6.7,\displaystyle\simeq 6.7, (80)
  • for σK=10MeV1\sqrt{\sigma_{K}}=10\,\text{MeV}^{-1} and R0=0.01MeV1R_{0}=0.01\,\text{MeV}^{-1} [Right panel of Fig. 12]:

    mϕWP|best-fit\displaystyle\left.m_{\phi}^{\text{WP}}\right|_{\text{best-fit}} =1019.9MeV,\displaystyle=1019.9\,\text{MeV}, 𝒩ee+ϕWP|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\phi}^{\text{WP}}\right|_{\text{best-fit}} =3.94×106MeV1,\displaystyle=3.94\times 10^{6}\,\text{MeV}^{-1},
    mϕPW-Parton|best-fit\displaystyle\left.m_{\phi}^{\text{PW-Parton}}\right|_{\text{best-fit}} =1019.4MeV,\displaystyle=1019.4\,\text{MeV}, 𝒩ee+ϕPW-Parton|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\phi}^{\text{PW-Parton}}\right|_{\text{best-fit}} =1.51×104MeV1,\displaystyle=1.51\times 10^{4}\,\text{MeV}^{-1},
    mϕPW-FF|best-fit\displaystyle\left.m_{\phi}^{\text{PW-FF}}\right|_{\text{best-fit}} =1019.6MeV,\displaystyle=1019.6\,\text{MeV}, 𝒩ee+ϕPW-FF|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\phi}^{\text{PW-FF}}\right|_{\text{best-fit}} =1.03×105MeV1,\displaystyle=1.03\times 10^{5}\,\text{MeV}^{-1}, (81)
    χϕ,WP2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{WP}}}{(\text{DOFs})}\right|_{\text{best-fit}} 14,\displaystyle\simeq 14, χϕ,PW-Parton2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{PW-Parton}}}{(\text{DOFs})}\right|_{\text{best-fit}} 6.3,\displaystyle\simeq 6.3, χϕ,PW-FF2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{PW-FF}}}{(\text{DOFs})}\right|_{\text{best-fit}} 14.\displaystyle\simeq 14. (82)

We comment on the difference between the plane-wave resonant shapes with and without the form factor. Without the factor, the shape obeys the Breit–Wigner distribution (73) and is symmetric under the reflection around the peak (M=mϕM=m_{\phi}), while taking into account it, the resonant shape becomes asymmetric under the reflection around the peak. The magnitude of the asymmetry is governed by the part R02(M24mK+2)R_{0}^{2}\left(M^{2}-4m_{K^{+}}^{2}\right) of the form factor. So, for a greater R0R_{0}, a more significant asymmetry will be realized, as observed in Fig. 12.

Refer to caption
Figure 13: We plot the variable χϕ2/(DOFs)|min\left.{\chi^{2}_{\phi}}/{(\text{DOFs})}\right|_{\text{min}} defined in Eq. (83) to compare the significance of the wave-packet calculation with the plane-wave one for various σK\sqrt{\sigma_{K}}. Here, R0R_{0} is fixed as 0.0015MeV10.0015\,\text{MeV}^{-1} and for each σK\sqrt{\sigma_{K}}, mϕm_{\phi} and 𝒩ee+ϕ{\cal N}_{e^{-}e^{+}\to\phi} are determined to (locally) minimize the corresponding χ2\chi^{2} function in Eq. (78). The black curve and blue dashed horizontal line describe the values in the wave-packet and plane-wave calculations without form factor, where the latter is manifestly independent of σK\sqrt{\sigma_{K}}.

Here, we comment on the origin of the “over-5σ5\sigma” values of χ2/(DOFs)\chi^{2}/(\text{DOFs}): this is because the resolution of the experimental results near the peak is very high, and the current simple scheme for σee+VPP¯\sigma_{e^{-}e^{+}\to V\to P\overline{P}} in Eqs. (75), (76) and (77) is not enough for discussing statistical significance precisely. On the other hand, however, we are able to discuss the relative significance between the wave-packet and plane-wave results. According to Eq. (80), the shape of the wave-packet resonant distribution is at least as good as that of the plane-wave resonant distribution at the focused parameter point, where we conclude that the wave-packet result at the first parameter point (for the left panel of Fig. 12) is consistent with the experiment. Note that at the first parameter point, R0R_{0} is taken as a typical value in the current scheme of the form factor (see Appendix A), and a wave packet with a greater size looks similar to the plane wave.

We also see the significance of the wave-packet results over a broad range of σK\sqrt{\sigma_{K}} under R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1}. In Fig. 13, we plot the “minimized” χ2/(DOFs)\chi^{2}/(\text{DOFs}) defined by

χϕ,WP/PW-Parton2(DOFs)|min:=minmϕ,𝒩ee+ϕWP/PW-Parton[χϕ,WP/PW-Parton2(DOFs)],\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{WP/PW-Parton}}}{(\text{DOFs})}\right|_{\text{min}}:=\min_{m_{\phi},\ {\cal N}_{e^{-}e^{+}\to\phi}^{\text{WP/PW-Parton}}}\left[\frac{\chi^{2}_{\phi,\,\text{WP/PW-Parton}}}{(\text{DOFs})}\right], (83)

which measures the statistical significance for σK\sqrt{\sigma_{K}}. We do not consider the PW-FF case since no sizable difference is generated when R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1}, as shown in the left panel of Fig. 12, and the form-factor part does not depend on σK\sqrt{\sigma_{K}}. Under the simple guideline that a wave-packet result is at least as good as the ordinary plane-wave one, from Fig. 13, we can put the lower bound on σK\sqrt{\sigma_{K}} as

σK3MeV1\displaystyle\sqrt{\sigma_{K}}\gtrsim 3\,\text{MeV}^{-1} (84)

for R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1}.

5.5 Result of ψ\psi

The BaBar, Belle, BES, and CLEO experimental collaborations have provided recent experimental data of the ψ\psi’s resonance produced by the ee+e^{-}e^{+} collision.

  • We will adopt the results on [30] of the converting experimental results to the exclusive initial-state-radiation scattering cross-section, e+e2De^{+}e^{-}\to 2D, where the following experimental papers are taken into account by BaBar [31], by Belle [32], and by CLEO [33, 34]. “2D2D” means the inclusive final states of D+DD^{+}D^{-} and D0D0¯D^{0}\overline{D^{0}}.272727 The exclusive initial-state-radiation scattering cross-sections of e+eD+De^{+}e^{-}\to D^{+}D^{-} and e+eD0D0¯e^{+}e^{-}\to D^{0}\overline{D^{0}} are also reported by Babar [35] and by Belle [36]. Since few data points are available inside the focused range [mψΓψ/2,mψ+Γψ/2][m_{\psi}-\Gamma_{\psi}/2,\,m_{\psi}+\Gamma_{\psi}/2], we do not adopt them for our analysis.

  • We also take account of the experimental results by BES of the inclusive hadron-production cross-section in [37, 38]. The original data is provided in the form,

    R(s)=σhad0(s)σμ+μ0(s),\displaystyle R(s)=\frac{\sigma^{0}_{\text{had}}(s)}{\sigma^{0}_{\mu^{+}\mu^{-}}(s)}, (85)

    where σμ+μ0(s)=4πα2(0)/3s\sigma^{0}_{\mu^{+}\mu^{-}}(s)=4\pi\alpha^{2}\!\left(0\right)/3s is the lowest-order QED cross section for muon pair production at the total center-of-mass energy E=sE=\sqrt{s}. α(0)1/137\alpha\!\left(0\right)\simeq 1/137 is the QED fine structure constant at the Thomson limit. Ruds(c)+ψ(3770)R_{uds(c)+\psi(3770)} and RudsR_{uds} are reported in [37] and in [38], respectively. Through the approximation Rψ(3770)Ruds(c)+ψ(3770)RudsR_{\psi(3770)}\simeq R_{uds(c)+\psi(3770)}-R_{uds}, we can recast the cross section of e+e2De^{+}e^{-}\to 2D, as done in [39].

Since the final state is inclusive as 2D2D, we adopt the following factorized form for the production cross section:

σee+ψ2DWP(M)\displaystyle\sigma_{e^{-}e^{+}\to\psi\to 2D}^{\text{WP}}\!\left(M\right) =𝒩ee+ψWP(dPψD+DdM+dPψD0D0¯dM),\displaystyle={\cal N}^{\text{WP}}_{e^{-}e^{+}\to\psi}\left(\frac{\text{d}P_{\psi\to D^{+}D^{-}}}{\text{d}M}+\frac{\text{d}P_{\psi\to D^{0}\overline{D^{0}}}}{\text{d}M}\right), (86)
σee+ψ2DPW-Parton(M)\displaystyle\sigma_{e^{-}e^{+}\to\psi\to 2D}^{\text{PW-Parton}}\!\left(M\right) =𝒩ee+ψPW-PartonfNR-BW(M),\displaystyle={\cal N}^{\text{PW-Parton}}_{e^{-}e^{+}\to\psi}\,f_{\text{NR-BW}}\!\left(M\right), (87)
σee+ψ2DPW-FF(M)\displaystyle\sigma_{e^{-}e^{+}\to\psi\to 2D}^{\text{PW-FF}}\!\left(M\right) =𝒩ee+ψPW-FFfNR-BW(M)\displaystyle={\cal N}^{\text{PW-FF}}_{e^{-}e^{+}\to\psi}\,f_{\text{NR-BW}}\!\left(M\right)
×[BrψD+D|11+R02(M24mD+2)4|2+BrψD0D0¯|11+R02(M24mD02)4|2],\displaystyle\quad\times\left[\text{Br}_{\psi\to D^{+}D^{-}}\left|\frac{1}{1+{\frac{R_{0}^{2}\left(M^{2}-4m_{D^{+}}^{2}\right)}{4}}}\right|^{2}+\text{Br}_{\psi\to D^{0}\overline{D^{0}}}\left|\frac{1}{1+{\frac{R_{0}^{2}\left(M^{2}-4m_{D^{0}}^{2}\right)}{4}}}\right|^{2}\right], (88)

where BrψD+D=0.41\text{Br}_{\psi\to D^{+}D^{-}}=0.41 and BrψD0D0¯=0.52\text{Br}_{\psi\to D^{0}\overline{D^{0}}}=0.52 are the corresponding branching ratios [1]. The χ2\chi^{2} functions are defined as

χψ,WP2\displaystyle\chi^{2}_{\psi,\,\text{WP}} :=I: BaBar, Belle, BES, CLEOiI(σiIWPσiIexp,I)2(δσiIexp,I)2,\displaystyle:=\sum_{\text{I: BaBar, Belle, BES, CLEO}}\sum_{i_{\text{I}}}\frac{\left(\sigma_{i_{\text{I}}}^{\text{WP}}-\sigma_{i_{\text{I}}}^{\text{exp,I}}\right)^{2}}{\left(\delta\sigma_{i_{\text{I}}}^{\text{exp,I}}\right)^{2}},
χψ,PW-Parton2\displaystyle\chi^{2}_{\psi,\,\text{PW-Parton}} :=I: BaBar, Belle, BES, CLEOiI(σiIPW-PartonσiIexp,I)2(δσiIexp,I)2,\displaystyle:=\sum_{\text{I: BaBar, Belle, BES, CLEO}}\sum_{i_{\text{I}}}\frac{\left(\sigma_{i_{\text{I}}}^{\text{PW-Parton}}-\sigma_{i_{\text{I}}}^{\text{exp,I}}\right)^{2}}{\left(\delta\sigma_{i_{\text{I}}}^{\text{exp,I}}\right)^{2}},
χψ,PW-FF2\displaystyle\chi^{2}_{\psi,\,\text{PW-FF}} :=I: BaBar, Belle, BES, CLEOiI(σiIPW-FFσiIexp,I)2(δσiIexp,I)2,\displaystyle:=\sum_{\text{I: BaBar, Belle, BES, CLEO}}\sum_{i_{\text{I}}}\frac{\left(\sigma_{i_{\text{I}}}^{\text{PW-FF}}-\sigma_{i_{\text{I}}}^{\text{exp,I}}\right)^{2}}{\left(\delta\sigma_{i_{\text{I}}}^{\text{exp,I}}\right)^{2}}, (89)

where σiIexp,I\sigma_{i_{\text{I}}}^{\text{exp,I}} and δσiIexp,I\delta\sigma_{i_{\text{I}}}^{\text{exp,I}} are the central and error of the experimentally-determined cross section at the point iIi_{\text{I}} of the experiment I, respectively. σiIWP\sigma_{i_{\text{I}}}^{\text{WP}}, σiIPW-Parton\sigma_{i_{\text{I}}}^{\text{PW-Parton}} and σiIPW-FF\sigma_{i_{\text{I}}}^{\text{PW-FF}} represent the theoretical values of the corresponding cross sections at the energy point identified by iI{i_{\text{I}}}.

Refer to caption
Refer to caption
Figure 14: The fitted resonance distributions of ϕ\phi in ee+ψ2De^{-}e^{+}\to\psi\to 2D are drawn for the fixed parameters σD=1MeV1\sqrt{\sigma_{D}}=1\,\text{MeV}^{-1} and R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1} (for the left panel), and σD=1MeV1\sqrt{\sigma_{D}}=1\,\text{MeV}^{-1} and R0=0.01MeV1R_{0}=0.01\,\text{MeV}^{-1} (for the right panel), where the mass of ψ\psi and the normalization factor 𝒩ee+ψ{\cal N}_{e^{-}e^{+}\to\psi} are determined through our statistical analysis based on the χ2\chi^{2} function defined in Eq. (89). The best-fit parameters and the χ2\chi^{2} functions for the left/right panel are shown in Eqs. (90) and (91) / in Eqs. (92) and (93), respectively. The other conventions are the same as those of Fig. 12.

Here, we will see the two examples for the same wave-packet size σD=1MeV1\sqrt{\sigma_{D}}=1\,\text{MeV}^{-1} but two different values for the form-factor parameter R0R_{0}. In the left and right panels of Fig. 14, the fitted distributions about the parameters {𝒩ee+ψ,mψ}\{{\cal N}_{e^{-}e^{+}\to\psi},\,m_{\psi}\} are shown for R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1} and R0=0.01MeV1R_{0}=0.01\,\text{MeV}^{-1}, respectively, where the valid ranges of 𝒩ee+ψ{\cal N}_{e^{-}e^{+}\to\psi} and mψm_{\psi} are fixed as

  • for σD=1MeV1\sqrt{\sigma_{D}}=1\,\text{MeV}^{-1} and R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1} [Left panel of Fig. 14]:

    mψWP|best-fit\displaystyle\left.m_{\psi}^{\text{WP}}\right|_{\text{best-fit}} =3770.4MeV,\displaystyle=3770.4\,\text{MeV}, 𝒩ee+ψWP|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\psi}^{\text{WP}}\right|_{\text{best-fit}} =1.02×105MeV1,\displaystyle=1.02\times 10^{5}\,\text{MeV}^{-1},
    mψPW-Parton|best-fit\displaystyle\left.m_{\psi}^{\text{PW-Parton}}\right|_{\text{best-fit}} =3774.6MeV,\displaystyle=3774.6\,\text{MeV}, 𝒩ee+ψPW-Parton|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\psi}^{\text{PW-Parton}}\right|_{\text{best-fit}} =3.90×102MeV1,\displaystyle=3.90\times 10^{2}\,\text{MeV}^{-1},
    mψPW-FF|best-fit\displaystyle\left.m_{\psi}^{\text{PW-FF}}\right|_{\text{best-fit}} =3775.5MeV,\displaystyle=3775.5\,\text{MeV}, 𝒩ee+ψPW-FF|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\psi}^{\text{PW-FF}}\right|_{\text{best-fit}} =5.77×102MeV1,\displaystyle=5.77\times 10^{2}\,\text{MeV}^{-1}, (90)
    χϕ,WP2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{WP}}}{(\text{DOFs})}\right|_{\text{best-fit}} 0.92,\displaystyle\simeq 0.92, χϕ,PW-Parton2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{PW-Parton}}}{(\text{DOFs})}\right|_{\text{best-fit}} 0.91,\displaystyle\simeq 0.91, χϕ,PW-FF2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{PW-FF}}}{(\text{DOFs})}\right|_{\text{best-fit}} 0.91.\displaystyle\simeq 0.91. (91)
  • for σD=1MeV1\sqrt{\sigma_{D}}=1\,\text{MeV}^{-1} and R0=0.01MeV1R_{0}=0.01\,\text{MeV}^{-1} [Right panel of Fig. 14]:

    mψWP|best-fit\displaystyle\left.m_{\psi}^{\text{WP}}\right|_{\text{best-fit}} =3775.7MeV,\displaystyle=3775.7\,\text{MeV}, 𝒩ee+ψWP|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\psi}^{\text{WP}}\right|_{\text{best-fit}} =4.92×106MeV1,\displaystyle=4.92\times 10^{6}\,\text{MeV}^{-1},
    mψPW-Parton|best-fit\displaystyle\left.m_{\psi}^{\text{PW-Parton}}\right|_{\text{best-fit}} =3774.6MeV,\displaystyle=3774.6\,\text{MeV}, 𝒩ee+ψPW-Parton|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\psi}^{\text{PW-Parton}}\right|_{\text{best-fit}} =3.90×102MeV1,\displaystyle=3.90\times 10^{2}\,\text{MeV}^{-1},
    mψPW-FF|best-fit\displaystyle\left.m_{\psi}^{\text{PW-FF}}\right|_{\text{best-fit}} =3780.0MeV,\displaystyle=3780.0\,\text{MeV}, 𝒩ee+ψPW-FF|best-fit\displaystyle\left.{\cal N}_{e^{-}e^{+}\to\psi}^{\text{PW-FF}}\right|_{\text{best-fit}} =3.36×104MeV1,\displaystyle=3.36\times 10^{4}\,\text{MeV}^{-1}, (92)
    χϕ,WP2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{WP}}}{(\text{DOFs})}\right|_{\text{best-fit}} 0.92,\displaystyle\simeq 0.92, χϕ,PW-Parton2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{PW-Parton}}}{(\text{DOFs})}\right|_{\text{best-fit}} 0.91,\displaystyle\simeq 0.91, χϕ,PW-FF2(DOFs)|best-fit\displaystyle\left.\frac{\chi^{2}_{\phi,\,\text{PW-FF}}}{(\text{DOFs})}\right|_{\text{best-fit}} 0.92.\displaystyle\simeq 0.92. (93)

From Fig. 14, when R0R_{0} is large as 102MeV1\sim 10^{-2}\,\text{MeV}^{-1}, the resonant distribution of the wave packet becomes identical with that of the plane-wave without taking into account the form factor. Meanwhile, when R0103MeV1R_{0}\sim 10^{-3}\,\text{MeV}^{-1}, where this size is favored with the agreement in RψR_{\psi}, we observe the deviation from the BW shape in the wave-packet distribution. Note that all three kinds of distributions agree with the experimental data for the larger and smaller R0R_{0}.

We comment on the large asymmetry observed in the right panel of Fig. 14, namely the large deviation of “BW with FF” in the low MM range. As mentioned in the previous subsection, the asymmetry under the reflection around the peak originates from the parts R02(M24mD+2)R_{0}^{2}\left(M^{2}-4m_{D^{+}}^{2}\right) and R02(M24mD02)R_{0}^{2}\left(M^{2}-4m_{D^{0}}^{2}\right) of the form factors. The realized asymmetry in R0=0.01MeV1R_{0}=0.01\,\text{MeV}^{-1} becomes extensive when MM is less than the range used for the statistical fit, so this case is considered to be disfavored even though the limited part near the resonant peak is fitted to the experimental results well.

Refer to caption
Figure 15: We plot the variable χψ2/(DOFs)|min\left.{\chi^{2}_{\psi}}/{(\text{DOFs})}\right|_{\text{min}} defined in Eq. (94) to compare the significance of the wave-packet calculation with the plane-wave one for various σD\sqrt{\sigma_{D}}. Here, R0R_{0} is fixed as 0.0015MeV10.0015\,\text{MeV}^{-1} and for each σD\sqrt{\sigma_{D}}, mψm_{\psi} and 𝒩ee+ψ{\cal N}_{e^{-}e^{+}\to\psi} are determined to (locally) minimize the corresponding χ2\chi^{2} function in Eq. (89). The black curve and the blue dashed horizontal line describe the values in the wave-packet and plane-wave calculations, respectively, where the latter is manifestly independent of σD\sqrt{\sigma_{D}}.

To clarify the experimentally-valid range for σD\sqrt{\sigma_{D}}, we see the curve of the “minimized” χ2/(DOFs)\chi^{2}/(\text{DOFs}) defined by

χψ,WP/PW-Parton2(DOFs)|min:=minmψ,𝒩ee+ψWP/PW-Parton[χψ,WP/PW-Parton2(DOFs)],\displaystyle\left.\frac{\chi^{2}_{\psi,\,\text{WP/PW-Parton}}}{(\text{DOFs})}\right|_{\text{min}}:=\min_{m_{\psi},\ {\cal N}_{e^{-}e^{+}\to\psi}^{\text{WP/PW-Parton}}}\left[\frac{\chi^{2}_{\psi,\,\text{WP/PW-Parton}}}{(\text{DOFs})}\right], (94)

for R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1}. Due to the same reason with the case of ϕ\phi for R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1}, we skip to consider the PW-FF case. From Fig. 15, we recognize that the range for σD\sqrt{\sigma_{D}} is consistent with the constraint,

χψ,WP/PW-Parton2(DOFs)|min<2,\displaystyle\left.\frac{\chi^{2}_{\psi,\,\text{WP/PW-Parton}}}{(\text{DOFs})}\right|_{\text{min}}<2,

even though the wave-packet shape does not exceed the BW shape in the goodness of fit. To summarize, within the current scheme for the production cross section, no significant bounds on σD\sqrt{\sigma_{D}} are imposed. This is because, as recognized from Fig. 14, the experimental results still have sizable errors for the ψ\psi’s resonant shape.

6 Summary and discussion

In this paper, we have discussed the long-standing anomaly in the ratio of the decay rates of the vector mesons ϕ\phi and ψ\psi, namely, Rϕ=Γ(ϕK+K)/Γ(ϕKL0KS0)R_{\phi}=\Gamma\!\left(\phi\to K^{+}K^{-}\right)/\Gamma\!\left(\phi\to K^{0}_{\text{L}}K^{0}_{\text{S}}\right) and Rψ=Γ(ψD+D)/Γ(ψD0D0¯)R_{\psi}=\Gamma\!\left(\psi\to D^{+}D^{-}\right)/\Gamma\!\bigl{(}\psi\to D^{0}\overline{D^{0}}\bigr{)}, where the strong interaction causes the decay channels, and they measure isospin breakings. If we estimate their theoretical values in the plane-wave formalism without considering the effects originating from the composite nature of the initial-state vector mesons, they are disfavored with the PDG’s central values at the level of 2.1σ2.1\,\sigma and 9.5σ9.5\,\sigma. In particular, there has been no explanation for the latter 9.5 σ\sigma anomaly so far.

The decay channels that we focus on are near the mass thresholds, where the velocities in the final state are small, and hence the localization of the overlap of the wave packets is more significant. Here, we fully take into account such effects in the Gaussian-wave-packet formalism. We carefully clarified the properties of one-to-two-body non-relativistic quantum transitions between normalizable physical states described by Gaussian wave packets under the presence of the decaying nature of the initial state, which is a full-fledged calculation taking into account the essences that are missing in the plane-wave calculations.

The result shows agreement with the PDG’s combined results within 1σ\sim 1\,\sigma confidence level. We conclude that the long-standing anomalies in RϕR_{\phi} and RψR_{\psi} are resolved.

In the calculation, the above-mentioned compositeness has been described by the form factor. The agreement is achieved when we appropriately take the form-factor parameter at around the physically reasonable value R0(500MeV)1R_{0}\sim\left(500\,\text{MeV}\right)^{-1}.

We also analyzed and made a comment on the bb¯b\overline{b}-vector-meson counterpart Υ\Upsilon, namely RΥ=Γ(ΥB+B)/Γ(ΥB0B0¯)R_{\Upsilon}=\Gamma\!\left(\Upsilon\to B^{+}B^{-}\right)/\Gamma\!\bigl{(}\Upsilon\to B^{0}\overline{B^{0}}\bigr{)}, where the plane-wave calculation without considering the above-mentioned composite nature already agrees at the 0.32σ0.32\,\sigma level with the corresponding PDG result due to the smallness of the mass difference between B±B^{\pm} and B0B^{0}. The wave-packet result agrees well with the PDG result around the same value of R0R_{0}.

We mention that the same form factors can be formally multiplied on the ratio of the plane-wave decay rates in order to partially take into account the wave-packet effects, though the wave-packet approach is more comprehensive in describing quantum transitions. By doing so, around the same value of R0R_{0}, the plane-wave results can also be made to agree with the PDG ones.

In general, the shape of a wave-packet resonance deviates from the Briet-Wigner shape, where the magnitude of the deviation depends on the size of wave packets. For ϕ\phi and ψ\psi, experimental data is available, and we put constraints on the size. We found that when the size of the wave packets is small, the derivation from the Briet-Wigner shape tends to be sizable. Both for ϕ\phi and ψ\psi, wide ranges of the wave-packet size are consistent with the experimental data.

In the decay channels of the vector mesons, the non-relativistic approximation works fine, which simplifies the integrations in the SS-matrices and the final-state phase spaces in the wave-packet formalism. Many other quantum transitions in high-energy physics are relativistic, and it is worthwhile to establish the general method to perform such integrations without relying on the non-relativistic approximation. Also, analyzing resonant productions precisely requires the full transition probabilities, including production parts. Doing more dedicated analyses on resonant shapes will be another important task.

Acknowledgments

K.I. thanks for useful communications with Dr. Elliott D. Bloom, Dr. Walter Toki, and Dr. Tomasz Skwarnicki on the crystal-ball functions. K.N. thanks Dr. Akimasa Ishikawa for the useful information on experimental details on RΥR_{\Upsilon} and Dr. Ryosuke Sato for useful comments on form factors. This work is partially supported by the JSPS KAKENHI Grant Nos. JP21H01107 (K.I., O.J., K.N., K.O.) and JP19H01899 (K.O.).

Appendix

Appendix A Vector-meson form factor

We focus on the following form for a non-relativistic bound state V(QQ¯)V\!\left(Q\overline{Q}\right) composed of a heavy constituent quark QQ and its anti-particle Q¯\overline{Q} at a certain time:

|V(QQ¯),𝑷:=s1,s2d3𝒙1d3𝒙2ei𝑷𝒙1+𝒙22Fs1,s2(𝒙1𝒙2)Q(𝒙1,s1)Q¯(𝒙2,s2)|0,\displaystyle\ket{V(Q\overline{Q}),\bm{P}}:={\sum_{s_{1},s_{2}}}\int\text{d}^{3}\bm{x}_{1}\,\text{d}^{3}\bm{x}_{2}\,e^{i\bm{P}\cdot\frac{\bm{x}_{1}+\bm{x}_{2}}{2}}F_{s_{1},s_{2}}\!\left(\bm{x}_{1}-\bm{x}_{2}\right)Q^{\dagger}\!\left(\bm{x}_{1},s_{1}\right)\overline{Q}^{\dagger}\!\left(\bm{x}_{2},s_{2}\right)\ket{0}, (95)

where Fs1,s2(𝒙1𝒙2)F_{s_{1},s_{2}}\!\left(\bm{x}_{1}-\bm{x}_{2}\right) is a wave function for the bound state; Q(𝒙1,s1)Q^{\dagger}\!\left(\bm{x}_{1},s_{1}\right) and Q¯(𝒙2,s2)\overline{Q}^{\dagger}\!\left(\bm{x}_{2},s_{2}\right) are the Fourier transforms of the momentum-space creation operators of QQ and Q¯\overline{Q}, respectively, with 𝒙1,2\bm{x}_{1,2} and s1,2s_{1,2} being their positions and spins; and |0\ket{0} is the vacuum; see e.g. Ref. [40].

We consider the following matrix element

Q(𝒑1,s1)Q¯(𝒑2,s2)|V(QQ¯),𝑷\displaystyle\Braket{Q\left(\bm{p}_{1},s_{1}\right)\overline{Q}\left(\bm{p}_{2},s_{2}\right)}{V(Q\overline{Q}),\bm{P}} =d3𝒙1d3𝒙2ei𝑷𝒙1+𝒙22Fs1,s2(𝒙1𝒙2)(12π)6ei𝒑1𝒙1i𝒙2𝒑2\displaystyle=\int\text{d}^{3}\bm{x}_{1}\,\text{d}^{3}\bm{x}_{2}\,e^{i\bm{P}\cdot\frac{\bm{x}_{1}+\bm{x}_{2}}{2}}F_{s_{1},s_{2}}\!\left(\bm{x}_{1}-\bm{x}_{2}\right)\left(\frac{1}{\sqrt{2\pi}}\right)^{6}e^{-i\bm{p}_{1}\cdot\bm{x}_{1}-i\bm{x}_{2}\cdot\bm{p}_{2}}
=d3𝑿d3𝒓ei(𝑷𝒑1𝒑2)𝑿i(𝒑1𝒑22)𝒓1(2π)3Fs1,s2(𝒓)\displaystyle=\int\text{d}^{3}\bm{X}\,\text{d}^{3}\bm{r}\,e^{i\left(\bm{P}-\bm{p}_{1}-\bm{p}_{2}\right)\cdot\bm{X}-i\left(\frac{\bm{p}_{1}-\bm{p}_{2}}{2}\right)\cdot\bm{r}}\frac{1}{\left(2\pi\right)^{3}}F_{s_{1},s_{2}}\!\left(\bm{r}\right)
=δ3(𝑷𝒑1𝒑2)d3𝒓ei(𝒑1𝒑22)𝒓Fs1,s2(𝒓)=:F~s1,s2(𝒑1𝒑22),\displaystyle=\delta^{3}\!\left(\bm{P}-\bm{p}_{1}-\bm{p}_{2}\right)\underbrace{\int\text{d}^{3}\bm{r}\,e^{-i\left(\frac{\bm{p}_{1}-\bm{p}_{2}}{2}\right)\cdot\bm{r}}F_{s_{1},s_{2}}\left(\bm{r}\right)}_{=:\widetilde{F}_{s_{1},s_{2}}\!\left(\frac{\bm{p}_{1}-\bm{p}_{2}}{2}\right)}, (96)

where 𝑿:=(𝒙1+𝒙2)/2\bm{X}:=(\bm{x}_{1}+\bm{x}_{2})/2, 𝒓:=𝒙1𝒙2\bm{r}:=\bm{x}_{1}-\bm{x}_{2}, and the normalization of the form factor F~s1,s2(𝒑1𝒑22)\widetilde{F}_{s_{1},s_{2}}\!\left(\frac{\bm{p}_{1}-\bm{p}_{2}}{2}\right) is irrelevant for our purpose.282828 What is relevant is only the product of the normalization of the effective coupling and that of the form factor. Such a normalization factor will be dropped out of the final ratio of the decay probabilities under the isospin-symmetric limit (8).

Hereafter, we assume the separable form

F~s1,s2(𝒑1𝒑22)=Ss1,s2F~(𝒑1𝒑22)\displaystyle\widetilde{F}_{s_{1},s_{2}}\!\left(\frac{\bm{p}_{1}-\bm{p}_{2}}{2}\right)=S_{s_{1},s_{2}}\,\widetilde{F}\!\left(\frac{\bm{p}_{1}-\bm{p}_{2}}{2}\right) (97)

for the heavy and non-relativistic quarks QQ¯Q\overline{Q}, which have negligible spin-orbital angular momentum interaction. Further, we drop the spin structure Ss1,s2S_{s_{1},s_{2}}, which will be canceled out in the ratio of the neutral to charged rates, and focus on the momentum part.292929 Concretely, the orbital and total angular momenta J\ell_{J} is S1S_{1} (=1\ell=1), D1D_{1} (=2\ell=2), and S1S_{1} for ϕ(1020)\phi\!\left(1020\right), ψ(3770)\psi\!\left(3770\right) and Υ(4S)\Upsilon\!\left(4S\right), respectively; see e.g. “Quark Model” section in Ref. [1].

We adopt an approximate form of the wave function of the (ss-wave) ground state under a Coulomb potential in the position space:

F(𝒓)=N2πR0erR0r,\displaystyle F\!\left(\bm{r}\right)=\frac{N}{\sqrt{2\pi R_{0}}}\frac{e^{-\frac{r}{R_{0}}}}{r}, (98)

where r:=|𝒓|r:=\left|\bm{r}\right|, the parameter R0R_{0} describes a typical length scale of the bound state discussed below, and NN is the irrelevant normalization factor mentioned above. Its Fourier-transform is

F~(𝒑)\displaystyle\widetilde{F}\!\left(\bm{p}\right) =Nd3𝒓(2π)3/2ei𝒑𝒓F(𝒓)=NπR01𝒑2+1R02,\displaystyle=N\int{\text{d}^{3}\bm{r}\over\left(2\pi\right)^{3/2}}\,e^{-i\bm{p}\cdot\bm{r}}F\!\left(\bm{r}\right)={N\over\pi\sqrt{R_{0}}}{1\over\bm{p}^{2}+{1\over R_{0}^{2}}}, (99)

where we used 0drsin(pr)erR0=pp2+1R02\int_{0}^{\infty}\text{d}r\sin\!\left(pr\right)e^{-\frac{r}{R_{0}}}=\frac{p}{p^{2}+\frac{1}{R_{0}^{2}}}. In this paper, we choose N=π/R03/2N=\pi/R_{0}^{3/2} such that F~(𝟎)=1\widetilde{F}\!\left(\bm{0}\right)=1:

F~(𝒑)\displaystyle\widetilde{F}\!\left(\bm{p}\right) =1R021𝒑2+1R02.\displaystyle={1\over R_{0}^{2}}{1\over\bm{p}^{2}+{1\over R_{0}^{2}}}. (100)

This form is also introduced in Ref. [3] to cut off a UV divergence in the plane-wave computation. The treatment in Refs. [4, 5] is equivalent to taking this form factor to be unity.

Here, a vector meson V(QQ¯)V\!\left(Q\overline{Q}\right) decays into two light pseudo-scalar mesons P(Qq¯)P\!\left(Q\overline{q}\right) and P¯(qQ¯)\overline{P}\!\left(q\overline{Q}\right). We approximate the mass and momentum for each psuedoscalar PP (P¯\overline{P}) by those of the constituent quark QQ (Q¯\overline{Q}): mPmQm_{P}\simeq m_{Q} and 𝒑P𝒑1\bm{p}_{P}\simeq\bm{p}_{1} (𝒑P¯𝒑2\bm{p}_{\overline{P}}\simeq\bm{p}_{2}), respectively. In this paper, we focus on the situation where the masses of the two pseudo-scalar mesons are almost the same (due to the approximated flavor-isospin SU(2)SU(2) symmetry), and the mass relation is near the decay threshold,

mV2mP.\displaystyle m_{V}\approx 2\,m_{P}. (101)

Therefore, we can treat the process as a non-relativistic one, and thus, we conclude that

𝒑1𝒑2mP(𝑽1𝑽2),\displaystyle\bm{p}_{1}-\bm{p}_{2}\approx m_{P}\left(\bm{V}_{1}-\bm{V}_{2}\right), (102)

thereby,

F~(𝒑1𝒑22)\displaystyle\widetilde{F}\!\left(\frac{\bm{p}_{1}-\bm{p}_{2}}{2}\right) =1(R0(𝒑1𝒑2)2)2+1\displaystyle=\frac{1}{\left(\frac{R_{0}\left(\bm{p}_{1}-\bm{p}_{2}\right)}{2}\right)^{2}+1} (103)
1(R0mP(𝑽1𝑽2)2)2+1,\displaystyle\Rightarrow\frac{1}{\left(\frac{R_{0}m_{P}\left(\bm{V}_{1}-\bm{V}_{2}\right)}{2}\right)^{2}+1}, (104)

where 𝑽1\bm{V}_{1} and 𝑽2\bm{V}_{2} are the (non-relativistic) velocities of PP and P¯\overline{P}, respectively.

Finally, we reach the spin-independent dimensionless function suitable for our purpose,

F~(|𝑽1𝑽2|):=1(R0mP(𝑽1𝑽2)2)2+1.\displaystyle\widetilde{F}\!\left(\left|\bm{V}_{1}-\bm{V}_{2}\right|\right):=\frac{1}{\left(\frac{R_{0}m_{P}\left(\bm{V}_{1}-\bm{V}_{2}\right)}{2}\right)^{2}+1}. (105)

This is the form factor shown in Eq. (15) for the matrix elements of the meson decays (with mV2mPm_{V}\approx 2\,m_{P}) in the rest system.

Now we estimate a typical value of the parameter R0R_{0} in Eq. (98). The quarkonium potential can be approximated by a sum of the confining linear potential and the QCD Coulomb potential

V(r)=ras2αsr,\displaystyle V(r)=\frac{r}{a_{s}^{2}}-\frac{\alpha_{s}}{r}, (106)

where asa_{s} is known as as=1.95GeV1a_{s}=1.95\,\text{GeV}^{-1} [40] and αs\alpha_{s} is the QCD fine structure constant. For the domain where rrc:=asαs1.5GeV1r\lesssim r_{c}:=a_{s}\sqrt{\alpha_{s}}\simeq 1.5\,\text{GeV}^{-1},303030 Here, we have put αs0.6\alpha_{s}\simeq 0.6 at the scale 0.50.5 GeV [41]. the wave function can be approximated by the Coulomb form (98). One can estimate rr by equating the potential and kinetic energies, V(r)KQV\!\left(r\right)\sim K_{Q}, where KQmV2mQ=𝒪(10)MeVK_{Q}\sim m_{V}-2m_{Q}={\cal O}(10)\,\text{MeV}. Since KQK_{Q} is much smaller than the typical energy scale as1=0.5a_{s}^{-1}=0.5 GeV of the potential, the typical QQ¯Q\overline{Q} distance can be estimated by equating two terms in the right-hand side of Eq. (106): rrcr\sim r_{c}. The use of Coulomb wave function (98) is marginally justified, which suffices for our current consideration. See e.g. Ref. [42] for further refinement.

Finally, a typical value of the parameter in Eq. (98) is

R0rc1.5GeV1=0.0015MeV1=1660MeV.\displaystyle R_{0}\sim r_{c}\simeq 1.5\,\text{GeV}^{-1}=0.0015\,\text{MeV}^{-1}={1\over 660\,\text{MeV}}. (107)

Appendix B Details on calculations of PVPP¯P_{V\to P\overline{P}}

We present detailed computation to obtain the decay probability integrated over the final-state positions and momenta under the rest-frame assumption 𝑷0=0\bm{P}_{0}=0.

B.1 Variables under non-relativistic limit

As preparation, we show concrete forms of the kinetic variables and parameters, whose physical meaning is given in Section 2.1, under the non-relativistic limit |𝑽1,2|1\left|\bm{V}_{1,2}\right|\ll 1. At first, we define the ‘light-cone’ variables for later convenience:

𝑷±\displaystyle\bm{P}_{\pm} :=𝑷1±𝑷2,\displaystyle:=\bm{P}_{1}\pm\bm{P}_{2}, 𝑽±\displaystyle\bm{V}_{\pm} :=𝑽1±𝑽2.\displaystyle:=\bm{V}_{1}\pm\bm{V}_{2}. (108)

The kinetic variables are

𝑷0=0\displaystyle\bm{P}_{0}=0 𝑽0=0,\displaystyle\Rightarrow\bm{V}_{0}=0, 𝑷1\displaystyle\bm{P}_{1} mP𝑽1,\displaystyle\Rightarrow m_{P}\bm{V}_{1}, 𝑷2\displaystyle\bm{P}_{2} mP𝑽2,\displaystyle\Rightarrow m_{P}\bm{V}_{2},
E0=mV\displaystyle E_{0}=m_{V} mV,\displaystyle\Rightarrow m_{V}, E1\displaystyle E_{1} mP+mP2𝑽12,\displaystyle\Rightarrow m_{P}+\frac{m_{P}}{2}\bm{V}_{1}^{2}, E2\displaystyle E_{2} mP+mP2𝑽22,\displaystyle\Rightarrow m_{P}+\frac{m_{P}}{2}\bm{V}_{2}^{2},
|geff|2¯=g23(𝑷1𝑷2)2\displaystyle\overline{|g_{\text{eff}}|^{2}}={\frac{g^{2}}{3}}(\bm{P}_{1}-\bm{P}_{2})^{2} g23(mP2𝑽2),\displaystyle\Rightarrow{\frac{g^{2}}{3}}\left(m_{P}^{2}\bm{V}_{-}^{2}\right), (109)

where the symbol \Rightarrow represents the non-relativistic approximation, which we will take later. We also have

𝑽¯\displaystyle\overline{\bm{V}} =σs(𝑽1σP+𝑽2σP)=σsσP𝑽+,\displaystyle=\sigma_{s}\left(\frac{\bm{V}_{1}}{\sigma_{P}}+\frac{\bm{V}_{2}}{\sigma_{P}}\right)=\frac{\sigma_{s}}{\sigma_{P}}\bm{V}_{+}, (110)
𝑽2¯\displaystyle\overline{\bm{V}^{2}} =σs(𝑽12σP+𝑽22σP)=σs2σP(𝑽+2+𝑽2),\displaystyle=\sigma_{s}\left(\frac{\bm{V}_{1}^{2}}{\sigma_{P}}+\frac{\bm{V}_{2}^{2}}{\sigma_{P}}\right)=\frac{\sigma_{s}}{2\sigma_{P}}\left(\bm{V}_{+}^{2}+\bm{V}_{-}^{2}\right), (111)
δ𝑷\displaystyle\delta\bm{P} =𝑷1+𝑷2mP𝑽+,\displaystyle=\bm{P}_{1}+\bm{P}_{2}\Rightarrow m_{P}\bm{V}_{+}, (112)
δE\displaystyle\delta E =mV+E1+E2(mV2mP)+12mP(𝑽12+𝑽22)\displaystyle=-m_{V}+E_{1}+E_{2}\Rightarrow-\left(m_{V}-2m_{P}\right)+\frac{1}{2}m_{P}\left(\bm{V}_{1}^{2}+\bm{V}_{2}^{2}\right)
=(mV2mP)+14mP(𝑽+2+𝑽2),\displaystyle\phantom{=-E_{0}+E_{1}+E_{2}}\ =-\left(m_{V}-2m_{P}\right)+\frac{1}{4}m_{P}\left(\bm{V}_{+}^{2}+\bm{V}_{-}^{2}\right), (113)
δω\displaystyle\delta\omega =δE𝑽¯δ𝑷δEmPσsσP𝑽+2,\displaystyle=\delta E-\overline{\bm{V}}\cdot\delta\bm{P}\Rightarrow\delta E-m_{P}\frac{\sigma_{s}}{\sigma_{P}}\bm{V}_{+}^{2}, (114)
σs\displaystyle\sigma_{s} =σVσP2σV+σP,\displaystyle=\frac{\sigma_{V}\sigma_{P}}{2\sigma_{V}+\sigma_{P}}, (115)
σt\displaystyle\sigma_{t} =σs𝑽2¯(𝑽¯)2=σP12(𝑽+2+𝑽2)σsσP𝑽+2.\displaystyle=\frac{\sigma_{s}}{\overline{\bm{V}^{2}}-\left(\overline{\bm{V}}\right)^{2}}=\frac{\sigma_{P}}{\frac{1}{2}\left(\bm{V}_{+}^{2}+\bm{V}_{-}^{2}\right)-\frac{\sigma_{s}}{\sigma_{P}}\bm{V}_{+}^{2}}. (116)

Also, we define the variables V+V_{+} and VV_{-}

V+\displaystyle V_{+} :=|𝑽+|,\displaystyle:=\left|\bm{V}_{+}\right|, V\displaystyle V_{-} :=|𝑽|.\displaystyle:=\left|\bm{V}_{-}\right|. (117)

B.2 Bulk contribution

We compute the bulk contribution in Eq. (37). First, we perform the position integrals +d3𝑿1d3𝑿2\int_{-\infty}^{+\infty}\text{d}^{3}\bm{X}_{1}\text{d}^{3}\bm{X}_{2}. As in Ref. [11], we obtain

d3𝑿1d3𝑿2\displaystyle\text{d}^{3}\bm{X}_{1}\text{d}^{3}\bm{X}_{2} =d5ydy0,\displaystyle=\text{d}^{5}y\,\text{d}y_{0}, (118)
d5ye\displaystyle\int\text{d}^{5}y\,e^{-\mathcal{R}} =π5σt(σ0σ1σ2σs)31(δ𝑽1)2+(δ𝑽2)2,\displaystyle=\sqrt{\frac{\pi^{5}}{\sigma_{t}}\left(\frac{\sigma_{0}\sigma_{1}\sigma_{2}}{\sigma_{s}}\right)^{3}}\frac{1}{\sqrt{\left(\delta\bm{V}_{1}\right)^{2}+\left(\delta\bm{V}_{2}\right)^{2}}}, (119)
𝔗ΓVσt2\displaystyle{\mathfrak{T}-{\Gamma_{V}\sigma_{t}\over 2}} =y0(δ𝑽1)2+(δ𝑽2)2+,\displaystyle=-\frac{y_{0}}{\sqrt{\left(\delta\bm{V}_{1}\right)^{2}+\left(\delta\bm{V}_{2}\right)^{2}}}+\cdots, (120)
d𝔗\displaystyle\text{d}\mathfrak{T} =dy0(δ𝑽1)2+(δ𝑽2)2,\displaystyle=-\frac{\text{d}y_{0}}{\sqrt{\left(\delta\bm{V}_{1}\right)^{2}+\left(\delta\bm{V}_{2}\right)^{2}}}, (121)

where y0y_{0} becomes a flat direction under the (unphysical) no-decay limit ΓV0\Gamma_{V}\to 0 (as considered in [11]), while the other five directions are not flat directions irrespective of ΓV\Gamma_{V}:

d3𝑿1d3𝑿2eΓV(𝔗T0)+ΓV2σt4(W(𝔗))2\displaystyle\int\text{d}^{3}\bm{X}_{1}\text{d}^{3}\bm{X}_{2}e^{-\mathcal{R}-\Gamma_{V}\left(\mathfrak{T}-T_{0}\right)+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}}\bigl{(}W(\mathfrak{T})\bigr{)}^{2} =π5σt(σ0σ1σ2σs)3Tin+ΓVσt2Tout+ΓVσt2d𝔗eΓV(𝔗T0)+ΓV2σt4\displaystyle=\sqrt{\frac{\pi^{5}}{\sigma_{t}}\left(\frac{\sigma_{0}\sigma_{1}\sigma_{2}}{\sigma_{s}}\right)^{3}}\,\int_{{T_{\text{in}}+{\Gamma_{V}\sigma_{t}\over 2}}}^{{T_{\text{out}}+{\Gamma_{V}\sigma_{t}\over 2}}}\text{d}\mathfrak{T}\,e^{-\Gamma_{V}\left(\mathfrak{T}-T_{0}\right)+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}}
=π5σt(σ0σ1σ2σs)3ΓV1eΓV(TinT0)ΓV2σt4,\displaystyle=\sqrt{\frac{\pi^{5}}{\sigma_{t}}\left(\frac{\sigma_{0}\sigma_{1}\sigma_{2}}{\sigma_{s}}\right)^{3}}\,\Gamma_{V}^{-1}e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right){-\frac{\Gamma_{V}^{2}\sigma_{t}}{4}}}, (122)

where we also used Eq. (31) and took the limit ToutT_{\text{out}}\to\infty. The range of the integration is given by the bulk window function W(𝔗)W(\mathfrak{T}) in Eq. (31). After the position integrations, we obtain

dPVPP¯bulk\displaystyle\text{d}P^{\text{bulk}}_{V\to P\overline{P}} =|geff|2NV2ΓV1eΓV(TinT0)ΓV2σt412E0d3𝑷1(2π)32E1d3𝑷2(2π)32E2\displaystyle=\left|g_{\text{eff}}\right|^{2}N_{V}^{2}\Gamma_{V}^{-1}e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right){-\frac{\Gamma_{V}^{2}\sigma_{t}}{4}}}\frac{1}{2E_{0}}\frac{\text{d}^{3}\bm{P}_{1}}{(2\pi)^{3}2E_{1}}\frac{\text{d}^{3}\bm{P}_{2}}{(2\pi)^{3}2E_{2}}
×(2π)4(σtπ(σsπ)3/2eσt(δω)2σs(δ𝑷)2)|F~(V)|2,\displaystyle\quad\times\left(2\pi\right)^{4}\left(\sqrt{\frac{\sigma_{t}}{\pi}}\left(\frac{\sigma_{s}}{\pi}\right)^{3/2}e^{-\sigma_{t}\left(\delta\omega\right)^{2}-\sigma_{s}\left(\delta\bm{P}\right)^{2}}\right)\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}, (123)

where the damping factors σtπeσt(δω)2\sqrt{\sigma_{t}\over\pi}\,e^{-\sigma_{t}\left(\delta\omega\right)^{2}} and (σsπ)3/2eσs(δ𝑷)2\left(\sigma_{s}\over\pi\right)^{3/2}e^{-\sigma_{s}\left(\delta\bm{P}\right)^{2}} provide the approximate conservation for the mean energy and momentum, respectively.313131 Of course the energy-momentum conservation is fulfilled in itself as a fundamental physical law of nature, in particular, for each partial plane-wave component in the Fourier transforms. So far the expression (123) does not assume 𝑷0=0\bm{P}_{0}=0 nor the non-relativistic approximation given in Sec. B.1.

Next, we perform the momentum integrals under the saddle-point approximation with 𝑷0=0\bm{P}_{0}=0 in the non-relativistic approximation given in Sec. B.1:

PVPP¯bulk\displaystyle P^{\text{bulk}}_{V\to P\overline{P}} (NV2ΓV1eΓV(TinT0))g2312mV1(2π)24E1E2mP68(04πV+2dV+)(04πV2dV)\displaystyle\Rightarrow\left(N_{V}^{2}\Gamma_{V}^{-1}e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right)}\right)\frac{g^{2}}{3}\frac{1}{2m_{V}}\frac{1}{(2\pi)^{2}4E_{1}E_{2}}\frac{m_{P}^{6}}{8}\left(\int_{0}^{\infty}4\pi V_{+}^{2}\text{d}V_{+}\right)\left(\int_{0}^{\infty}4\pi V_{-}^{2}\text{d}V_{-}\right)
×1π2σt1/2σs3/2(mP2V2)eFbulk(V+,V)|F~(V)|2,\displaystyle\quad\times\frac{1}{\pi^{2}}\sigma_{t}^{1/2}\sigma_{s}^{3/2}\left(m_{P}^{2}V_{-}^{2}\right)e^{-F_{\text{bulk}}(V_{+},V_{-})}\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}, (124)

where the factor 1/8=1/231/8=1/2^{3} is from the Jacobian and the exponent is

Fbulk(V+,V)\displaystyle F_{\text{bulk}}(V_{+},V_{-}) :=σs(δ𝑷)2+σt(δω)2+ΓV2σt4\displaystyle:=\sigma_{s}\left(\delta\bm{P}\right)^{2}+\sigma_{t}\left(\delta\omega\right)^{2}{+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}}
=σsmP2V+2\displaystyle=\sigma_{s}m_{P}^{2}V_{+}^{2}
+2σP(12σsσP)V+2+V2{[mP(14σsσP)V+2+14mPV2(mV2mP)]2+ΓV24}.\displaystyle\quad+\frac{2\sigma_{P}}{\left(1-\frac{2\sigma_{s}}{\sigma_{P}}\right)V_{+}^{2}+V_{-}^{2}}\left\{\left[m_{P}\left(\frac{1}{4}-\frac{\sigma_{s}}{\sigma_{P}}\right)V_{+}^{2}+\frac{1}{4}m_{P}V_{-}^{2}-\left(m_{V}-2m_{P}\right)\right]^{2}+{{\Gamma_{V}^{2}\over 4}}\right\}. (125)

We see that

12σsσP=12σV/σP+1>0,\displaystyle 1-\frac{2\sigma_{s}}{\sigma_{P}}=\frac{1}{2\sigma_{V}/\sigma_{P}+1}>0, (126)

which implies

σP2σs>0.\displaystyle\sigma_{P}-{2\sigma_{s}}>0. (127)

Also, we see that

σP4σs=σP(σP2σs)σP+2σs>0,\displaystyle\sigma_{P}-4\sigma_{s}=\frac{\sigma_{P}\left(\sigma_{P}-2\sigma_{s}\right)}{\sigma_{P}+2\sigma_{s}}>0, (128)

which leads to

14σsσP=14σP(σP4σs)>0.\displaystyle\frac{1}{4}-\frac{\sigma_{s}}{\sigma_{P}}=\frac{1}{4\sigma_{P}}\left(\sigma_{P}-4\sigma_{s}\right)>0. (129)

Thereby,

eFbulk\displaystyle e^{-F_{\text{bulk}}} V+0,\displaystyle\underbrace{\to}_{V_{+}\to\infty}0, eFbulk\displaystyle e^{-F_{\text{bulk}}} V0.\displaystyle\underbrace{\to}_{V_{-}\to\infty}0. (130)

The stationary point (V+s,Vs)(V_{+}^{s},V_{-}^{s}) that satisfies

Fbulk(V+s,Vs)V+=0andFbulk(V+s,Vs)V=0\displaystyle\frac{\partial F_{\text{bulk}}(V_{+}^{s},V_{-}^{s})}{\partial V_{+}}=0\quad\text{and}\quad\frac{\partial F_{\text{bulk}}(V_{+}^{s},V_{-}^{s})}{\partial V_{-}}=0 (131)

is found to be

(V+s,Vs)\displaystyle\left(V_{+}^{s},V_{-}^{s}\right) {(0,2((mV2mP)2+ΓV24)1/4mP),\displaystyle\in\left\{\left(0,\,\frac{2\left(\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}\right)^{1/4}}{\sqrt{m_{P}}}\right)\right., (0,2((mV2mP)2+ΓV24)1/4mP),\displaystyle\left(0,\,\frac{-2\left(\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}\right)^{1/4}}{\sqrt{m_{P}}}\right),
(0,2i((mV2mP)2+ΓV24)1/4mP),\displaystyle\qquad\left(0,\,\frac{2i\left(\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}\right)^{1/4}}{\sqrt{m_{P}}}\right), (0,2i((mV2mP)2+ΓV24)1/4mP),\displaystyle\left(0,\,\frac{-2i\left(\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}\right)^{1/4}}{\sqrt{m_{P}}}\right),
(2((mV2mP)2+ΓV24)1/4mP, 0),\displaystyle\qquad\left(\frac{2\left(\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}\right)^{1/4}}{\sqrt{m_{P}}},\,0\right), (2((mV2mP)2+ΓV24)1/4mP, 0),\displaystyle\left(\frac{-2\left(\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}\right)^{1/4}}{\sqrt{m_{P}}},\,0\right),
(2i((mV2mP)2+ΓV24)1/4mP, 0),\displaystyle\qquad\left(\frac{2i\left(\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}\right)^{1/4}}{\sqrt{m_{P}}},\,0\right), (2i((mV2mP)2+ΓV24)1/4mP, 0)}.\displaystyle\left.\left(\frac{-2i\left(\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}\right)^{1/4}}{\sqrt{m_{P}}},\,0\right)\right\}. (132)

Now, we focus on the kinetic region (V+0V_{+}\simeq 0 and V>0V_{-}>0) and thus only the first one is relevant in our calculation,

(V+B,VB)\displaystyle\left(V_{+}^{\text{B}},V_{-}^{\text{B}}\right) :=(0,2((mV2mP)2+ΓV24)1/4mP=2[(mV2mP)2mP2+ΓV24mP2]1/4).\displaystyle:=\left(0,\,\frac{2\left(\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}\right)^{1/4}}{\sqrt{m_{P}}}=2\left[\frac{\left(m_{V}-2m_{P}\right)^{2}}{m_{P}^{2}}+{\Gamma_{V}^{2}\over 4m_{P}^{2}}\right]^{1/4}\right). (133)

Note that

Fbulk0:=Fbulk(V+B,VB)\displaystyle{F_{\text{bulk}}^{0}}:=F_{\text{bulk}}\left(V_{+}^{\text{B}},V_{-}^{\text{B}}\right) =mPσP((mV2mP)+(mV2mP)2+ΓV24),\displaystyle=m_{P}\sigma_{P}\left(-\left(m_{V}-2m_{P}\right)+\sqrt{\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}}\right), (134)

which goes to zero in the (unphysical) limit ΓV0\Gamma_{V}\to 0. Also, we can see

2FbulkV+2|(V+B,VB)\displaystyle\left.\frac{\partial^{2}F_{\text{bulk}}}{\partial V_{+}^{2}}\right|_{\left(V_{+}^{\text{B}},V_{-}^{\text{B}}\right)} =2mP2σs[12+mV2mP2(mV2mP)2+ΓV24]=:2mP2σsAbulk(>0),\displaystyle=2m_{P}^{2}\sigma_{s}\left[\frac{1}{2}+\frac{m_{V}-2m_{P}}{2\sqrt{\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 4}}}\right]=:2m_{P}^{2}\sigma_{s}{A_{\text{bulk}}}\quad(>0),
2FbulkV2|(V+B,VB)\displaystyle\left.\frac{\partial^{2}F_{\text{bulk}}}{\partial V_{-}^{2}}\right|_{\left(V_{+}^{\text{B}},V_{-}^{\text{B}}\right)} =mP2σP(>0),\displaystyle=m_{P}^{2}\sigma_{P}\quad(>0),
2FbulkV+V|(V+B,VB)\displaystyle\left.\frac{\partial^{2}F_{\text{bulk}}}{\partial V_{+}\partial V_{-}}\right|_{\left(V_{+}^{\text{B}},V_{-}^{\text{B}}\right)} =0,\displaystyle=0, (135)

where Abulk1A_{\text{bulk}}\to 1 under the (unphysical) limit ΓV0\Gamma_{V}\to 0.

We perform the integrals under the saddle-point approximation (putting E1=E2=mPE_{1}=E_{2}=m_{P} in the overall factor) as follows:

PVPP¯bulk\displaystyle P^{\text{bulk}}_{V\to P\overline{P}} gV2312mV1(2π)24mP2mP68(04πV+2dV+)(04π(VB)2dV)\displaystyle\simeq\frac{g_{V}^{2}}{3}\frac{1}{2m_{V}}\frac{1}{(2\pi)^{2}4m_{P}^{2}}\,\frac{m_{P}^{6}}{8}\left(\int_{0}^{\infty}4\pi V_{+}^{2}\text{d}V_{+}\right)\left(\int_{0}^{\infty}4\pi\left(V_{-}^{\text{B}}\right)^{2}\text{d}V_{-}\right)
×1π2[σt1/2]V+V+B,VVBσs3/2[mP2(VB)2]\displaystyle\quad\times\frac{1}{\pi^{2}}\left[\sigma_{t}^{1/2}\right]_{V_{+}\to V_{+}^{\text{B}},\,V_{-}\to{V_{-}^{\text{B}}}}\,\sigma_{s}^{3/2}\left[m_{P}^{2}\left(V_{-}^{\text{B}}\right)^{2}\right]
×eFbulk0e12(2mP2σs)Abulk(V+V+B)2e12(mP2σP)(VVB)2(NV2ΓVeΓV(TinT0))|F~(VB)|2\displaystyle\quad\times{e^{-F^{0}_{\text{bulk}}}}e^{-\frac{1}{2}\left(2m_{P}^{2}\sigma_{s}\right){A_{\text{bulk}}}\left(V_{+}-V_{+}^{\text{B}}\right)^{2}}e^{-\frac{1}{2}\left(m_{P}^{2}\sigma_{P}\right)\left(V_{-}-V_{-}^{\text{B}}\right)^{2}}\left({N_{V}^{2}\over\Gamma_{V}}e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right)}\right)\left|\widetilde{F}\!\left(V_{-}^{\text{B}}\right)\right|^{2}
=g2mP3NV2eΓV(TinT0)12πmVmP2mPΓV[(mV2mP)2mP2+ΓV24mP2]3/4\displaystyle=\frac{g^{2}m_{P}^{3}N_{V}^{2}e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right)}}{12\pi m_{V}m_{P}^{2}}\,{m_{P}\over\Gamma_{V}}\left[\frac{\left(m_{V}-2m_{P}\right)^{2}}{m_{P}^{2}}+{\Gamma_{V}^{2}\over 4m_{P}^{2}}\right]^{3/4}
×12[1+erf(mPσPVB2)]eFbulk0Abulk3/2|F~(VB)|2,\displaystyle\qquad\times\frac{1}{2}\left[1+\text{erf}\left(\frac{m_{P}\sqrt{\sigma_{P}}{V_{-}^{\text{B}}}}{\sqrt{2}}\right)\right]\ {\frac{e^{-F^{0}_{\text{bulk}}}}{A_{\text{bulk}}^{3/2}}}\left|\widetilde{F}\!\left(V_{-}^{\text{B}}\right)\right|^{2}, (136)

where we have used the formulas for a>0a>0,

04πr2ea2r2dr\displaystyle\int_{0}^{\infty}4\pi\,r^{2}e^{-\frac{a}{2}r^{2}}\text{d}r =22π3/2a3/2,\displaystyle=\frac{2\sqrt{2}\,\pi^{3/2}}{a^{3/2}}, (137)
0ea2(rr0)2dr\displaystyle\int_{0}^{\infty}e^{-\frac{a}{2}\left(r-r_{0}\right)^{2}}\text{d}r =π2a[1+erf(ar02)],\displaystyle=\sqrt{\frac{\pi}{2a}}\left[1+\text{erf}\left(\frac{\sqrt{a}r_{0}}{\sqrt{2}}\right)\right], (138)

and the Taylor expansion around one of the stationary points (V+,V)=(V+s,Vs)(V_{+},V_{-})=(V_{+}^{s},V_{-}^{s})

F(V+,V)F(V+s,Vs)+12i,j=+,2F(V+s,Vs)ViVj(ViVis)(VjVjs).\displaystyle F(V_{+},V_{-})\simeq F(V_{+}^{s},V_{-}^{s})+\frac{1}{2}\sum_{i,j=+,-}\frac{\partial^{2}F(V_{+}^{s},V_{-}^{s})}{\partial V_{i}\,\partial V_{j}}\left(V_{i}-V_{i}^{s}\right)\left(V_{j}-V_{j}^{s}\right). (139)

B.3 Boundary contribution

We compute the boundary contribution in Eq. (37):

d3𝑿1d3𝑿2eΓV(𝔗Tin)+ΓV2σt4(𝔗TinΓVσt2)2σt2σtπ1(𝔗TinΓVσt2)2+(σtδω)2\displaystyle\int\text{d}^{3}\bm{X}_{1}\text{d}^{3}\bm{X}_{2}e^{-{\cal R}-\Gamma_{V}\left(\mathfrak{T}-T_{\text{in}}\right)+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}-\frac{\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)^{2}}{\sigma_{t}}}\frac{2\sigma_{t}}{\pi}\frac{1}{\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)^{2}+\left(\sigma_{t}\delta\omega\right)^{2}}
=π5σt(σ0σ1σ2σs)3eΓV(TinT0)2σtπd𝔗e1σt(𝔗)21(𝔗ΓVσt2)2+(σtδω)2,\displaystyle=\sqrt{\frac{\pi^{5}}{\sigma_{t}}\left(\frac{\sigma_{0}\sigma_{1}\sigma_{2}}{\sigma_{s}}\right)^{3}}e^{-\Gamma_{V}\left(T_{\text{in}}-T_{0}\right)}\frac{2\sigma_{t}}{\pi}\int_{-\infty}^{\infty}\text{d}\mathfrak{T}^{\prime}e^{-\frac{1}{\sigma_{t}}\left(\mathfrak{T}^{\prime}\right)^{2}}\frac{1}{\left(\mathfrak{T}^{\prime}-{\Gamma_{V}\sigma_{t}\over 2}\right)^{2}+\left(\sigma_{t}\delta\omega\right)^{2}}, (140)

where 𝔗:=𝔗Tin\mathfrak{T}^{\prime}:=\mathfrak{T}-T_{\text{in}} and the range of the integration is (,+)(-\infty,+\infty) since there exists no window function W(𝔗)W(\mathfrak{T}) for this boundary term other than the exponential factor.

Now we perform the Taylor expansion of 𝔗\mathfrak{T}^{\prime} around the saddle point 𝔗=0\mathfrak{T}^{\prime}=0 up to the second order

e1σt(𝔗)2(𝔗ΓVσt2)2+(σtδω)2\displaystyle\frac{e^{-\frac{1}{\sigma_{t}}\left(\mathfrak{T}^{\prime}\right)^{2}}}{\left(\mathfrak{T}^{\prime}-{\Gamma_{V}\sigma_{t}\over 2}\right)^{2}+\left(\sigma_{t}\delta\omega\right)^{2}} (1(ΓVσt2)2+(σtδω)2+2(ΓVσt2)𝔗((ΓVσt2)2+(σtδω)2)2\displaystyle\simeq\left(\frac{1}{\left({\Gamma_{V}\sigma_{t}\over 2}\right)^{2}+\left(\sigma_{t}\delta\omega\right)^{2}}+\frac{2\left({\Gamma_{V}\sigma_{t}\over 2}\right)\mathfrak{T}^{\prime}}{\left(\left({\Gamma_{V}\sigma_{t}\over 2}\right)^{2}+\left(\sigma_{t}\delta\omega\right)^{2}\right)^{2}}\right.
+(3(ΓVσt2)2(σtδω)2)(𝔗)2((ΓVσt2)2+(σtδω)2)3)e1σt(𝔗)2,\displaystyle\quad+\left.\frac{\left(3\left({\Gamma_{V}\sigma_{t}\over 2}\right)^{2}-\left(\sigma_{t}\delta\omega\right)^{2}\right)\left(\mathfrak{T}^{\prime}\right)^{2}}{\left(\left({\Gamma_{V}\sigma_{t}\over 2}\right)^{2}+\left(\sigma_{t}\delta\omega\right)^{2}\right)^{3}}\right)e^{-\frac{1}{\sigma_{t}}\left(\mathfrak{T}^{\prime}\right)^{2}}, (141)

to obtain

d𝔗e1σt(𝔗)2(𝔗ΓVσt2)2+(σtδω)2\displaystyle\int_{-\infty}^{\infty}\text{d}\mathfrak{T}^{\prime}\frac{e^{-\frac{1}{\sigma_{t}}\left(\mathfrak{T}^{\prime}\right)^{2}}}{\left(\mathfrak{T}^{\prime}-{\Gamma_{V}\sigma_{t}\over 2}\right)^{2}+\left(\sigma_{t}\delta\omega\right)^{2}} πσtσt2(1(ΓV2)2+(δω)2+(3(ΓV2)2(δω)2)((ΓV2)2+(δω)2)32σt),\displaystyle\simeq\frac{\sqrt{\pi\sigma_{t}}}{\sigma_{t}^{2}}\left(\frac{1}{\left({\Gamma_{V}\over 2}\right)^{2}+\left(\delta\omega\right)^{2}}+\frac{\left(3\left({\Gamma_{V}\over 2}\right)^{2}-\left(\delta\omega\right)^{2}\right)}{\left(\left({\Gamma_{V}\over 2}\right)^{2}+\left(\delta\omega\right)^{2}\right)^{3}2\sigma_{t}}\right), (142)

which yields

PVPP¯bdry\displaystyle P^{\text{bdry}}_{V\to P\overline{P}} (NV2eΓV(TinT0))gV2312mV1(2π)24E1E2mP68(04πV+2dV+)(04πdV)\displaystyle\simeq\left(N_{V}^{2}e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right)}\right)\frac{g_{V}^{2}}{3}\frac{1}{2m_{V}}\frac{1}{(2\pi)^{2}4E_{1}E_{2}}\frac{m_{P}^{6}}{8}\left(\int_{0}^{\infty}4\pi V_{+}^{2}\text{d}V_{+}\right)\left(\int_{0}^{\infty}4\pi\text{d}V_{-}\right)
×12π5/2σs3/2mP2eσsmP2V+2\displaystyle\quad\times\frac{1}{2\pi^{5/2}}\sigma_{s}^{3/2}\,{m_{P}^{2}}\,e^{-\sigma_{s}m_{P}^{2}V_{+}^{2}}
×{V4(ΓV2)2+(δω)2+V4[3(ΓV2)2(δω)2][(ΓV2)2+(δω)2]32σt}|F~(V)|2.\displaystyle\quad\times\left\{\frac{V_{-}^{4}}{\left({\Gamma_{V}\over 2}\right)^{2}+\left(\delta\omega\right)^{2}}+\frac{V_{-}^{4}\left[3\left({\Gamma_{V}\over 2}\right)^{2}-\left(\delta\omega\right)^{2}\right]}{\left[\left({\Gamma_{V}\over 2}\right)^{2}+\left(\delta\omega\right)^{2}\right]^{3}2\sigma_{t}}\right\}\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}. (143)

There is no VV_{-} in the exponent, and we will perform the numerical computation for the VV_{-} integral. On the other hand, the saddle point of V+V_{+} is located at V+=0V_{+}=0. With it in mind, we approximate the polynomial part of the integrand by setting V+=0V_{+}=0 other than V+2V_{+}^{2}:

PVPP¯bdry\displaystyle P^{\text{bdry}}_{V\to P\overline{P}} (NV2eΓV(TinT0))gV2312mV1(2π)24mP2mP6804πV+2dV+04πdV\displaystyle\simeq\left(N_{V}^{2}e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right)}\right)\frac{g_{V}^{2}}{3}\frac{1}{2m_{V}}\frac{1}{(2\pi)^{2}4m_{P}^{2}}\frac{m_{P}^{6}}{8}\int_{0}^{\infty}4\pi V_{+}^{2}\text{d}V_{+}\int_{0}^{\infty}4\pi\text{d}V_{-}
×12π5/2σs3/2mP2eσsmP2V+2(mP4)2f~bdry(V),\displaystyle\quad\times\frac{1}{2\pi^{5/2}}\sigma_{s}^{3/2}m_{P}^{2}e^{-\sigma_{s}m_{P}^{2}V_{+}^{2}}\left(\frac{m_{P}}{4}\right)^{-2}\widetilde{f}_{\text{bdry}}\!\left(V_{-}\right), (144)

where we define a dimensionless function

f~bdry(V)\displaystyle\widetilde{f}_{\text{bdry}}\!\left(V_{-}\right) :=(mP4)2{V4(ΓV2)2+(δω)2+V4[3(ΓV2)2(δω)2][(ΓV2)2+(δω)2]32σt}V+0|F~(V)|2\displaystyle:=\left(\frac{m_{P}}{4}\right)^{2}\left\{\frac{V_{-}^{4}}{\left({\Gamma_{V}\over 2}\right)^{2}+\left(\delta\omega\right)^{2}}+\frac{V_{-}^{4}\left[3\left({\Gamma_{V}\over 2}\right)^{2}-\left(\delta\omega\right)^{2}\right]}{\left[\left({\Gamma_{V}\over 2}\right)^{2}+\left(\delta\omega\right)^{2}\right]^{3}2\sigma_{t}}\right\}_{V_{+}\to 0}\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}
={V4(V24mV2mPmP)2+42mP2(ΓV24)\displaystyle=\left\{\frac{V_{-}^{4}}{\left(V_{-}^{2}-4\frac{m_{V}-2m_{P}}{m_{P}}\right)^{2}+\frac{4^{2}}{m_{P}^{2}}\left({\Gamma_{V}^{2}\over 4}\right)}\right.
V4([V24mV2mPmP]234ΓV2mP2)2(σt|V+0)(mP4)2[(V24mV2mPmP)2+4ΓV2mP2]3}|F~(V)|2,\displaystyle\qquad-\left.\frac{V_{-}^{4}\left(\left[V_{-}^{2}-4\frac{m_{V}-2m_{P}}{m_{P}}\right]^{2}-3\frac{4\Gamma_{V}^{2}}{m_{P}^{2}}\right)}{2\left(\sigma_{t}|_{V_{+}\to 0}\right)\left(\frac{m_{P}}{4}\right)^{2}\left[\left(V_{-}^{2}-4\frac{m_{V}-2m_{P}}{m_{P}}\right)^{2}+\frac{4\Gamma_{V}^{2}}{m_{P}^{2}}\right]^{3}}\right\}\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}, (145)

with

σt|V+0:=2σPV2.\displaystyle\sigma_{t}|_{V_{+}\to 0}:=\frac{2\sigma_{P}}{V_{-}^{2}}. (146)

Now we execute the V+V_{+} integral using the formula (137):

PVPP¯bdry\displaystyle P^{\text{bdry}}_{V\to P\overline{P}} =g2mP3NV2eΓV(TinT0)12πmVmP2Ibdry2π,\displaystyle={\frac{g^{2}m_{P}^{3}N_{V}^{2}e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right)}}{12\pi m_{V}m_{P}^{2}}}\frac{I_{\text{bdry}}}{2\pi}, (147)

where the integral

Ibdry\displaystyle I_{\text{bdry}} :=0dVf~bdry(V)\displaystyle:=\int_{0}^{\infty}\text{d}V_{-}\widetilde{f}_{\text{bdry}}\!\left(V_{-}\right) (148)

is convergent.

B.4 Interference contribution

We compute the interference contribution in Eq. (37). We focus on the part including the factor

e+iδω(𝔗TinΓVσt2)𝔗TinΓVσt2iσtδω,\displaystyle\frac{e^{+i\delta\omega\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)}}{\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}-i\sigma_{t}\delta\omega},

since the other part can be obtained by taking complex conjugation. At first, we perform the square completion of the 𝔗\mathfrak{T} part:

Iintf:=\displaystyle I_{\text{intf}}:=\ d3𝑿1d3𝑿2eΓV(𝔗T0)+ΓV2σt4(𝔗TinΓVσt2)22σt+iδω(𝔗TinΓVσt2)1𝔗TinΓVσt2iσtδωW(𝔗)\displaystyle\int\text{d}^{3}\bm{X}_{1}\text{d}^{3}\bm{X}_{2}e^{-{\cal R}-\Gamma_{V}\left(\mathfrak{T}-{T_{0}}\right)+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}-\frac{\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)^{2}}{{\color[rgb]{0,0,1}2}\sigma_{t}}+i\delta\omega\left(\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}\right)}\frac{1}{\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}-i\sigma_{t}\delta\omega}W\!\left(\mathfrak{T}\right)
=\displaystyle=\ π5σt(σ0σ1σ2σs)3Tin+ΓVσt2Tout+ΓVσt2d𝔗e12σt(𝔗(TinΓVσt2+iσtδω))2ΓV(TinT0)12σt(δω)2+ΓV2σt4iΓVσtδω\displaystyle\sqrt{\frac{\pi^{5}}{\sigma_{t}}\left(\frac{\sigma_{0}\sigma_{1}\sigma_{2}}{\sigma_{s}}\right)^{3}}\int_{T_{\text{in}}+{\Gamma_{V}\sigma_{t}\over 2}}^{T_{\text{out}}+{\Gamma_{V}\sigma_{t}\over 2}}\text{d}\mathfrak{T}e^{-\frac{1}{2\sigma_{t}}\left(\mathfrak{T}-\left(T_{\text{in}}{-}{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega\right)\right)^{2}-\Gamma_{V}\left(T_{\text{in}}-T_{0}\right)-\frac{1}{2}\sigma_{t}\left(\delta\omega\right)^{2}+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}-i\Gamma_{V}\sigma_{t}\delta\omega}
×1𝔗TinΓVσt2iσtδω\displaystyle\quad\times\frac{1}{\mathfrak{T}-T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}-i\sigma_{t}\delta\omega}
\displaystyle\to\ π5σt(σ0σ1σ2σs)3Tind𝔗e12σt(𝔗(TinΓVσt2+iσtδω))2ΓV(TinT0)12σt(δω)2+ΓV2σt4iΓVσtδω\displaystyle\sqrt{\frac{\pi^{5}}{\sigma_{t}}\left(\frac{\sigma_{0}\sigma_{1}\sigma_{2}}{\sigma_{s}}\right)^{3}}\int_{T_{\text{in}}}^{\infty}\text{d}\mathfrak{T}^{\prime}e^{-\frac{1}{2\sigma_{t}}\left(\mathfrak{T}^{\prime}-\left(T_{\text{in}}{-{\Gamma_{V}\sigma_{t}\over 2}}+i\sigma_{t}\delta\omega\right)\right)^{2}-\Gamma_{V}\left(T_{\text{in}}-T_{0}\right)-\frac{1}{2}\sigma_{t}\left(\delta\omega\right)^{2}+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}-i\Gamma_{V}\sigma_{t}\delta\omega}
×1𝔗Tiniσtδω,\displaystyle\quad\times\frac{1}{\mathfrak{T}^{\prime}-T_{\text{in}}-i\sigma_{t}\delta\omega}, (149)

where in the last fine, we changed the variable to 𝔗:=𝔗ΓVσt/2\mathfrak{T}^{\prime}:=\mathfrak{T}-\Gamma_{V}\sigma_{t}/2 and took the limit ToutT_{\text{out}}\to\infty.

In order to use the analytic formula for σt>0\sigma_{t}>0 and α\alpha\in\mathbb{C},323232 One of the necessary conditions for this relation is “(Tin+α)&((Tin)(α))&((Tin)(α))\left(T_{\text{in}}+\alpha\not\in\mathbb{R}\right)\&\left(\Im\!\left(T_{\text{in}}\right)\not=\Im\!\left(\alpha\right)\right)\&\left(\Re\!\left(T_{\text{in}}\right)\geq\Re\!\left(\alpha\right)\right)”. In our case, the set of these conditions is manifestly fulfilled since α\alpha corresponds to “TinΓVσt/2+iσtδωT_{\text{in}}-\Gamma_{V}\sigma_{t}/2+i\sigma_{t}\delta\omega”.

Tindt1tαe12σt(tα)2=12Ei(12σt(Tinα)2),\displaystyle\int_{T_{\text{in}}}^{\infty}\text{d}t\frac{1}{t-\alpha}e^{-\frac{1}{2\sigma_{t}}\left(t-\alpha\right)^{2}}=-\frac{1}{2}\,\operatorname{Ei}\!\left(-\frac{1}{2\sigma_{t}}\left(T_{\text{in}}-\alpha\right)^{2}\right), (150)

where Ei(z)\operatorname{Ei}\!\left(z\right) is the exponential integral function defined by the principal value of

Ei(z):=zdtett,\displaystyle\operatorname{Ei}\!\left(z\right):=-\int_{-z}^{\infty}\text{d}t\frac{e^{-t}}{t}, (151)

we add an extra term in the denominator of the integrand of IintfI_{\text{intf}} such that

Iintf\displaystyle I_{\text{intf}} π5σt(σ0σ1σ2σs)3Tind𝔗e12σt(𝔗(TinΓVσt2+iσtδω))2ΓV(TinT0)12σt(δω)2+ΓV2σt4iΓVσtδω\displaystyle\sim\sqrt{\frac{\pi^{5}}{\sigma_{t}}\left(\frac{\sigma_{0}\sigma_{1}\sigma_{2}}{\sigma_{s}}\right)^{3}}\int_{T_{\text{in}}}^{\infty}\text{d}\mathfrak{T}^{\prime}e^{-\frac{1}{2\sigma_{t}}\left(\mathfrak{T}^{\prime}-\left(T_{\text{in}}{-{\Gamma_{V}\sigma_{t}\over 2}}+i\sigma_{t}\delta\omega\right)\right)^{2}-\Gamma_{V}\left(T_{\text{in}}-T_{0}\right)-\frac{1}{2}\sigma_{t}\left(\delta\omega\right)^{2}+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}-i\Gamma_{V}\sigma_{t}\delta\omega}
×1𝔗(TinΓVσt2+iσtδω),\displaystyle\quad\times\frac{1}{\mathfrak{T}^{\prime}-{\left(T_{\text{in}}-{\Gamma_{V}\sigma_{t}\over 2}+i\sigma_{t}\delta\omega\right)}}, (152)

which would underestimate the integral to some extent. Now, we reach the following analytic form

Iintfπ5σt(σ0σ1σ2σs)3eΓV(TinT0)12σt(δω)2+ΓV2σt4iΓVσtδω×(12Ei(X)),\displaystyle I_{\text{intf}}\sim\sqrt{\frac{\pi^{5}}{\sigma_{t}}\left(\frac{\sigma_{0}\sigma_{1}\sigma_{2}}{\sigma_{s}}\right)^{3}}e^{-\Gamma_{V}\left(T_{\text{in}}-T_{0}\right)-\frac{1}{2}\sigma_{t}\left(\delta\omega\right)^{2}+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}-i\Gamma_{V}\sigma_{t}\delta\omega}\times\left(-\frac{1}{2}\operatorname{Ei}\!\left(-X^{\prime}\right)\right), (153)

with

X:=σt2(ΓV+iδω)2.\displaystyle X^{\prime}:=\frac{\sigma_{t}}{2}\left(\Gamma_{V}+i\delta\omega\right)^{2}. (154)

Here, we will take the leading term of the expansion around infinity for XX^{\prime},

Ei(X)=eX+𝒪(1X2)[1X+𝒪(1X2)],\displaystyle-\operatorname{Ei}\!\left(-X^{\prime}\right)=e^{-X^{\prime}+{\cal O}\left(\frac{1}{{X^{\prime}}^{2}}\right)}\left[\frac{1}{X^{\prime}}+{\cal O}\!\left(\frac{1}{{X^{\prime}}^{2}}\right)\right], (155)

which leads to

Iintf+(Iintf)\displaystyle I_{\text{intf}}+\left(I_{\text{intf}}\right)^{\ast} 12π5σt(σ0σ1σ2σs)3eΓV(TinT0)ΓV2σt4[e2iΓVσtδωσt2(ΓV+iδω)2+e2iΓVσtδωσt2(ΓViδω)2],\displaystyle\sim\frac{1}{2}\sqrt{\frac{\pi^{5}}{\sigma_{t}}\left(\frac{\sigma_{0}\sigma_{1}\sigma_{2}}{\sigma_{s}}\right)^{3}}e^{-\Gamma_{V}\left(T_{\text{in}}-T_{0}\right)-\frac{\Gamma_{V}^{2}\sigma_{t}}{4}}\left[\frac{e^{-2i\Gamma_{V}\sigma_{t}\delta\omega}}{{\sigma_{t}\over 2}\left(\Gamma_{V}+i\delta\omega\right)^{2}}+\frac{e^{2i\Gamma_{V}\sigma_{t}\delta\omega}}{{\sigma_{t}\over 2}\left(\Gamma_{V}-i\delta\omega\right)^{2}}\right], (156)

and

dPVPP¯intf\displaystyle dP^{\text{intf}}_{V\to P\overline{P}} |geff|2NV212E0d3𝑷1(2π)32E1d3𝑷2(2π)32E2(2π)4(σtπeσt(δω)2)[(σsπ)3/2eσs(δ𝑷)2]\displaystyle\sim-|g_{\text{eff}}|^{2}N_{V}^{2}\frac{1}{2E_{0}}\frac{\text{d}^{3}\bm{P}_{1}}{(2\pi)^{3}2E_{1}}\frac{\text{d}^{3}\bm{P}_{2}}{(2\pi)^{3}2E_{2}}(2\pi)^{4}\left(\sqrt{\frac{\sigma_{t}}{\pi}}e^{-\sigma_{t}\left(\delta\omega\right)^{2}}\right)\left[\left(\frac{\sigma_{s}}{\pi}\right)^{3/2}e^{-\sigma_{s}\left(\delta\bm{P}\right)^{2}}\right]
×eΓV(TinT0)ΓV2σt4+12σt(δω)2\displaystyle\quad\times e^{-\Gamma_{V}\left(T_{\text{in}}-T_{0}\right)-\frac{\Gamma_{V}^{2}\sigma_{t}}{4}+\frac{1}{2}\sigma_{t}\left(\delta\omega\right)^{2}}
×2σtπ(ΓV2(δω)2)cos(2ΓVσtδω)2ΓVδωsin(2ΓVσtδω)σt(ΓV2+(δω)2)2|F~(V)|2.\displaystyle\quad\times\sqrt{\frac{2\sigma_{t}}{\pi}}\frac{\left(\Gamma_{V}^{2}-\left(\delta\omega\right)^{2}\right)\cos\!\left(2\Gamma_{V}\sigma_{t}\delta\omega\right)-2\Gamma_{V}\delta\omega\sin\!\left(2\Gamma_{V}\sigma_{t}\delta\omega\right)}{\sigma_{t}\left(\Gamma_{V}^{2}+\left(\delta\omega\right)^{2}\right)^{2}}\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}. (157)

We perform the momentum integrals in the non-relativistic limit:

PVPP¯intf\displaystyle P^{\text{intf}}_{V\to P\overline{P}} NV2eΓV(TinT0)gV2312mV1(2π)24E1E2×mP68(04πV+2dV+)(04πV2dV)\displaystyle\Rightarrow-N_{V}^{2}e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right)}\frac{g_{V}^{2}}{3}\frac{1}{2m_{V}}\frac{1}{(2\pi)^{2}4E_{1}E_{2}}\times\frac{m_{P}^{6}}{8}\left(\int_{0}^{\infty}4\pi V_{+}^{2}\text{d}V_{+}\right)\left(\int_{0}^{\infty}4\pi V_{-}^{2}\text{d}V_{-}\right)
×1π2σt1/2σs3/2(mP2V2)eFintf(V+,V)2σtπ\displaystyle\quad\times\frac{1}{\pi^{2}}\sigma_{t}^{1/2}\sigma_{s}^{3/2}\left(m_{P}^{2}V_{-}^{2}\right)e^{-F_{\text{intf}}(V_{+},V_{-})}\sqrt{\frac{2\sigma_{t}}{\pi}}
×(ΓV2(δω)2)cos(2ΓVσtδω)2ΓVδωsin(2ΓVσtδω)σt(ΓV2+(δω)2)2|F~(V)|2,\displaystyle\quad\times\frac{\left(\Gamma_{V}^{2}-\left(\delta\omega\right)^{2}\right)\cos\!\left(2\Gamma_{V}\sigma_{t}\delta\omega\right)-2\Gamma_{V}\delta\omega\sin\!\left(2\Gamma_{V}\sigma_{t}\delta\omega\right)}{\sigma_{t}\left(\Gamma_{V}^{2}+\left(\delta\omega\right)^{2}\right)^{2}}\left|\widetilde{F}\!\left(V_{-}\right)\right|^{2}, (158)

where

Fintf(V+,V):=\displaystyle F_{\text{intf}}(V_{+},V_{-}):= σs(δ𝑷)2+12σt(δω)2+ΓV2σt4\displaystyle\,\sigma_{s}\left(\delta\bm{P}\right)^{2}+\frac{1}{2}\sigma_{t}\left(\delta\omega\right)^{2}+\frac{\Gamma_{V}^{2}\sigma_{t}}{4}
=\displaystyle= σsmP2V+2\displaystyle\,\sigma_{s}m_{P}^{2}V_{+}^{2}
+2σP(12σsσP)V+2+V2=σt{12[mP(14σsσP)V+2+14mPV2(mV2mP)]2+ΓV24}.\displaystyle+\underbrace{\frac{2\sigma_{P}}{\left(1-\frac{2\sigma_{s}}{\sigma_{P}}\right)V_{+}^{2}+V_{-}^{2}}}_{=\sigma_{t}}\left\{\frac{1}{2}\left[m_{P}\left(\frac{1}{4}-\frac{\sigma_{s}}{\sigma_{P}}\right)V_{+}^{2}+\frac{1}{4}m_{P}V_{-}^{2}-\left(m_{V}-2m_{P}\right)\right]^{2}+{\Gamma_{V}^{2}\over 4}\right\}. (159)

This function is similar to FbulkF_{\text{bulk}}, but a factor of the half comes in front of σt(δω)2\sigma_{t}\left(\delta\omega\right)^{2}. As in the case of the bulk part, we can see

eFintfV+0,eFintfV0.\displaystyle e^{-F_{\text{intf}}}\underbrace{\to}_{V_{+}\to\infty}0,\quad e^{-F_{\text{intf}}}\underbrace{\to}_{V_{-}\to\infty}0. (160)

Stationary points (V+s,Vs)(V_{+}^{s},V_{-}^{s}) are defined by

Fintf(V+s,Vs)V+=0andFintf(V+s,Vs)V=0,\displaystyle\frac{\partial F_{\text{intf}}(V_{+}^{s},V_{-}^{s})}{\partial V_{+}}=0\quad\text{and}\quad\frac{\partial F_{\text{intf}}(V_{+}^{s},V_{-}^{s})}{\partial V_{-}}=0, (161)

and we find one stationary point in the kinetic region (V+0V_{+}\simeq 0 and V>0V_{-}>0),

(V+I,VI)\displaystyle\left(V_{+}^{\text{I}},V_{-}^{\text{I}}\right) :=(0,2[(mV2mP)2+ΓV22]1/4mP).\displaystyle:=\left(0,\,\frac{2\left[\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 2}\right]^{1/4}}{\sqrt{m_{P}}}\right). (162)

Note that

Fintf0=Fintf(V+I,VI)\displaystyle{F_{\text{intf}}^{0}}=F_{\text{intf}}\left(V_{+}^{\text{I}},V_{-}^{\text{I}}\right) :=mPσP2[(mV2mP)+(mV2mP)2+ΓV22],\displaystyle:=\frac{m_{P}\sigma_{P}}{2}\left[-\left(m_{V}-2m_{P}\right)+\sqrt{\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 2}}\right], (163)

which takes a positive value when ΓV\Gamma_{V} is finite, and goes to zero in the (unphysical) limit ΓV0\Gamma_{V}\to 0. Also, we can see

2FintfV+2|(V+I,VI)\displaystyle\left.\frac{\partial^{2}F_{\text{intf}}}{\partial V_{+}^{2}}\right|_{\left(V_{+}^{\text{I}},V_{-}^{\text{I}}\right)} =12mP2σs[3+mV2mP(mV2mP)2+ΓV22]=:(12mP2σs)4Aintf,\displaystyle=\frac{1}{2}m_{P}^{2}\sigma_{s}\left[3+\frac{m_{V}-2m_{P}}{\sqrt{\left(m_{V}-2m_{P}\right)^{2}+{\Gamma_{V}^{2}\over 2}}}\right]=:\left(\frac{1}{2}m_{P}^{2}\sigma_{s}\right)4{A_{\text{intf}}},
2FintfV2|(V+I,VI)\displaystyle\left.\frac{\partial^{2}F_{\text{intf}}}{\partial V_{-}^{2}}\right|_{\left(V_{+}^{\text{I}},V_{-}^{\text{I}}\right)} =12mP2σP,\displaystyle=\frac{1}{2}m_{P}^{2}\sigma_{P},
2FintfV+V|(V+I,VI)\displaystyle\left.\frac{\partial^{2}F_{\text{intf}}}{\partial V_{+}\partial V_{-}}\right|_{\left(V_{+}^{\text{I}},V_{-}^{\text{I}}\right)} =0,\displaystyle=0, (164)

where Aintf1A_{\text{intf}}\to 1 under the (unphysical) limit ΓV0\Gamma_{V}\to 0.

Now, we evaluate the non-relativistic integral in Eq. (158). By use of Eqs. (137) and (138), we reach

PVPP¯intf\displaystyle P^{\text{intf}}_{V\to P\overline{P}} (gV2mP3NV2eΓV(TinT0)12πmVE1E2)mPσP22π(VI)2\displaystyle\sim-\left(\frac{g_{V}^{2}m_{P}^{3}N_{V}^{2}e^{-\Gamma_{V}\left(T_{\text{in}}-{T_{0}}\right)}}{12\pi m_{V}E_{1}E_{2}}\right)\frac{m_{P}\sqrt{\sigma_{P}}}{2\sqrt{2}\sqrt{\pi}}\left(V_{-}^{\text{I}}\right)^{2}
×12[1+erf(mPσPVI2)]eFintf0Aintf3/2|F~(VI)|2\displaystyle\quad\times\frac{1}{2}\left[1+\text{erf}\left(\frac{m_{P}\sqrt{\sigma_{P}}{V_{-}^{\text{I}}}}{2}\right)\right]{\frac{e^{-F^{0}_{\text{intf}}}}{A_{\text{intf}}^{3/2}}}\,\left|\widetilde{F}\!\left(V^{\text{I}}_{-}\right)\right|^{2}
×[(ΓV2(δω~)2)cos(2ΓVσt~δω~)2ΓVδω~sin(2ΓVσt~δω~)σt~(ΓV2+(δω~)2)2],\displaystyle\quad\times\left[\frac{\left(\Gamma_{V}^{2}-\left(\widetilde{\delta\omega}\right)^{2}\right)\cos\!\left(2\Gamma_{V}\widetilde{\sigma_{t}}\widetilde{\delta\omega}\right)-2\Gamma_{V}\widetilde{\delta\omega}\sin\!\left(2\Gamma_{V}\widetilde{\sigma_{t}}\widetilde{\delta\omega}\right)}{\widetilde{\sigma_{t}}\left(\Gamma_{V}^{2}+\left(\widetilde{\delta\omega}\right)^{2}\right)^{2}}\right], (165)

where the factor σP\sqrt{\sigma_{P}} comes in the overall factor (instead of ΓV1\Gamma_{V}^{-1} compared with the bulk result in Eq. (136)) and we defined the parameters

σt~\displaystyle\widetilde{\sigma_{t}} :=σt|V+V+I,VVI=2σP(12σsσP)V+2+V2|V+V+I,VVI=2σP(VI)2,\displaystyle:=\sigma_{t}\Big{|}_{V_{+}\to V_{+}^{\text{I}},\,V_{-}\to V_{-}^{\text{I}}}=\left.\frac{2\sigma_{P}}{\left(1-\frac{2\sigma_{s}}{\sigma_{P}}\right)V_{+}^{2}+V_{-}^{2}}\right|_{V_{+}\to V_{+}^{\text{I}},\,V_{-}\to V_{-}^{\text{I}}}=\frac{2\sigma_{P}}{\left(V_{-}^{\text{I}}\right)^{2}}, (166)
δω~\displaystyle\widetilde{\delta\omega} :=δω|V+V+I,VVI=[mP(14σsσP)V+2+14mPV2(mV2mP)]|V+V+I,VVI\displaystyle:=\delta\omega\Big{|}_{V_{+}\to V_{+}^{\text{I}},\,V_{-}\to V_{-}^{\text{I}}}=\left.\left[m_{P}\left(\frac{1}{4}-\frac{\sigma_{s}}{\sigma_{P}}\right)V_{+}^{2}+\frac{1}{4}m_{P}V_{-}^{2}-\left(m_{V}-2m_{P}\right)\right]\right|_{V_{+}\to V_{+}^{\text{I}},\,V_{-}\to V_{-}^{\text{I}}}
=14mP(VI)2(mV2mP).\displaystyle=\frac{1}{4}m_{P}\left(V_{-}^{\text{I}}\right)^{2}-\left(m_{V}-2m_{P}\right). (167)

Appendix C Plane-wave decay rate

From the effective Hamiltonian (7), under the use of the form factor (105), it is immediate to get the following form in the plane-wave formalism for the resting VV,

ΓVPP¯plane\displaystyle\Gamma^{\text{plane}}_{V\to P\overline{P}} =23(gVP24π)|𝒌P|3mV2|F~(𝒌1𝒌22)|2\displaystyle=\frac{2}{3}\left(\frac{{g_{VP}}^{2}}{4\pi}\right)\frac{|\bm{k}_{P}|^{3}}{m_{V}^{2}}\left|\widetilde{F}\!\left(\frac{\bm{k}_{1}-\bm{k}_{2}}{2}\right)\right|^{2}
=gVP248πmV2(mV24mP2)3/2|1(R0(𝒌1𝒌2)2)2+1|2,\displaystyle=\frac{{g_{VP}}^{2}}{48\pi m_{V}^{2}}\left(m_{V}^{2}-4m_{P}^{2}\right)^{3/2}\left|\frac{1}{\left(\frac{R_{0}\left(\bm{k}_{1}-\bm{k}_{2}\right)}{2}\right)^{2}+1}\right|^{2}, (168)

where gVPg_{VP} represents gV+g_{V+} (for P=P+P=P^{+}) or gV0g_{V0} (for P=P0P=P^{0}) and the magnitude of the final-state momenta in the center-of-mass frame is given as

|𝒌P|=|𝒌1|=|𝒌2|\displaystyle\left|\bm{k}_{P}\right|=\left|\bm{k}_{1}\right|=\left|\bm{k}_{2}\right| =12(mV24mP2)1/2.\displaystyle=\frac{1}{2}\left(m_{V}^{2}-4m_{P}^{2}\right)^{1/2}. (169)

The form factor part does not take the non-relativistic limit as in Eq. (105), and currently, 𝒌1=𝒌2\bm{k}_{1}=-\bm{k}_{2} and thus

|𝒌1𝒌2|=2|𝒌1|=(mV24mP2)1/2.\displaystyle\left|\bm{k}_{1}-\bm{k}_{2}\right|=2\left|\bm{k}_{1}\right|=\left(m_{V}^{2}-4m_{P}^{2}\right)^{1/2}. (170)

Note that under the non-relativistic approximation taken in Eqs (101) and (102), it is easy to obtain the approximated form:

ΓVPP¯plane\displaystyle\Gamma^{\text{plane}}_{V\to P\overline{P}} ΓVPP¯plane,non-rel:=gVP2mP212πmV(mV2mPmP)3/2|1(R0mP(𝑽1𝑽2)2)2+1|2,\displaystyle\Rightarrow\Gamma^{\text{plane,non-rel}}_{V\to P\overline{P}}:=\frac{{g_{VP}}^{2}m_{P}^{2}}{12\pi m_{V}}\left(\frac{m_{V}-2m_{P}}{m_{P}}\right)^{3/2}\left|\frac{1}{\left(\frac{R_{0}m_{P}\left(\bm{V}_{1}-\bm{V}_{2}\right)}{2}\right)^{2}+1}\right|^{2}, (171)
|𝑽1𝑽2|\displaystyle\left|\bm{V}_{1}-\bm{V}_{2}\right| 2(mV2mP)1/2mP.\displaystyle\approx\frac{2\left(m_{V}-2m_{P}\right)^{1/2}}{\sqrt{m_{P}}}. (172)

Appendix D Comparison with plane-wave decay rate

As shown in Sec. 3, the boundary contribution dominates over the bulk and interference ones in all regions of the parameter space R0R_{0} and σP\sigma_{P}. This is because of the fast decay of the vector meson VV due to strong interactions. The fast decay suppresses the contribution from the bulk so that the contribution from the initial time boundary becomes more significant compared to the bulk one.

As mentioned above, the plane-wave decay width ΓV\Gamma_{V} is dependent on other theory parameters such as gVg_{V}, mPm_{P}, and mVm_{V}. Therefore, it is meaningless to take an “on-shell” limit ΓV0\Gamma_{V}\to 0 with other parameters being fixed. Nevertheless, one might pretend that one can take this unphysical limit, and then one extracts the plane-wave decay width (171):

ΓVPP¯plane,non-rel\displaystyle\Gamma_{V\to P\overline{P}}^{\text{plane,non-rel}} =limσP(limΓV0ΓVPVPP¯)=limσP(limΓV0ΓVPVPP¯bulk),\displaystyle=\lim_{\sigma_{P}\to\infty}\left(\lim_{\Gamma_{V}\to 0}\Gamma_{V}P_{V\to P\overline{P}}\right)=\lim_{\sigma_{P}\to\infty}\left(\lim_{\Gamma_{V}\to 0}\Gamma_{V}P_{V\to P\overline{P}}^{\text{bulk}}\right), (173)

where the limit of large wave-packet size σP\sigma_{P}\to\infty is taken after the limit ΓV0\Gamma_{V}\to 0.333333 As said above, σV\sigma_{V} anyway drops out of the result at this order of the saddle-point approximation. The second equality in Eq. (173) is derived as follows: Under the limit ΓV0\Gamma_{V}\to 0, the two ratios ΓVPVPP¯bdry\Gamma_{V}P^{\text{bdry}}_{V\to P\overline{P}} and ΓVPVPP¯intf\Gamma_{V}P^{\text{intf}}_{V\to P\overline{P}} approach zero since PVPP¯bdryP^{\text{bdry}}_{V\to P\overline{P}} and PVPP¯intfP^{\text{intf}}_{V\to P\overline{P}} are not proportional to ΓV1\Gamma_{V}^{-1} as shown in Eqs. (45) and (48). When taking the above limits ΓV0\Gamma_{V}\to 0 and σP\sigma_{P}\to\infty, the following is satisfied:

VB\displaystyle V_{-}^{\text{B}} 2mV2mPmP,\displaystyle\to 2\frac{\sqrt{m_{V}-2m_{P}}}{\sqrt{m_{P}}}, erf(mPσPVB2)\displaystyle\text{erf}\left(\frac{m_{P}\sqrt{\sigma_{P}}{V_{-}^{\text{B}}}}{\sqrt{2}}\right) 1,\displaystyle\to 1,
Fbulk0\displaystyle F_{\text{bulk}}^{0} 0,\displaystyle\to 0, Abulk\displaystyle A_{\text{bulk}} 1,\displaystyle\to 1, (174)

as well as the physical requirement NV1N_{V}\to 1 from ΓV0\Gamma_{V}\to 0.

In the actual setup, the (unphysical) ΓV0\Gamma_{V}\to 0 limit is not good, and we see that the boundary contribution dominates over the bulk one for the parameters corresponding to the real experiments mainly because of the exponential suppression factor eFbulk0e^{-F^{0}_{\text{bulk}}}.

Appendix E A brief comment on the isospin breaking of the ρ\rho system

We will make a brief comment on the isospin breaking of the ρ\rho system. Here, we define the following ratio,

RρWP\displaystyle R_{\rho}^{\text{WP}} :=Pρ+π+π0Pρ0π+π,\displaystyle:={P_{\rho^{+}\to\pi^{+}\pi^{0}}\over P_{\rho^{0}\to\pi^{+}\pi^{-}}}, Rρplane\displaystyle R_{\rho}^{\text{plane}} :=Γρ+π+π0planeΓρ0π+πplane,\displaystyle:=\frac{\Gamma^{\text{plane}}_{\rho^{+}\to\pi^{+}\pi^{0}}}{\Gamma^{\text{plane}}_{\rho^{0}\to\pi^{+}\pi^{-}}}, Rρparton\displaystyle R_{\rho}^{\text{parton}} :=Γρ+π+π0plane|without form factorΓρ0π+πplane|without form factor,\displaystyle:=\frac{\left.\Gamma^{\text{plane}}_{\rho^{+}\to\pi^{+}\pi^{0}}\right|_{\text{without form factor}}}{\left.\Gamma^{\text{plane}}_{\rho^{0}\to\pi^{+}\pi^{-}}\right|_{\text{without form factor}}}, (175)

where we replace mPm_{P} to (mπ++mπ0)/2\left(m_{\pi^{+}}+m_{\pi^{0}}\right)/2 for the calculations of ρ+π+π0\rho^{+}\to\pi^{+}\pi^{0}. In Fig. 16, the RρR_{\rho} in terms of the wave packet and the relativistic plane wave is depicted for R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1} (Left panel) and R0=0.01MeV1R_{0}=0.01\,\text{MeV}^{-1} (Right panel); see Sections 2.3 and 4.3 for the details of the wave-packet decay probabilities and the plane-wave decay rates, respectively.

First, we mention the experimental inputs that we adopt. As official results reported by the PDG [1],

  • Six digits are reported as typical mass scales of the broad-resonant ρ\rho system, where their central values are located from 769.0MeV769.0\,\text{MeV} to 775.26MeV775.26\,\text{MeV}.

  • Also, six digits are shown as typical width scales of the ρ\rho system, where their central values are located from 147.4MeV147.4\,\text{MeV} to 151.5MeV151.5\,\text{MeV}.

  • The difference between mρ0m_{\rho^{0}} and mρ+m_{\rho^{+}} takes mρ0mρ+=0.7±0.8MeVm_{\rho^{0}}-m_{\rho^{+}}=-0.7\pm 0.8\,\text{MeV}.

  • The difference between Γρ0\Gamma_{\rho^{0}} and Γρ+\Gamma_{\rho^{+}} takes Γρ0Γρ+=0.3±1.3MeV\Gamma_{\rho^{0}}-\Gamma_{\rho^{+}}=0.3\pm 1.3\,\text{MeV}.

Since RρR_{\rho} measures isospin-violating effects, and thus it is insensitive to what is a typical mass scale of the system. So, we simply take mρ0=770±1MeVm_{\rho^{0}}=770\pm 1\,\text{MeV} and Γρ0=147.4±0.8MeV\Gamma_{\rho^{0}}=147.4\pm 0.8\,\text{MeV}. Based on the mass scale, mρ+m_{\rho^{+}} is estimated by use of the above data on mρ0m_{\rho^{0}} and mρ0mρ+m_{\rho^{0}}-m_{\rho^{+}} as

mρ+\displaystyle m_{\rho^{+}} =770.7MeV(central),\displaystyle=770.7\,\text{MeV}\,\text{(central)}, mρ+\displaystyle m_{\rho^{+}} =771.5MeV(+1σ),\displaystyle=771.5\,\text{MeV}\,(+1\sigma), mρ+\displaystyle m_{\rho^{+}} =772.3MeV(+2σ),\displaystyle=772.3\,\text{MeV}\,(+2\sigma),

while the central value of Γρ+\Gamma_{\rho^{+}} is similarly estimated as 147.7MeV147.7\,\text{MeV}. As expected and as shown in Fig. 16, RρR_{\rho} can depend on the mass difference sizably.

Refer to caption
Refer to caption
Figure 16: Values of RρR_{\rho} defined in Eq. (175) are depicted for R0=0.0015MeV1R_{0}=0.0015\,\text{MeV}^{-1} (Left panel) and R0=0.01MeV1R_{0}=0.01\,\text{MeV}^{-1} (Right panel). The captions “Wave-Packet (full)”, “Plane-Wave (rel)”, and “Plane-Wave parton-level” correspond to RρWPR_{\rho}^{\text{WP}}, RρplaneR_{\rho}^{\text{plane}}, and RρpartonR_{\rho}^{\text{parton}}, respectively. See the main text of this section for other details.

Next, we discuss the numerical result shown in Fig. 16. As expected, the experimental result is located very near the unity since the ρ\rho vector mesons do not contain heavy quarks. The plane-wave theoretical predictions with and without the form-factor effect show similar results. The wave-packet predictions deviate from the unity several percent upward, showing fewer agreements. Nevertheless, this does not necessarily mean that wave-packet formalism works less effectively for the ρ\rho system than plane-wave formalism because the current wave-packet result is precise only for non-relativistic systems (due to the usage of non-relativistic approximations). Note that the decays ρ+π+π0\rho^{+}\to\pi^{+}\pi^{0} and ρ0π+π\rho^{0}\to\pi^{+}\pi^{-} are fully relativistic due to the large difference between the total masses of the initial state and final state. Several percent of theoretical errors are expected from the fully-relativistic wave-packet prediction, which is beyond the scope of this paper. The typical scale of the form factor under the current scheme R0=0.0015MeVR_{0}=0.0015\,\text{MeV} works well compared with R0=0.01MeVR_{0}=0.01\,\text{MeV}.

Also, we comment on the dependence on R0R_{0} in RρR_{\rho}. As typically observed in the curves of “Plane-Wave (rel)” of Fig. 16, the form-factor part of RρR_{\rho} is not sensitive to R0R_{0} since the total mass differences between the initial and the final states take almost the same values in ρ+π+π0\rho^{+}\to\pi^{+}\pi^{0} and ρ0π+π\rho^{0}\to\pi^{+}\pi^{-} and their final-state phase spaces are wide.

References