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

Hidden-charm pentaquarks with triple strangeness due to the Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} interactions

Fu-Lai Wang1,2 wangfl2016@lzu.edu.cn    Xin-Dian Yang1,2 yangxd20@lzu.edu.cn    Rui Chen4,5 chen rui@pku.edu.cn    Xiang Liu1,2,3111Corresponding author xiangliu@lzu.edu.cn 1School of Physical Science and Technology, Lanzhou University, Lanzhou 730000, China
2Research Center for Hadron and CSR Physics, Lanzhou University and Institute of Modern Physics of CAS, Lanzhou 730000, China
3Lanzhou Center for Theoretical Physics, Key Laboratory of Theoretical Physics of Gansu Province, and Frontiers Science Center for Rare Isotopes, Lanzhou University, Lanzhou 730000, China
4Center of High Energy Physics, Peking University, Beijing 100871, China
5School of Physics and State Key Laboratory of Nuclear Physics and Technology, Peking University, Beijing 100871, China
Abstract

Motivated by the successful interpretation of these observed PcP_{c} and PcsP_{cs} states under the meson-baryon molecular picture, we systematically investigate the possible hidden-charm molecular pentaquark states with triple strangeness which is due to the Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} interactions. We perform a dynamical calculation of the possible hidden-charm molecular pentaquarks with triple strangeness by the one-boson-exchange model, where the SS-DD wave mixing effect and the coupled channel effect are taken into account in our calculation. Our results suggest that the ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} state with JP=3/2J^{P}={3}/{2}^{-} and the ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} state with JP=5/2J^{P}={5}/{2}^{-} can be recommended as the candidates of the hidden-charm molecular pentaquark with triple strangeness. Furthermore, we discuss the two-body hidden-charm strong decay behaviors of these possible hidden-charm molecular pentaquarks with triple strangeness by adopting the quark-interchange model. These predictions are expected to be tested at the LHCb, which can be as a potential research issue with more accumulated experimental data in near future.

I Introduction

As is well known, the study of the matter spectrum is an important way to explore the relevant matter structures and the involved interaction properties. In the hadron physics, since the discovery of the X(3872)X(3872) by the Belle Collaboration Choi:2003ue , a series of exotic states has been observed benefiting from the accumulation of more and more experimental data with high precision, and the exotic hadrons have stimulated extensive studies in the past two decades (see the review articles Chen:2016qju ; Liu:2019zoy ; Olsen:2017bmm ; Guo:2017jvc ; Liu:2013waa ; Hosaka:2016pey ; Brambilla:2019esw for learning the relevant processes). Exploring these exotic hadronic states not only gives new insights for revealing the hadron structures, but also provides useful hints to deepening our understanding of the nonperturbative behavior of the quantum chromodynamics (QCD) in the low energy regions.

In fact, investigating the pentaquark states has been a long history, which can be tracked back to the birth of the quark model GellMann:1964nj ; Zweig:1981pd . Among exotic hadronic states, the hidden-charm molecular pentaquarks have attracted much attention as early as 2010 Li:2014gra ; Karliner:2015ina ; Wu:2010jy ; Wang:2011rga ; Yang:2011wz ; Wu:2012md ; Chen:2015loa and become a hot topic with the discovery of the Pc(4380)P_{c}(4380) and Pc(4450)P_{c}(4450) in the ΛbJ/ψpK\Lambda_{b}\to J/\psi pK process by the LHCb Collaboration Aaij:2015tga . In 2019, there was a new progress about the observation of three narrow structures [Pc(4312)P_{c}(4312), Pc(4440)P_{c}(4440), and Pc(4457)P_{c}(4457)] by revisiting the process ΛbJ/ψpK\Lambda_{b}\to J/\psi pK based on more collected data Aaij:2019vzc , and they are just below the corresponding thresholds of the SS-wave charmed baryon and SS-wave anticharmed meson. This provides strong evidence to support the existence of the hidden-charm meson-baryon molecular states. More recently, the LHCb Collaboration reported a possible hidden-charm pentaquark with strangeness Pcs(4459)P_{cs}(4459) Aaij:2020gdg , and this structure can be assigned as the ΞcD¯\Xi_{c}\bar{D}^{*} molecular state Chen:2016ryt ; Wu:2010vk ; Hofmann:2005sw ; Anisovich:2015zqa ; Wang:2015wsa ; Feijoo:2015kts ; Lu:2016roh ; Xiao:2019gjd ; Shen:2020gpw ; Chen:2015sxa ; Zhang:2020cdi ; Wang:2019nvm ; Chen:2020uif ; Peng:2020hql ; 1830432 ; 1830426 ; Liu:2020hcv ; 1839195 .

Facing the present status of exploring the hidden-charm molecular pentaquarks Chen:2016qju ; Liu:2019zoy ; Olsen:2017bmm ; Guo:2017jvc , we naturally propose a meaningful question: why are we interested in the hidden-charm molecular pentaquark states? The hidden-charm pentaquark states are relatively easy to produce via the bottom baryon weak decays in the experimental facilities Aaij:2019vzc ; Aaij:2020gdg , and the hidden-charm quantum number is a crucial condition for the existence of the hadronic molecules Li:2014gra ; Karliner:2015ina . In addition, it is worth indicating that the heavy hadrons are more likely to generate the bound states due to the relatively small kinetic terms, and the interactions between the charmed baryon and the anticharmed meson may be mediated by exchanging a series of allowed light mesons Chen:2016qju ; Liu:2019zoy . Indeed, these announced hidden-charm pentaquark states have a (cc¯c\bar{c}) pair Chen:2016qju ; Liu:2019zoy ; Olsen:2017bmm ; Guo:2017jvc .

Based on the present research progress on the hidden-charm pentaquarks Chen:2016qju ; Liu:2019zoy ; Olsen:2017bmm ; Guo:2017jvc , the theorists should pay more attention to making the reliable prediction of various types of the hidden-charm molecular pentaquarks and give more abundant suggestions to searching for the hidden-charm molecular pentaquarks accessible at the forthcoming experiment. Generally speaking, there are two important approaches to construct the family of the hidden-charm molecular pentaquark states which is very special in the hadron spectroscopy. Firstly, we propose that there may exist a series of hidden-charm molecular pentaquarks with the different strangeness. Secondly, we also have enough reason to believe that there may exist more hidden-charm molecular pentaquark states with higher mass. In fact, we already studied the Ξc(,)D¯s()\Xi_{c}^{(\prime,*)}\bar{D}_{s}^{(*)} systems with double strangeness Wang:2020bjt and the c()T¯\mathcal{B}_{c}^{(*)}\bar{T} systems with c()=Λc/Σc()\mathcal{B}_{c}^{(*)}=\Lambda_{c}/\Sigma_{c}^{(*)} and T¯=D¯1/D¯2\bar{T}=\bar{D}_{1}/\bar{D}_{2}^{*} Wang:2019nwt , and predicted a series of possible candidates of the hidden-charm molecular pentaquarks. In fact, the triple-strangeness hidden-charm pentaquarks may be regarded as systems that can be used to reveal the binding mechanism and the importance of the scalar-meson exchange as they are not expected to exist in the treatment of Ref. 1839195 . Thus, we investigate the possible hidden-charm molecular pentaquarks with triple strangeness from the Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} interactions, which will be a main task of the present work.

In the present work, we perform a dynamical calculation with the possible hidden-charm molecular pentaquark states with triple strangeness by adopting the one-boson-exchange (OBE) model Chen:2016qju ; Liu:2019zoy , which involves the interactions between an SS-wave charmed baryon Ωc()\Omega_{c}^{(*)} and an SS-wave anticharmed-strange meson D¯s()\bar{D}_{s}^{(*)}. In concrete calculation, the SS-DD wave mixing effect and the coupled channel effect are taken into account. Furthermore, we study the two-body hidden-charm strong decay behaviors of these possible hidden-charm molecular pentaquarks. Here, we adopt the quark-interchange model to estimate the transition amplitudes for the decay widths Barnes:1991em ; Barnes:1999hs ; Barnes:2000hu , which is widely used to give the decay information of the exotic hadronic states during the last few decades Wang:2018pwi ; Wang:2019spc ; Xiao:2019spy ; Wang:2020prk ; Hilbert:2007hc . We hope that the present investigation is a key step to complement the family of the hidden-charm molecular pentaquark state and may provide crucial information of searching for possible hidden-charm molecular pentaquarks with triple strangeness. With higher statistic data accumulation at Run III of the LHC and after High-Luminosity-LHC upgrade Bediaga:2018lhg , it is highly probable that these possible hidden-charm molecular pentaquarks with triple strangeness can be detected at the LHCb Collaboration in the near future, which will be full of opportunities and challenges.

The remainder of this paper is organized as follows. In Sec. II, we introduce how to deduce the effective potentials and present the bound state properties of these investigated Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} systems. We present the quark-interchange model and the two-body strong decay behaviors of these possible molecular pentaquarks in Sec. III. Finally, a short summary follows in Sec. IV.

II The Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} interactions

II.1 OBE effective potentials

In the present work, we study the interactions between an SS-wave charmed baryon Ωc()\Omega_{c}^{(*)} and an SS-wave anticharmed-strange meson D¯s()\bar{D}_{s}^{(*)}. Here, we adopt the OBE model Chen:2016qju ; Liu:2019zoy , and consider the effective potentials from the f0(980)f_{0}(980), η\eta, and ϕ\phi exchanges. In particular, we need to emphasize that the light scalar meson f0(980)f_{0}(980) exchange provides effective interaction for these investigated systems, and we do not consider the σ\sigma and a0(980)a_{0}(980) exchanges in our calculation, since the σ\sigma is usually considered as a meson with only up and down quarks and the a0(980)a_{0}(980) is the light isovector scalar meson. In this subsection, we construct the relevant wave functions and effective Lagrangians, and deduce the OBE effective potentials in the coordinate space for all of the investigated systems.

Firstly, we introduce the flavor and spin-orbital wave functions involved in our calculation. For the Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} systems, the flavor wave function |I,I3|I,I_{3}\rangle is quite simple and reads as |0,0=|Ωc()0Ds()|0,0\rangle=|\Omega_{c}^{(*)0}{D}_{s}^{(*)-}\rangle, where II and I3I_{3} are the isospin and its third component of the discussed system. In addition, the spin-orbital wave functions |LJ2S+1|{}^{2S+1}L_{J}\rangle for these investigated Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} systems are explicitly written as

|ΩcD¯s(LJ2S+1)\displaystyle\left|\Omega_{c}\bar{D}_{s}\left({}^{2S+1}L_{J}\right)\right\rangle =\displaystyle= m,mLC12m,LmLJ,Mχ12m|YL,mL,\displaystyle\sum_{m,m_{L}}C^{J,M}_{\frac{1}{2}m,Lm_{L}}\chi_{\frac{1}{2}m}\left|Y_{L,m_{L}}\right\rangle,
|ΩcD¯s(LJ2S+1)\displaystyle\left|\Omega_{c}^{*}\bar{D}_{s}\left({}^{2S+1}L_{J}\right)\right\rangle =\displaystyle= m,mLC32m,LmLJ,MΦ32m|YL,mL,\displaystyle\sum_{m,m_{L}}C^{J,M}_{\frac{3}{2}m,Lm_{L}}\Phi_{\frac{3}{2}m}\left|Y_{L,m_{L}}\right\rangle,
|ΩcD¯s(LJ2S+1)\displaystyle\left|\Omega_{c}\bar{D}_{s}^{*}\left({}^{2S+1}L_{J}\right)\right\rangle =\displaystyle= m,m,mS,mLC12m,1mS,mSCSmS,LmLJ,Mχ12mϵmμ|YL,mL,\displaystyle\sum_{m,m^{\prime},m_{S},m_{L}}C^{S,m_{S}}_{\frac{1}{2}m,1m^{\prime}}C^{J,M}_{Sm_{S},Lm_{L}}\chi_{\frac{1}{2}m}\epsilon_{m^{\prime}}^{\mu}\left|Y_{L,m_{L}}\right\rangle,
|ΩcD¯s(LJ2S+1)\displaystyle\left|\Omega_{c}^{*}\bar{D}_{s}^{*}\left({}^{2S+1}L_{J}\right)\right\rangle =\displaystyle= m,m,mS,mLC32m,1mS,mSCSmS,LmLJ,MΦ32mϵmμ|YL,mL.\displaystyle\sum_{m,m^{\prime},m_{S},m_{L}}C^{S,m_{S}}_{\frac{3}{2}m,1m^{\prime}}C^{J,M}_{Sm_{S},Lm_{L}}\Phi_{\frac{3}{2}m}\epsilon_{m^{\prime}}^{\mu}\left|Y_{L,m_{L}}\right\rangle.

In the above expressions, SS, LL, and JJ denote the spin, orbit angular momentum, and total angular momentum for the discussed system, respectively. The constant Cab,cde,fC^{e,f}_{ab,cd} is the Clebsch-Gordan coefficient, and |YL,mL|Y_{L,m_{L}}\rangle is the spherical harmonics function. In the static limit, the polarization vector ϵmμ(m=0,±1)\epsilon_{m}^{\mu}\,(m=0,\,\pm 1) with the spin-1 field can be expressed as ϵ0μ=(0,0,0,1)\epsilon_{0}^{\mu}=\left(0,0,0,-1\right) and ϵ±μ=(0,±1,i, 0)/2\epsilon_{\pm}^{\mu}=\left(0,\,\pm 1,\,i,\,0\right)/\sqrt{2}. χ12m\chi_{\frac{1}{2}m} stands for the spin wave function of the charmed baryon with spin S=1/2S={1}/{2}, and the polarization tensor Φ32m\Phi_{\frac{3}{2}m} of the charmed baryon with spin quantum number S=3/2S={3}/{2} can be written in a general form, i.e.,

Φ32m=m1,m2C12m1,1m232,mχ12m1ϵm2μ.\displaystyle\Phi_{\frac{3}{2}m}=\sum_{m_{1},m_{2}}C^{\frac{3}{2},m}_{\frac{1}{2}m_{1},1m_{2}}\chi_{\frac{1}{2}m_{1}}\epsilon_{m_{2}}^{\mu}. (2.2)

In order to write out the relevant scattering amplitudes quantitatively, we usually adopt the effective Lagrangian approach. To be convenient, we construct two types of super-fields 𝒮μ\mathcal{S}_{\mu} and Ha(Q¯)H^{(\overline{Q})}_{a} via the heavy quark limit Wise:1992hn . The superfield 𝒮μ\mathcal{S}_{\mu} is expressed as a combination of the charmed baryons 6\mathcal{B}_{6} with JP=1/2+J^{P}=1/2^{+} and 6\mathcal{B}^{*}_{6} with JP=3/2+J^{P}=3/2^{+} in the 6F6_{F} flavor representation Chen:2017xat , and the superfield Ha(Q¯)H^{(\overline{Q})}_{a} includes the anticharmed-strange vector meson D¯s\bar{D}^{*}_{s} with JP=1J^{P}=1^{-} and the pseudoscalar meson D¯s\bar{D}_{s} with JP=0J^{P}=0^{-} Ding:2008gr . The general expressions of the super-fields 𝒮μ\mathcal{S}_{\mu} and Ha(Q¯)H^{(\overline{Q})}_{a} can be given by

𝒮μ\displaystyle\mathcal{S}_{\mu} =\displaystyle= 13(γμ+vμ)γ56+6μ,\displaystyle-\sqrt{\frac{1}{3}}(\gamma_{\mu}+v_{\mu})\gamma^{5}\mathcal{B}_{6}+\mathcal{B}_{6\mu}^{*},
Ha(Q¯)\displaystyle H^{(\overline{Q})}_{a} =\displaystyle= (D¯a(Q¯)μγμD¯a(Q¯)γ5)1/v2.\displaystyle\left(\bar{D}^{*(\overline{Q})\mu}_{a}\gamma_{\mu}-\bar{D}^{(\overline{Q})}_{a}\gamma_{5}\right)\frac{1-/\!\!\!v}{2}. (2.3)

Here, vμ=(1,𝟎)v_{\mu}=(1,\bm{0}) is the four velocity under the nonrelativistic approximation.

With the above preparation, we construct the relevant effective Lagrangians to describe the interactions among the heavy hadrons 6()/D¯s()\mathcal{B}_{6}^{(*)}/\bar{D}_{s}^{(*)} and the light scalar, pseudoscalar, or vector mesons as Ding:2008gr ; Chen:2017xat

6()\displaystyle\mathcal{L}_{\mathcal{B}^{(*)}_{6}} =\displaystyle= lS𝒮¯μf0𝒮μ32g1εμνλκvκ𝒮¯μ𝒜ν𝒮λ\displaystyle l_{S}\langle\bar{\mathcal{S}}_{\mu}f_{0}\mathcal{S}^{\mu}\rangle-\frac{3}{2}g_{1}\varepsilon^{\mu\nu\lambda\kappa}v_{\kappa}\langle\bar{\mathcal{S}}_{\mu}{\mathcal{A}}_{\nu}\mathcal{S}_{\lambda}\rangle
+iβS𝒮¯μvα(𝒱αρα)𝒮μ+λS𝒮¯μFμν(ρ)𝒮ν,\displaystyle+i\beta_{S}\langle\bar{\mathcal{S}}_{\mu}v_{\alpha}\left(\mathcal{V}^{\alpha}-\rho^{\alpha}\right)\mathcal{S}^{\mu}\rangle+\lambda_{S}\langle\bar{\mathcal{S}}_{\mu}F^{\mu\nu}(\rho)\mathcal{S}_{\nu}\rangle,
H\displaystyle\mathcal{L}_{H} =\displaystyle= gSH¯a(Q¯)f0Ha(Q¯)+igH¯a(Q¯)γμ𝒜abμγ5Hb(Q¯)\displaystyle g_{S}\langle\bar{H}^{(\overline{Q})}_{a}f_{0}H^{(\overline{Q})}_{a}\rangle+ig\langle\bar{H}^{(\overline{Q})}_{a}\gamma_{\mu}{\mathcal{A}}_{ab}^{\mu}\gamma_{5}H^{(\overline{Q})}_{b}\rangle (2.4)
iβH¯a(Q¯)vμ(𝒱μρμ)abHb(Q¯)\displaystyle-i\beta\langle\bar{H}^{(\overline{Q})}_{a}v_{\mu}\left(\mathcal{V}^{\mu}-\rho^{\mu}\right)_{ab}H^{(\overline{Q})}_{b}\rangle
+iλH¯a(Q¯)σμνFμν(ρ)abHb(Q¯),\displaystyle+i\lambda\langle\bar{H}^{(\overline{Q})}_{a}\sigma_{\mu\nu}F^{\mu\nu}(\rho)_{ab}H^{(\overline{Q})}_{b}\rangle,

which satisfy the requirement of the heavy quark symmetry, the chiral symmetry, and the hidden local symmetry Casalbuoni:1992gi ; Casalbuoni:1996pg ; Yan:1992gz ; Harada:2003jx ; Bando:1987br . The axial current 𝒜μ\mathcal{A}_{\mu} and the vector current 𝒱μ{\cal V}_{\mu} can be defined as 𝒜μ=(ξμξξμξ)/2{\mathcal{A}}_{\mu}=\left(\xi^{\dagger}\partial_{\mu}\xi-\xi\partial_{\mu}\xi^{\dagger}\right)/2 and 𝒱μ=(ξμξ+ξμξ)/2{\mathcal{V}}_{\mu}=\left(\xi^{\dagger}\partial_{\mu}\xi+\xi\partial_{\mu}\xi^{\dagger}\right)/2, respectively. Here, the pseudo-Goldstone field can be written as ξ=exp(i/fπ)\xi=\exp(i\mathbb{P}/f_{\pi}), where fπf_{\pi} is the pion decay constant. In the above formulas, the vector meson field ρμ\rho_{\mu} and its strength tensor Fμν(ρ)F_{\mu\nu}(\rho) are ρμ=igV𝕍μ/2\rho_{\mu}=i{g_{V}}\mathbb{V}_{\mu}/{\sqrt{2}} and Fμν(ρ)=μρννρμ+[ρμ,ρν]F_{\mu\nu}(\rho)=\partial_{\mu}\rho_{\nu}-\partial_{\nu}\rho_{\mu}+[\rho_{\mu},\rho_{\nu}], respectively. Here, 6()\mathcal{B}_{6}^{(*)}, 𝕍μ\mathbb{V}_{\mu}, and {\mathbb{P}} are the matrices of the charmed baryon in the 6F6_{F} flavor representation, light vector meson, and light pseudoscalar meson, which can be written as

6()=(Σc()++Σc()+2Ξc(,)+2Σc()+2Σc()0Ξc(,)02Ξc(,)+2Ξc(,)02Ωc()0),𝕍μ=(ρ02+ω2ρ+K+ρρ02+ω2K0KK¯0ϕ)μ,=(π02+η6π+K+ππ02+η6K0KK¯023η),\displaystyle\left.\begin{array}[]{c}\mathcal{B}_{6}^{(*)}=\left(\begin{array}[]{ccc}\Sigma_{c}^{{(*)}++}&\frac{\Sigma_{c}^{{(*)}+}}{\sqrt{2}}&\frac{\Xi_{c}^{(^{\prime},*)+}}{\sqrt{2}}\\ \frac{\Sigma_{c}^{{(*)}+}}{\sqrt{2}}&\Sigma_{c}^{{(*)}0}&\frac{\Xi_{c}^{(^{\prime},*)0}}{\sqrt{2}}\\ \frac{\Xi_{c}^{(^{\prime},*)+}}{\sqrt{2}}&\frac{\Xi_{c}^{(^{\prime},*)0}}{\sqrt{2}}&\Omega_{c}^{(*)0}\end{array}\right),\\ {\mathbb{V}}_{\mu}={\left(\begin{array}[]{ccc}\frac{\rho^{0}}{\sqrt{2}}+\frac{\omega}{\sqrt{2}}&\rho^{+}&K^{*+}\\ \rho^{-}&-\frac{\rho^{0}}{\sqrt{2}}+\frac{\omega}{\sqrt{2}}&K^{*0}\\ K^{*-}&\bar{K}^{*0}&\phi\end{array}\right)}_{\mu},\\ {\mathbb{P}}={\left(\begin{array}[]{ccc}\frac{\pi^{0}}{\sqrt{2}}+\frac{\eta}{\sqrt{6}}&\pi^{+}&K^{+}\\ \pi^{-}&-\frac{\pi^{0}}{\sqrt{2}}+\frac{\eta}{\sqrt{6}}&K^{0}\\ K^{-}&\bar{K}^{0}&-\sqrt{\frac{2}{3}}\eta\end{array}\right)},\end{array}\right. (2.17)

respectively. By expanding the compact effective Lagrangians to the leading order of the pseudo-Goldstone field ξ\xi, we can further obtain the concrete effective Lagrangians. The effective Lagrangians for 6()\mathcal{B}_{6}^{(*)} and the light mesons are expressed as

6()6()f0\displaystyle\mathcal{L}_{\mathcal{B}_{6}^{(*)}\mathcal{B}_{6}^{(*)}f_{0}} =\displaystyle= lS¯6f06+lS¯6μf06μ\displaystyle-l_{S}\langle\bar{\mathcal{B}}_{6}f_{0}\mathcal{B}_{6}\rangle+l_{S}\langle\bar{\mathcal{B}}_{6\mu}^{*}f_{0}\mathcal{B}_{6}^{*\mu}\rangle (2.18)
lS3¯6μf0(γμ+vμ)γ56+h.c.,\displaystyle-\frac{l_{S}}{\sqrt{3}}\langle\bar{\mathcal{B}}_{6\mu}^{*}f_{0}\left(\gamma^{\mu}+v^{\mu}\right)\gamma^{5}\mathcal{B}_{6}\rangle+h.c.,
6()6()\displaystyle\mathcal{L}_{\mathcal{B}_{6}^{(*)}\mathcal{B}_{6}^{(*)}\mathbb{P}} =\displaystyle= ig12fπεμνλκvκ¯6γμγλν6\displaystyle i\frac{g_{1}}{2f_{\pi}}\varepsilon^{\mu\nu\lambda\kappa}v_{\kappa}\langle\bar{\mathcal{B}}_{6}\gamma_{\mu}\gamma_{\lambda}\partial_{\nu}\mathbb{P}\mathcal{B}_{6}\rangle (2.19)
i3g12fπεμνλκvκ¯6μν6λ\displaystyle-i\frac{3g_{1}}{2f_{\pi}}\varepsilon^{\mu\nu\lambda\kappa}v_{\kappa}\langle\bar{\mathcal{B}}_{6\mu}^{*}\partial_{\nu}\mathbb{P}\mathcal{B}_{6\lambda}^{*}\rangle
+i3g12fπvκεμνλκ¯6μνγλγ56+h.c.,\displaystyle+i\frac{\sqrt{3}g_{1}}{2f_{\pi}}v_{\kappa}\varepsilon^{\mu\nu\lambda\kappa}\langle\bar{\mathcal{B}}_{6\mu}^{*}\partial_{\nu}\mathbb{P}{\gamma_{\lambda}\gamma^{5}}\mathcal{B}_{6}\rangle+h.c.,
6()6()𝕍\displaystyle\mathcal{L}_{\mathcal{B}_{6}^{(*)}\mathcal{B}_{6}^{(*)}\mathbb{V}} =\displaystyle= βSgV2¯6v𝕍6\displaystyle-\frac{\beta_{S}g_{V}}{\sqrt{2}}\langle\bar{\mathcal{B}}_{6}v\cdot\mathbb{V}\mathcal{B}_{6}\rangle (2.20)
iλSgV32¯6γμγν(μ𝕍νν𝕍μ)6\displaystyle-i\frac{\lambda_{S}g_{V}}{3\sqrt{2}}\langle\bar{\mathcal{B}}_{6}\gamma_{\mu}\gamma_{\nu}\left(\partial^{\mu}\mathbb{V}^{\nu}-\partial^{\nu}\mathbb{V}^{\mu}\right)\mathcal{B}_{6}\rangle
βSgV6¯6μv𝕍(γμ+vμ)γ56\displaystyle-\frac{\beta_{S}g_{V}}{\sqrt{6}}\langle\bar{\mathcal{B}}_{6\mu}^{*}v\cdot\mathbb{V}\left(\gamma^{\mu}+v^{\mu}\right)\gamma^{5}\mathcal{B}_{6}\rangle
iλSgV6¯6μ(μ𝕍νν𝕍μ)(γν+vν)γ56\displaystyle-i\frac{\lambda_{S}g_{V}}{\sqrt{6}}\langle\bar{\mathcal{B}}_{6\mu}^{*}\left(\partial^{\mu}\mathbb{V}^{\nu}-\partial^{\nu}\mathbb{V}^{\mu}\right)\left(\gamma_{\nu}+v_{\nu}\right)\gamma^{5}\mathcal{B}_{6}\rangle
+βSgV2¯6μvV6μ\displaystyle+\frac{\beta_{S}g_{V}}{\sqrt{2}}\langle\bar{\mathcal{B}}_{6\mu}^{*}v\cdot{V}\mathcal{B}_{6}^{*\mu}\rangle
+iλSgV2¯6μ(μ𝕍νν𝕍μ)6ν+h.c.,\displaystyle+i\frac{\lambda_{S}g_{V}}{\sqrt{2}}\langle\bar{\mathcal{B}}_{6\mu}^{*}\left(\partial^{\mu}\mathbb{V}^{\nu}-\partial^{\nu}\mathbb{V}^{\mu}\right)\mathcal{B}_{6\nu}^{*}\rangle+h.c.,

and the effective Lagrangians to describe the SS-wave anticharmed-strange mesons D¯s()\bar{D}_{s}^{(*)} and the light scalar, pseudoscalar, or vector mesons are

D¯()D¯()f0\displaystyle\mathcal{L}_{{\bar{D}}^{(*)}{\bar{D}}^{(*)}f_{0}} =\displaystyle= 2gSD¯aD¯af0+2gSD¯aμD¯aμf0,\displaystyle-2g_{S}{\bar{D}}_{a}{\bar{D}}_{a}^{{\dagger}}f_{0}+2g_{S}{\bar{D}}_{a\mu}^{*}{\bar{D}}_{a}^{*\mu{\dagger}}f_{0}, (2.21)
D¯()D¯()\displaystyle\mathcal{L}_{{\bar{D}}^{(*)}{\bar{D}}^{(*)}\mathbb{P}} =\displaystyle= 2igfπvαεαμνλD¯aμD¯bλνab\displaystyle\frac{2ig}{f_{\pi}}v^{\alpha}\varepsilon_{\alpha\mu\nu\lambda}{\bar{D}}_{a}^{*\mu{\dagger}}{\bar{D}}_{b}^{*\lambda}\partial^{\nu}{\mathbb{P}}_{ab} (2.22)
+2gfπ(D¯aμD¯b+D¯aD¯bμ)μab,\displaystyle+\frac{2g}{f_{\pi}}\left({\bar{D}}_{a}^{*\mu{\dagger}}{\bar{D}}_{b}+{\bar{D}}_{a}^{{\dagger}}{\bar{D}}_{b}^{*\mu}\right)\partial_{\mu}{\mathbb{P}}_{ab},
D¯()D¯()𝕍\displaystyle\mathcal{L}_{{\bar{D}}^{(*)}{\bar{D}}^{(*)}\mathbb{V}} =\displaystyle= 2βgVD¯aD¯bv𝕍ab2βgVD¯aμD¯bμv𝕍ab\displaystyle\sqrt{2}\beta g_{V}{\bar{D}}_{a}{\bar{D}}_{b}^{{\dagger}}v\cdot\mathbb{V}_{ab}-\sqrt{2}\beta g_{V}{\bar{D}}_{a\mu}^{*}{\bar{D}}_{b}^{*\mu{\dagger}}v\cdot\mathbb{V}_{ab}
22iλgVD¯aμD¯bν(μ𝕍νν𝕍μ)ab\displaystyle-2\sqrt{2}i\lambda g_{V}{\bar{D}}_{a}^{*\mu{\dagger}}{\bar{D}}_{b}^{*\nu}\left(\partial_{\mu}\mathbb{V}_{\nu}-\partial_{\nu}\mathbb{V}_{\mu}\right)_{ab}
22λgVvλελμαβ(D¯aμD¯b+D¯aD¯bμ)α𝕍abβ.\displaystyle-2\sqrt{2}\lambda g_{V}v^{\lambda}\varepsilon_{\lambda\mu\alpha\beta}\left({\bar{D}}_{a}^{*\mu{\dagger}}{\bar{D}}_{b}+{\bar{D}}_{a}^{{\dagger}}{\bar{D}}_{b}^{*\mu}\right)\partial^{\alpha}\mathbb{V}^{\beta}_{ab}.

In the above effective Lagrangians, the coupling constants can be either extracted from the experimental data or calculated by the theoretical models, and the signs of these coupling constants are fixed via the quark model Riska:2000gd . The values of these coupling constants are lS=6.20l_{S}=6.20, gS=0.76g_{S}=0.76,222In this work, we consider the contribution from light scalar meson f0(980)f_{0}(980) exchange. Here, the corresponding coupling constant involved in effective Lagrangians [Eq. (2.18) and Eq. (2.21)] is approximately taken as the same as that for the case of light scalar σ\sigma. g1=0.94g_{1}=0.94, g=0.59g=0.59, fπ=132MeVf_{\pi}=132~{}\rm{MeV}, βSgV=10.14\beta_{S}g_{V}=10.14, βgV=5.25\beta g_{V}=-5.25, λSgV=19.2GeV1\lambda_{S}g_{V}=19.2~{}\rm{GeV}^{-1}, and λgV=3.27GeV1\lambda g_{V}=-3.27~{}\rm{GeV}^{-1} Chen:2019asm , which are widely used to discuss the hadronic molecular states Wang:2020bjt ; Chen:2017xat ; Wang:2019nwt ; Chen:2019asm ; He:2015cea ; He:2019ify ; Chen:2018pzd . In particular, we need to emphasize that these input coupling constants can well reproduce the masses of the Pc(4312)P_{c}(4312), Pc(4440)P_{c}(4440), and Pc(4457)P_{c}(4457) Aaij:2019vzc under the meson-baryon molecular picture when adopting the OBE model Chen:2019asm ; He:2019ify .

We follow the standard strategy to deduce the effective potentials in the coordinate space in Refs. Wang:2020dya ; Wang:2019nwt ; Wang:2019aoc , which is a lengthy and tedious calculation. In Fig. 1, we present the relevant Feynman diagram for the Ωc()D¯s()Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)}\to\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} scattering processes.

Refer to caption
Figure 1: Relevant Feynman diagram for the Ωc()D¯s()Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)}\to\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} scattering processes.

At the hadronic level, we firstly write out the scattering amplitude (h1h2h3h4)\mathcal{M}(h_{1}h_{2}\to h_{3}h_{4}) of the scattering process h1h2h3h4h_{1}h_{2}\to h_{3}h_{4} by considering the effective Lagrangian approach. And then, the effective potential in momentum space 𝒱h1h2h3h4(𝒒)\mathcal{V}^{h_{1}h_{2}\to h_{3}h_{4}}(\bm{q}) can be related to the scattering amplitude (h1h2h3h4)\mathcal{M}(h_{1}h_{2}\to h_{3}h_{4}) with the help of the Breit approximation Breit:1929zz ; Breit:1930zza and the nonrelativistic normalization, i.e.,

𝒱Eh1h2h3h4(𝒒)\displaystyle\mathcal{V}_{E}^{h_{1}h_{2}\to h_{3}h_{4}}(\bm{q}) =\displaystyle= (h1h2h3h4)i2mif2mf,\displaystyle-\frac{\mathcal{M}(h_{1}h_{2}\to h_{3}h_{4})}{\sqrt{\prod_{i}2m_{i}\prod_{f}2m_{f}}}, (2.24)

where mim_{i} and mfm_{f} are the masses of the initial states (h1,h2)(h_{1},\,h_{2}) and final states (h3,h4)(h_{3},\,h_{4}), respectively. By performing the Fourier transformation, the effective potential in the coordinate space 𝒱h1h2h3h4(𝒓)\mathcal{V}^{h_{1}h_{2}\to h_{3}h_{4}}(\bm{r}) can be deduced

𝒱Eh1h2h3h4(𝒓)=d3𝒒(2π)3ei𝒒𝒓𝒱Eh1h2h3h4(𝒒)2(q2,mE2).\displaystyle\mathcal{V}^{h_{1}h_{2}\to h_{3}h_{4}}_{E}(\bm{r})=\int\frac{d^{3}\bm{q}}{(2\pi)^{3}}e^{i\bm{q}\cdot\bm{r}}\mathcal{V}_{E}^{h_{1}h_{2}\to h_{3}h_{4}}(\bm{q})\mathcal{F}^{2}(q^{2},m_{E}^{2}).

In order to reflect the finite size effect of the discussed hadrons and compensate the off-shell effect of the exchanged light mesons Wang:2020dya , we need to introduce the monopole form factor (q2,mE2)=(Λ2mE2)/(Λ2q2)\mathcal{F}(q^{2},m_{E}^{2})=(\Lambda^{2}-m_{E}^{2})/(\Lambda^{2}-q^{2}) in the interaction vertex Tornqvist:1993ng ; Tornqvist:1993vu . Here, Λ\Lambda, mEm_{E}, and qq are the cutoff parameter, the mass, and the four momentum of the exchanged light meson, respectively.

In addition, we also need a series of normalization relations for the heavy hadrons DsD_{s}, DsD_{s}^{*}, Ωc\Omega_{c}, and Ωc\Omega_{c}^{*}, i.e.,

0|Ds|cs¯(0)\displaystyle\langle 0|D_{s}|c\bar{s}\left(0^{-}\right)\rangle =\displaystyle= MDs,\displaystyle\sqrt{M_{D_{s}}}, (2.26)
0|Dsμ|cs¯(1)\displaystyle\langle 0|D_{s}^{*\mu}|c\bar{s}\left(1^{-}\right)\rangle =\displaystyle= MDsϵμ,\displaystyle\sqrt{M_{D_{s}^{*}}}\epsilon^{\mu}, (2.27)
0|Ωc|css(1/2+)\displaystyle\langle 0|\Omega_{c}|css\left({1}/{2}^{+}\right)\rangle =\displaystyle= 2MΩc(χ12m,σ𝐩2MΩcχ12m)T,\displaystyle\sqrt{2M_{\Omega_{c}}}{\left(\chi_{\frac{1}{2}m},\frac{\bf{\sigma}\cdot\bf{p}}{2M_{\Omega_{c}}}\chi_{\frac{1}{2}m}\right)^{T}}, (2.28)
0|Ωcμ|css(3/2+)\displaystyle\langle 0|\Omega_{c}^{*\mu}|css\left({3}/{2}^{+}\right)\rangle =\displaystyle= m1,m2C1/2,m1;1,m23/2,m1+m22MΩc\displaystyle\sum_{m_{1},m_{2}}C_{1/2,m_{1};1,m_{2}}^{3/2,m_{1}+m_{2}}\sqrt{2M_{\Omega_{c}^{*}}} (2.29)
×(χ12m1,σ𝐩2MΩcχ12m1)Tϵm2μ.\displaystyle\times\left(\chi_{\frac{1}{2}m_{1}},\frac{\bf{\sigma}\cdot\bf{p}}{2M_{\Omega_{c}^{*}}}\chi_{\frac{1}{2}m_{1}}\right)^{T}\epsilon^{\mu}_{m_{2}}.

With the above preparation, we can deduce the OBE effective potentials in the coordinate space for all of the investigated processes, which are collected in the A.

II.2 Finding bound state solutions for discussed systems

Now, we attempt to find the loosely bound state solutions of these discussed Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} systems by solving the coupled channel Schrödinger equation, i.e.,

12μ(2(+1)r2)ψ(r)+V(r)ψ(r)=Eψ(r)\displaystyle-\frac{1}{2\mu}\left(\nabla^{2}-\frac{\ell(\ell+1)}{r^{2}}\right)\psi(r)+V(r)\psi(r)=E\psi(r) (2.30)

with 2=1r2rr2r\nabla^{2}=\frac{1}{r^{2}}\frac{\partial}{\partial r}r^{2}\frac{\partial}{\partial r}, where μ=m1m2m1+m2\mu=\frac{m_{1}m_{2}}{m_{1}+m_{2}} is the reduced mass for the discussed system. The bound state solutions include the binding energy EE, the root-mean-square radius rRMSr_{\rm RMS}, and the probability of the individual channel PiP_{i}, which provides us with valuable information to analyze whether the loosely bound state exists. In this work, we are interested in the SS-wave Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} systems since there exists the repulsive centrifugal potential for the higher partial wave states 1\ell\geqslant 1.

In our calculation, the masses of these involved hadrons are mf0=990.00m_{f_{0}}=990.00 MeV, mη=547.85m_{\eta}=547.85 MeV, mϕ=1019.46m_{\phi}=1019.46 MeV, mDs=1968.34m_{D_{s}}=1968.34 MeV, mDs=2112.20m_{D_{s}^{*}}=2112.20 MeV, mΩc=2695.20m_{\Omega_{c}}=2695.20 MeV, and mΩc=2765.90m_{\Omega_{c}^{*}}=2765.90 MeV, which are taken from the Particle Data Group (PDG) Zyla:2020zbs . As the remaining phenomenological parameter, we take the cutoff value from 1.00 to 4.00 GeV. Usually, a loosely bound state with the cutoff parameter closed to 1.00 GeV can be suggested as the possible hadronic molecular candidate according to the experience of the deuteron Tornqvist:1993ng ; Tornqvist:1993vu ; Wang:2019nwt ; Chen:2017jjn . For an ideal hadronic molecular candidate, the reasonable binding energy should be at most tens of MeV, and the typical size should be larger than the size of all the included component hadrons Chen:2017xat .

In addition, the SS-DD wave mixing effect is considered in this work, which plays an important role to modify the bound state properties of the deuteron Wang:2019nwt . The relevant channels |LJ2S+1|{}^{2S+1}L_{J}\rangle are summarized in Table 1.

Table 1: The relevant channels |LJ2S+1|{}^{2S+1}L_{J}\rangle involved in our calculation. Here, “...” means that the SS-wave component for the corresponding channel does not exist.
JPJ^{P} ΩcD¯s\Omega_{c}\bar{D}_{s} ΩcD¯s\Omega_{c}^{*}\bar{D}_{s} ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} ΩcD¯s\Omega_{c}^{*}\bar{D}^{*}_{s}
1/2{1}/{2}^{-} |𝕊1/22|{}^{2}\mathbb{S}_{1/2}\rangle ... |𝕊1/22/|𝔻1/24|{}^{2}\mathbb{S}_{1/2}\rangle/|{}^{4}\mathbb{D}_{1/2}\rangle |𝕊1/22/|𝔻1/24,6|{}^{2}\mathbb{S}_{1/2}\rangle/|{}^{4,6}\mathbb{D}_{1/2}\rangle
3/2{3}/{2}^{-} ... |𝕊3/24/|𝔻3/24|{}^{4}\mathbb{S}_{3/2}\rangle/|{}^{4}\mathbb{D}_{3/2}\rangle |𝕊3/24/|𝔻3/22,4|{}^{4}\mathbb{S}_{3/2}\rangle/|{}^{2,4}\mathbb{D}_{3/2}\rangle |𝕊3/24/|𝔻3/22,4,6|{}^{4}\mathbb{S}_{3/2}\rangle/|{}^{2,4,6}\mathbb{D}_{3/2}\rangle
5/2{5}/{2}^{-} ... ... ... |𝕊5/26/|𝔻5/22,4,6|{}^{6}\mathbb{S}_{5/2}\rangle/|{}^{2,4,6}\mathbb{D}_{5/2}\rangle

Before performing numerical calculation, we analyze the OBE effective potentials for these discussed Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} systems as below:

  • For the ΩcD¯s\Omega_{c}\bar{D}_{s} and ΩcD¯s\Omega_{c}^{*}\bar{D}_{s} systems, only the f0f_{0} and ϕ\phi exchange interactions are allowed. Meanwhile, the tensor force from the SS-DD wave mixing effect disappears in the effective potentials, and thus the contribution of the SS-DD wave mixing effect does not affect the bound state properties of the ΩcD¯s\Omega_{c}\bar{D}_{s} and ΩcD¯s\Omega_{c}^{*}\bar{D}_{s} systems.

  • For the ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} and ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} systems, in addition to the f0f_{0} and ϕ\phi exchange interactions, the η\eta exchange interaction and the SS-DD wave mixing effect need to be taken into account.

II.2.1 The ΩcD¯s\Omega_{c}\bar{D}_{s} and ΩcD¯s\Omega_{c}^{*}\bar{D}_{s} systems

For the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s} state with JP=1/2J^{P}={1}/{2}^{-} and the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s} state with JP=3/2J^{P}={3}/{2}^{-}, we fail to find their bound state solutions by varying the cutoff parameter in the range of 1.001.00-4.00GeV4.00~{}{\rm GeV} with the single channel analysis. Nevertheless, we can further take into account the coupled channel effect. In the coupled channel analysis, the binding energy of the bound state is determined by the lowest mass threshold among various investigated channels Chen:2017xat .

For the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s} state with JP=1/2J^{P}={1}/{2}^{-}, we consider the coupled channel effect from the ΩcD¯s\Omega_{c}\bar{D}_{s}, ΩcD¯s\Omega_{c}\bar{D}_{s}^{*}, and ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} channels. In Table 2, we present the obtained bound state solutions by performing the coupled channel analysis. When we set the cutoff parameter Λ\Lambda around 2.92 GeV, the loosely bound state solutions can be obtained, and the ΩcD¯s\Omega_{c}\bar{D}_{s} channel is dominant with almost 90% probabilities. Since the cutoff parameter Λ\Lambda is obviously different from 1.00 GeV Tornqvist:1993ng ; Tornqvist:1993vu , the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s} state with JP=1/2J^{P}={1}/{2}^{-} is not priority for recommending the hadronic molecular candidate.

Table 2: Bound state solutions of the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s} state with JP=1/2J^{P}={1}/{2}^{-} by performing coupled channel analysis. Here, the cutoff parameter Λ\Lambda, binding energy EE, and root-mean-square radius rRMSr_{RMS} are in units of GeV\rm{GeV}, MeV\rm{MeV}, and fm\rm{fm}, respectively.
Λ\Lambda EE rRMSr_{\rm RMS} P(ΩcD¯s/ΩcD¯s/ΩcD¯s\Omega_{c}\bar{D}_{s}/\Omega_{c}\bar{D}_{s}^{*}/\Omega_{c}^{*}\bar{D}_{s}^{*})
2.92 3.71-3.71 1.26 92.92/4.77/2.31
2.93 12.65-12.65 0.64 90.11/6.66/3.23

In Table 3, we list the bound state solutions of the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s} state with JP=3/2J^{P}={3}/{2}^{-} with the coupled channel analysis. Our numerical results show that the bound state solutions can be obtained by choosing the cutoff parameter Λ\Lambda around 1.78 GeV or even larger, and the ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} system is the dominant channel with the probabilities over 80%. However, we find the size (rRMS0.33fmr_{\rm RMS}\sim 0.33~{}{\rm{fm}}) of this bound state is too small,333 We notice that the obtained values of rRMSr_{RMS} are too small, which is due to the fact that this sysmtem is dominated by the ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} channel as shown in the last column of Table 3. which is not consistent with a loosely molecular state picture Chen:2017xat . Thus, we tentatively exclude the possibility of the existence of the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s} molecular state with JP=3/2J^{P}={3}/{2}^{-}.

Table 3: Bound state solutions of the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s} state with JP=3/2J^{P}={3}/{2}^{-} when the coupled channel effect is introduced. The units are the same as Table 2.
Λ\Lambda EE rRMSr_{\rm RMS} P(ΩcD¯s/ΩcD¯s/ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}/\Omega_{c}\bar{D}_{s}^{*}/\Omega_{c}^{*}\bar{D}_{s}^{*})
1.78 6.15-6.15 0.33 0.01/86.64/13.36
1.79 17.41-17.41 0.32 0.01/86.37/13.63

II.2.2 The ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} and ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} systems

For the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} system, the relevant numerical results are collected in Table 4. For JP=1/2J^{P}={1}/{2}^{-}, there do not exist bound states until we increase the cutoff parameter to be around 4.00 GeV, even if we consider the coupled channel effect. Thus, we conclude that our quantitative analysis does not support the existence of the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} molecular state with JP=1/2J^{P}={1}/{2}^{-}.

Table 4: Bound state solutions of the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} system. The units are the same as Table 2.
Effect Single channel SS-DD wave mixing effect Coupled channel
JPJ^{P} Λ\Lambda EE rRMSr_{\rm RMS} Λ\Lambda EE rRMSr_{\rm RMS} P(𝕊324/𝔻322/𝔻324){}^{4}\mathbb{S}_{\frac{3}{2}}/{}^{2}\mathbb{D}_{\frac{3}{2}}/{}^{4}\mathbb{D}_{\frac{3}{2}}) Λ\Lambda EE rRMSr_{\rm RMS} P(ΩcD¯s/ΩcD¯s\Omega_{c}\bar{D}_{s}^{*}/\Omega_{c}^{*}\bar{D}_{s}^{*})
3/2{3}/{2}^{-} 1.96 0.19-0.19 4.76 1.96 0.33-0.33 4.14 99.94/0.01/0.05 1.67 1.36-1.36 2.27 95.84/4.16
1.98 5.36-5.36 1.09 1.98 5.71-5.71 1.06 99.92/0.02/0.06 1.69 8.38-8.38 0.90 92.17/7.83
1.99 9.44-9.44 0.82 1.99 9.84-9.84 0.81 99.93/0.02/0.05 1.70 13.35-13.35 0.72 91.00/9.00

For the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} state with JP=3/2J^{P}={3}/{2}^{-}, we notice that the effective potentials from the f0f_{0}, η\eta, and ϕ\phi exchanges provide the attractive forces, and there exist the bound state solutions with the cutoff parameter around 1.96 GeV by performing the single channel analysis. More interestingly, the bound state properties will change accordingly after including the coupled channels ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} and ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*}, where we can obtain the loosely bound state solutions when the cutoff parameter Λ\Lambda around 1.67 GeV. Moreover, this bound state is mainly composed of the ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} channel with the probabilities over 90%. Based on our numerical results, the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} state with JP=3/2J^{P}={3}/{2}^{-} can be recommended as a good candidate of the hidden-charm molecular pentaquark with triple strangeness.

Comparing the numerical results, it is obvious that the DD-wave probabilities are less than 1% and the SS-DD mixing effect can be ignored in forming the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} bound states, but the coupled channel effect is obvious in generating the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} bound states, especially for the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} molecular candidate with JP=3/2J^{P}={3}/{2}^{-}.

For the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} system, the bound state properties are collected in Table 5. Here, we still scan the Λ\Lambda parameter range from 1.00 GeV to 4.00 GeV. For JP=1/2J^{P}=1/2^{-}, the binding energy is a few MeV and the root-mean-square radii are around 1.00 fm with the cutoff parameter Λ\Lambda larger than 3.59 GeV when only considering the SS-wave channel, and we can also obtain the bound state solutions when the cutoff value Λ\Lambda is lowered down 3.51 GeV after adding the contribution of the DD-wave channels. Because the obtained cutoff parameter Λ\Lambda is far away from 1.00 GeV Tornqvist:1993ng ; Tornqvist:1993vu , our numerical results disfavor the existence of the molecular candidate for the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} state with JP=1/2J^{P}=1/2^{-}. For JP=3/2J^{P}=3/2^{-}, there do not exist the bound state solutions when the cutoff parameter varies from 1.00 GeV to 4.00 GeV. This situation does not change when the SS-DD wave mixing effect is considered. Thus, we can exclude the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} state with JP=3/2J^{P}=3/2^{-} as the hadronic molecular candidate. For JP=5/2J^{P}=5/2^{-}, we notice that the total effective potentials due to the f0f_{0}, η\eta, and ϕ\phi exchanges are attractive. We can obtain the loosely bound state solutions by taking the cutoff value around 1.64 GeV when only considering the contribution of the SS-wave channel, and the bound state solutions also can be found with the cutoff parameter around 1.64 GeV after considering the SS-DD wave mixing effect. As a result, the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} state with JP=5/2J^{P}=5/2^{-} can be regarded as the hidden-charm molecular pentaquark candidate with triple strangeness.

Table 5: Bound state solutions of the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} system. The units are the same as Table 2.
Effect Single channel SS-DD wave mixing effect
JPJ^{P} Λ\Lambda EE rRMSr_{\rm RMS} Λ\Lambda EE rRMSr_{\rm RMS} P(𝕊122/𝔻124/𝔻126){}^{2}\mathbb{S}_{\frac{1}{2}}/{}^{4}\mathbb{D}_{\frac{1}{2}}/{}^{6}\mathbb{D}_{\frac{1}{2}})
1/2{1}/{2}^{-} 3.593.59 0.27-0.27 4.964.96 3.51 0.29-0.29 4.89 99.97/0.02/0.01
3.803.80 1.18-1.18 2.962.96 3.76 1.74-1.74 2.52 99.92/0.05/0.03
4.004.00 2.63-2.63 2.112.11 4.00 4.35-4.35 1.72 99.87/0.08/0.05
JPJ^{P} Λ\Lambda EE rRMSr_{\rm RMS} Λ\Lambda EE rRMSr_{\rm RMS} P(𝕊526/𝔻522/𝔻524/𝔻526){}^{6}\mathbb{S}_{\frac{5}{2}}/{}^{2}\mathbb{D}_{\frac{5}{2}}/{}^{4}\mathbb{D}_{\frac{5}{2}}/{}^{6}\mathbb{D}_{\frac{5}{2}})
5/2{5}/{2}^{-} 1.64 0.31-0.31 4.27 1.64 0.80-0.80 2.97 99.81/0.02/0.01/0.15
1.66 4.93-4.93 1.19 1.66 5.81-5.81 1.11 99.76/0.03/0.01/0.20
1.68 13.13-13.13 0.74 1.67 9.55-9.55 0.87 99.77/0.03/0.01/0.19

To summarize, we predict two types of hidden-charm molecular pentaquark states with triple strangeness, i.e., the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} molecular state with JP=3/2J^{P}={3}/{2}^{-} and the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} molecular state with JP=5/2J^{P}=5/2^{-}. Here, we want to indicate that the effective potentials from the ϕ\phi and η\eta exchanges are attractive for the ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} system with JP=3/2J^{P}={3}/{2}^{-} and the ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} system with JP=5/2J^{P}=5/2^{-}, which is due to the contributions from the 𝐪2{\bf q}^{2} terms in the deduced effective potentials. In fact, this issue has been discussed in Ref. 1839195 .

III Decay behaviors of these possible Ωc()D¯s\Omega_{c}^{(*)}\bar{D}_{s}^{*} molecular states

In order to further reveal the inner structures and properties of the possible hidden-charm molecular pentaquarks with triple strangeness, we calculate the strong decay behaviors of these possible molecular candidates. In this work, we discuss the hidden-charm decay mode, the corresponding final states including the ηc(1S)Ω\eta_{c}(1S)\Omega and J/ψΩJ/\psi\Omega. Different with the binding of the possible hidden-charm molecular pentaquarks with triple strangeness, the interactions in the very short range distance contribute to the hidden-charm decay processes. Thus, the quark-interchange model Barnes:1991em ; Barnes:1999hs can be as a reasonable theoretical framework.

III.1 The quark-interchange model

When using the quark-interchange model to estimate the transition amplitudes in calculating the decay widths, we usually adopt the nonrelativistic quark model to describe the quark-quark interaction Wang:2019spc ; Xiao:2019spy , which is expressed as Wong:2001td

Vij(q2)=λi2λj2(4παsq2+6πbq48παs3mimjeq24σ2𝐬i𝐬j),V_{ij}(q^{2})=\frac{\lambda_{i}}{2}\cdot\frac{\lambda_{j}}{2}\left(\frac{4\pi\alpha_{s}}{q^{2}}+\frac{6\pi b}{q^{4}}-\frac{8\pi\alpha_{s}}{3m_{i}m_{j}}e^{-{\frac{q^{2}}{4\sigma^{2}}}}{\bf{s}}_{i}\cdot{\bf{s}}_{j}\right), (3.1)

where λi(λj)\lambda_{i}(\lambda_{j}), mi(mj)m_{i}(m_{j}), and 𝐬i(𝐬j){\bf{s}}_{i}({\bf{s}}_{j}) represent the color factor, the mass, and the spin operator of the interacting quarks, respectively. αs\alpha_{s} denotes the running coupling constant, which reads as Wong:2001td

αs(Q2)=12π(322nf)ln(A+Q2B2),\alpha_{s}(Q^{2})=\frac{12\pi}{\left(32-2n_{f}\right){\rm ln}\left(A+\frac{Q^{2}}{B^{2}}\right)}, (3.2)

where Q2Q^{2} is the square of the invariant mass of the interacting quarks, and the relevant parameters Wong:2001td in Eqs. (3.1) and  (3.2) are collected in Table 6.

Table 6: The parameters of the nonrelativistic quark model Wong:2001td and the oscillating parameters of the Gaussian function Wang:2019spc .
Quark model b(GeV2)b~{}(\rm{GeV}^{2}) σ(GeV)\sigma~{}(\rm{GeV}) AA
0.180 0.897 10
BB (GeV) ms(GeV)m_{s}~{}(\rm{GeV}) mc(GeV)m_{c}~{}(\rm{GeV})
0.310 0.575 1.776
Oscillating parameters βDs(GeV)\beta_{D^{\ast}_{s}}~{}(\rm{GeV}) βηc(GeV)\beta_{\eta_{c}}~{}(\rm{GeV}) βJ/ψ(GeV)\beta_{J/\psi}~{}(\rm{GeV})
0.562 0.838 0.729
αλΩ(GeV)\alpha_{\lambda\Omega}~{}(\rm{GeV}) αρΩ(GeV)\alpha_{\rho\Omega}~{}(\rm{GeV}) αλΩc(GeV)\alpha_{\lambda\Omega_{c}}~{}(\rm{GeV})
0.466 0.407 0.583
αρΩc(GeV)\alpha_{\rho\Omega_{c}}~{}(\rm{GeV}) αλΩc(GeV)\alpha_{\lambda\Omega^{\ast}_{c}}~{}(\rm{GeV}) αρΩc(GeV)\alpha_{\rho\Omega^{\ast}_{c}}~{}(\rm{GeV})
0.444 0.537 0.423

To get the transition amplitudes within the quark-interchange model, we take the same convention as the previous work Wang:2019spc ; Xiao:2019spy ; Wang:2020prk . The transition amplitude for the process A(css)+B(sc¯)C(sss)+D(cc¯)A(css)+B(s\bar{c})\to C(sss)+D(c\bar{c}) can be decomposed as four processes in the hadronic molecular picture, which are illustrated in Fig. 2.

Refer to caption
Figure 2: Quark-interchange diagrams for the process A(css)+B(sc¯)C(sss)+D(cc¯)A(css)+B(s\bar{c})\to C(sss)+D(c\bar{c}) in the hadronic molecular picture.

The Hamiltonian of the initial hidden-charm molecular pentaquark state can be written as Wang:2019spc

HInitial=HA0+HB0+VAB,H_{\rm{Initial}}=H^{0}_{A}+H^{0}_{B}+V_{AB}, (3.3)

where HA0H^{0}_{A} and HB0H^{0}_{B} are the Hamiltonian of the free baryon A and meson B, and VABV_{AB} denotes the interaction between the baryon A and the meson B.

Furthermore, we define the color wave function ωcolor\omega_{\rm{color}}, the flavor wave function χflavor\chi_{\rm{flavor}}, the spin wave function χspin\chi_{\rm{spin}}, and the momentum space wave function ϕ(𝐩)\phi(\bf{p}), respectively. Thus, the total wave function can be expressed as

ψtotal=ωcolorχflavorχspinϕ(𝐩).\psi_{\rm{total}}=\omega_{\rm{color}}\chi_{\rm{flavor}}\chi_{\rm{spin}}\phi(\bf{p}). (3.4)

In this work, we take the Gaussian functions to approximate the momentum space wave functions for the baryon, meson, and molecule. The more explicit forms of the relevant Gaussian function can be found in Ref. Wang:2019spc , and the oscillating parameters of the meson and baryon are estimated by fitting their mass spectrum in the Godfrey-Isgur model Godfrey:1985xj , which are listed in Table 6. For an SS-wave loosely bound state composed of two hadrons A and B, the oscillating parameter β\beta can be related to the mass of the molecular state mm, i.e., β=3μ(mA+mBm)\beta=\sqrt{3\mu(m_{A}+m_{B}-m)} with μ=mAmBmA+mB\mu=\frac{m_{A}m_{B}}{m_{A}+m_{B}} Weinberg:1962hj ; Weinberg:1963zza ; Guo:2017jvc . And then, the TT-matrix TfiT_{fi} represents the relevant effective potential in the quark-interchange diagrams, which can be factorized as

Tfi=IcolorIflavorIspinIspace,T_{fi}=I_{\rm{color}}I_{\rm{flavor}}I_{\rm{spin}}I_{\rm{space}}, (3.5)

where IiI_{i} with the subscripts color, flavor, spin, and space stand for the corresponding factors, and the calculation details of these factors Ii(i=color,flavor,spin,space)I_{i}\,(i=\rm{color},\,\rm{flavor},\,\rm{spin},\,\rm{space}) are referred to in Ref. Wang:2019spc .

For the two-body strong decay widths of these discussed molecular candidates, they can be explicitly expressed as

Γ=|𝐏C|32π2m2(2J+1)𝑑Ω||2.\displaystyle\Gamma=\frac{|{\bf{P}}_{C}|}{32\pi^{2}m^{2}(2J+1)}\int d\Omega|\mathcal{M}|^{2}. (3.6)

In the above expression, 𝐏C{\bf{P}}_{C}, mm, and \mathcal{M} stand for the momentum of the final state, the mass of the molecular state, and the transition amplitude of the discussed process, respectively. Here, we want to emphasize that there exists a relation of the transition amplitude \mathcal{M} and the TT-matrix TfiT_{fi}, i.e.,

=(2π)322m2EC2EDTfi,\displaystyle\mathcal{M}=-(2\pi)^{\frac{3}{2}}\sqrt{2m2E_{C}2E_{D}}T_{fi}, (3.7)

where ECE_{C} and EDE_{D} are the energies of the final states C and D, respectively. Through the above preparation, we can calculate the two-body hidden-charm strong decay widths of these proposed Ωc()D¯s\Omega_{c}^{(*)}\bar{D}_{s}^{*} molecular states.

III.2 Two-body hidden-charm strong decay widths of these proposed Ωc()D¯s\Omega_{c}^{(*)}\bar{D}_{s}^{*} molecular states

In the above section, our results suggest that the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} state with JP=3/2J^{P}={3}/{2}^{-} and the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} state with JP=5/2J^{P}=5/2^{-} can be regarded as the hidden-charm molecular pentaquark candidates with triple strangeness. Thus, we will study the two-body strong decay property of these possible hidden-charm molecular pentaquarks with triple strangeness, which provides valuable information to search for these proposed molecular candidates in experiment.

In this work, we focus on the two-body hidden-charm strong decay channels for these predicted hidden-charm molecular pentaquarks with triple strangeness. For the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} molecular state with JP=3/2J^{P}={3}/{2}^{-}, it can decay into the J/ψΩJ/\psi\,\Omega and ηcΩ\eta_{c}\,\Omega channels through the SS-wave interaction. For the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} molecular state with JP=5/2J^{P}={5}/{2}^{-}, we only take into account the J/ψΩJ/\psi\,\Omega decay channel via the SS-wave coupling, while the ηcΩ\eta_{c}\,\Omega channel is suppressed since it is a DD-wave decay Wang:2019spc .

In order to intuitively clarify the uncertainty of the binding energies, we present the binding energies dependence of the decay widths for the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} molecular state with JP=3/2J^{P}={3}/{2}^{-} and the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} molecular state with JP=5/2J^{P}={5}/{2}^{-} in Fig. 3. As stressed in Sec. II, the hadronic molecule is a loosely bound state Chen:2017xat , so the binding energies of these hidden-charm molecular pentaquarks with triple strangeness change from 20-20 to 1-1 MeV in calculating the decay widths. With increasing the absolute values of the binding energy, the decay widths become larger, which is consistent with other theoretical calculations Chen:2017xat ; Lin:2017mtz ; Lin:2018kcc ; Lin:2018nqd ; Shen:2019evi ; Lin:2019qiv ; Lin:2019tex ; Dong:2019ofp ; Dong:2020rgs ; Xiao:2019mvs ; Wu:2018xaa ; Chen:2017abq ; Xiao:2016mho .

Refer to caption
Refer to caption
Figure 3: The binding energies dependence of the decay widths for the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} molecular state with JP=3/2J^{P}={3}/{2}^{-} and the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} molecular state with JP=5/2J^{P}={5}/{2}^{-}.

As illustrated in Fig. 3, when the binding energies are taken as 15-15 MeV with typical values, the dominant decay channel is the J/ψΩJ/\psi\,\Omega around one MeV for the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} molecular state with JP=3/2J^{P}={3}/{2}^{-}, and the decay width of the J/ψΩJ/\psi\,\Omega channel is predicted to be around several MeV for the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} molecule with JP=5/2J^{P}={5}/{2}^{-}. Thus, the J/ψΩJ/\psi\,\Omega should be the promising channel to observe the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} molecular state with JP=3/2J^{P}={3}/{2}^{-} and the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} molecular state with JP=5/2J^{P}={5}/{2}^{-}. Meanwhile, it is interesting to note that the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} molecular state with JP=3/2J^{P}={3}/{2}^{-} prefers to decay into the J/ψΩJ/\psi\,\Omega channel, but the decay width of the ηcΩ\eta_{c}\,\Omega channel is comparable to the J/ψΩJ/\psi\,\Omega channel, which indicates that the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} molecule with JP=3/2J^{P}={3}/{2}^{-} can be detected in the ηcΩ\eta_{c}\,\Omega channel in future experiment.

In the heavy quark symmetry, the relative partial decay branch ratio between the ηc(1S)Ω\eta_{c}(1S)\Omega and J/ψΩJ/\psi\Omega for the ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} state with JP=3/2J^{P}=3/2^{-} can be estimated as

HQS=Γ(ΩcD¯sηc(1S)Ω)Γ(ΩcD¯sJ/ψΩ)=0.6,\displaystyle\mathcal{R}_{\text{HQS}}=\frac{\Gamma(\Omega_{c}\bar{D}_{s}^{*}\to\eta_{c}(1S)\Omega)}{\Gamma(\Omega_{c}\bar{D}_{s}^{*}\to J/\psi\Omega)}=0.6, (3.8)

since the relative momentum in the ηc(1S)Ω\eta_{c}(1S)\Omega channel is larger than that in the J/ψΩJ/\psi\Omega channel, (E)\mathcal{R}(E) should be a little larger than HQS=0.6\mathcal{R}_{\text{HQS}}=0.6, where EE is the binding energy. In our calculation, we obtain

(5MeV)=(10MeV)=0.62,(15MeV)=0.67.\displaystyle\mathcal{R}(-5~{}\text{MeV})=\mathcal{R}(-10~{}\text{MeV})=0.62,\quad\mathcal{R}(-15~{}\text{MeV})=0.67.

Obviously, our results are consistent with the estimation in the heavy quark limit.

IV Summary

Searching for exotic hadronic state is an interesting and important research topic of hadron physics. With accumulation of experimental data, the LHCb observed three narrow Pc(4312)P_{c}(4312), Pc(4440)P_{c}(4440), and Pc(4457)P_{c}(4457) in 2019 Aaij:2019vzc , and found the evidence of the Pcs(4459)P_{cs}(4459) as a hidden-charm pentaquark with strangeness Aaij:2020gdg . These progresses make us have reason to believe that there should exist a zoo of the hidden-charm molecular pentaquark. At present, the hidden-charm molecular pentaquark with triple strangeness is still missing, which inspires our interest in exploring how to find these intriguing hidden-charm molecular pentaquark states with triple strangeness.

Mass spectrum information is crucial to searching for them. In this work, we perform the dynamical calculation of the possible hidden-charm molecular pentaquark states with triple strangeness from the Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} interactions, where the effective potentials can be obtained by the OBE model. By finding bound state solutions of these discussed systems, we find that the most promising hidden-charm molecular pentaquarks with triple strangeness are the SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} state with JP=3/2J^{P}={3}/{2}^{-} and the SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} state with JP=5/2J^{P}=5/2^{-}. Besides mass spectrum study, we also discuss their two-body hidden-charm strong decay behaviors within the quark-interchange model. In concrete calculation, we mainly focus on the J/ψΩJ/\psi\,\Omega and ηcΩ\eta_{c}\,\Omega decay modes for the predicted SS-wave ΩcD¯s\Omega_{c}\bar{D}_{s}^{*} molecule with JP=3/2J^{P}={3}/{2}^{-} and the J/ψΩJ/\psi\,\Omega decay channel for the predicted SS-wave ΩcD¯s\Omega_{c}^{*}\bar{D}_{s}^{*} molecule with JP=5/2J^{P}={5}/{2}^{-}.

In the following years, the LHCb Collaboration will collect more experimental data at Run III and upgrade the High-Luminosity-LHC Bediaga:2018lhg . Experimental searches for these predicted hidden-charm molecular pentaquarks with triple strangeness are an area full of opportunities and challenges in future experiments.

ACKNOWLEDGMENTS

We would like to thank Z. W. Liu, G. J. Wang, L. Y. Xiao, and S. Q. Luo for very helpful discussions. This work is supported by the China National Funds for Distinguished Young Scientists under Grant No. 11825503, National Key Research and Development Program of China under Contract No. 2020YFA0406400, the 111 Project under Grant No. B20063, and the National Natural Science Foundation of China under Grant No. 12047501. R. C. is supported by the National Postdoctoral Program for Innovative Talent.

Appendix A Relevant subpotentials

Through the standard strategy Wang:2020dya ; Wang:2019nwt ; Wang:2019aoc , we can derive the effective potentials in the coordinate space for these investigated Ωc()D¯s()\Omega_{c}^{(*)}\bar{D}_{s}^{(*)} systems, i.e.,

𝒱ΩcD¯sΩcD¯s\displaystyle\mathcal{V}^{\Omega_{c}\bar{D}_{s}\rightarrow\Omega_{c}\bar{D}_{s}} =\displaystyle= AYf0C2Yϕ,\displaystyle-AY_{f_{0}}-\frac{C}{2}Y_{\phi}, (1.1)
𝒱ΩcD¯sΩcD¯s\displaystyle\mathcal{V}^{\Omega_{c}^{*}\bar{D}_{s}\rightarrow\Omega_{c}^{*}\bar{D}_{s}} =\displaystyle= A𝒜1Yf0C2𝒜1Yϕ,\displaystyle-A\mathcal{A}_{1}Y_{f_{0}}-\frac{C}{2}\mathcal{A}_{1}Y_{\phi}, (1.2)
𝒱ΩcD¯sΩcD¯s\displaystyle\mathcal{V}^{\Omega_{c}\bar{D}_{s}^{*}\rightarrow\Omega_{c}\bar{D}_{s}^{*}} =\displaystyle= A𝒜2Yf0+2B9[𝒜3𝒪r+𝒜4𝒫r]Yη\displaystyle-A\mathcal{A}_{2}Y_{f_{0}}+\frac{2B}{9}\left[\mathcal{A}_{3}\mathcal{O}_{r}+\mathcal{A}_{4}\mathcal{P}_{r}\right]Y_{\eta} (1.3)
C2𝒜2Yϕ2D9[2𝒜3𝒪r𝒜4𝒫r]Yϕ,\displaystyle-\frac{C}{2}\mathcal{A}_{2}Y_{\phi}-\frac{2D}{9}\left[2\mathcal{A}_{3}\mathcal{O}_{r}-\mathcal{A}_{4}\mathcal{P}_{r}\right]Y_{\phi},
𝒱ΩcD¯sΩcD¯s\displaystyle\mathcal{V}^{\Omega_{c}^{*}\bar{D}_{s}^{*}\rightarrow\Omega_{c}^{*}\bar{D}_{s}^{*}} =\displaystyle= A𝒜5Yf0B3[𝒜6𝒪r+𝒜7𝒫r]Yη\displaystyle-A\mathcal{A}_{5}Y_{f_{0}}-\frac{B}{3}\left[\mathcal{A}_{6}\mathcal{O}_{r}+\mathcal{A}_{7}\mathcal{P}_{r}\right]Y_{\eta} (1.4)
C2𝒜5Yϕ+D3[2𝒜6𝒪r𝒜7𝒫r]Yϕ,\displaystyle-\frac{C}{2}\mathcal{A}_{5}Y_{\phi}+\frac{D}{3}\left[2\mathcal{A}_{6}\mathcal{O}_{r}-\mathcal{A}_{7}\mathcal{P}_{r}\right]Y_{\phi},
𝒱ΩcD¯sΩcD¯s\displaystyle\mathcal{V}^{\Omega_{c}\bar{D}_{s}\rightarrow\Omega_{c}^{*}\bar{D}_{s}} =\displaystyle= A3𝒜8Yf01+C23𝒜8Yϕ1,\displaystyle\frac{A}{\sqrt{3}}\mathcal{A}_{8}Y_{f_{0}1}+\frac{C}{2\sqrt{3}}\mathcal{A}_{8}Y_{\phi 1}, (1.5)
𝒱ΩcD¯sΩcD¯s\displaystyle\mathcal{V}^{\Omega_{c}\bar{D}_{s}\rightarrow\Omega_{c}\bar{D}_{s}^{*}} =\displaystyle= 2B9[𝒜9𝒪r+𝒜10𝒫r]Yη2\displaystyle\frac{2B}{9}\left[\mathcal{A}_{9}\mathcal{O}_{r}+\mathcal{A}_{10}\mathcal{P}_{r}\right]Y_{\eta 2} (1.6)
+2D9[2𝒜9𝒪r𝒜10𝒫r]Yϕ2,\displaystyle+\frac{2D}{9}\left[2\mathcal{A}_{9}\mathcal{O}_{r}-\mathcal{A}_{10}\mathcal{P}_{r}\right]Y_{\phi 2},
𝒱ΩcD¯sΩcD¯s\displaystyle\mathcal{V}^{\Omega_{c}\bar{D}_{s}\rightarrow\Omega_{c}^{*}\bar{D}_{s}^{*}} =\displaystyle= B33[𝒜11𝒪r+𝒜12𝒫r]Yη3\displaystyle-\frac{B}{3\sqrt{3}}\left[\mathcal{A}_{11}\mathcal{O}_{r}+\mathcal{A}_{12}\mathcal{P}_{r}\right]Y_{\eta 3} (1.7)
D33[2𝒜11𝒪r𝒜12𝒫r]Yϕ3,\displaystyle-\frac{D}{3\sqrt{3}}\left[2\mathcal{A}_{11}\mathcal{O}_{r}-\mathcal{A}_{12}\mathcal{P}_{r}\right]Y_{\phi 3},
𝒱ΩcD¯sΩcD¯s\displaystyle\mathcal{V}^{\Omega_{c}^{*}\bar{D}_{s}\rightarrow\Omega_{c}\bar{D}_{s}^{*}} =\displaystyle= B33[𝒜13𝒪r+𝒜14𝒫r]Yη4\displaystyle\frac{B}{3\sqrt{3}}\left[\mathcal{A}_{13}\mathcal{O}_{r}+\mathcal{A}_{14}\mathcal{P}_{r}\right]Y_{\eta 4} (1.8)
+D33[2𝒜13𝒪r𝒜14𝒫r]Yϕ4,\displaystyle+\frac{D}{3\sqrt{3}}\left[2\mathcal{A}_{13}\mathcal{O}_{r}-\mathcal{A}_{14}\mathcal{P}_{r}\right]Y_{\phi 4},
𝒱ΩcD¯sΩcD¯s\displaystyle\mathcal{V}^{\Omega_{c}^{*}\bar{D}_{s}\rightarrow\Omega_{c}^{*}\bar{D}_{s}^{*}} =\displaystyle= B3[𝒜15𝒪r+𝒜16𝒫r]Yη5\displaystyle\frac{B}{3}\left[\mathcal{A}_{15}\mathcal{O}_{r}+\mathcal{A}_{16}\mathcal{P}_{r}\right]Y_{\eta 5} (1.9)
+D3[2𝒜15𝒪r𝒜16𝒫r]Yϕ5,\displaystyle+\frac{D}{3}\left[2\mathcal{A}_{15}\mathcal{O}_{r}-\mathcal{A}_{16}\mathcal{P}_{r}\right]Y_{\phi 5},
𝒱ΩcD¯sΩcD¯s\displaystyle\mathcal{V}^{\Omega_{c}\bar{D}_{s}^{*}\rightarrow\Omega_{c}^{*}\bar{D}_{s}^{*}} =\displaystyle= A3𝒜17Yf06+B33[𝒜18𝒪r+𝒜19𝒫r]Yη6\displaystyle\frac{A}{\sqrt{3}}\mathcal{A}_{17}Y_{f_{0}6}+\frac{B}{3\sqrt{3}}\left[\mathcal{A}_{18}\mathcal{O}_{r}+\mathcal{A}_{19}\mathcal{P}_{r}\right]Y_{\eta 6}
+C𝒜1723Yϕ6D33[2𝒜18𝒪r𝒜19𝒫r]Yϕ6.\displaystyle+\frac{C\mathcal{A}_{17}}{2\sqrt{3}}Y_{\phi 6}-\frac{D}{3\sqrt{3}}\left[2\mathcal{A}_{18}\mathcal{O}_{r}-\mathcal{A}_{19}\mathcal{P}_{r}\right]Y_{\phi 6}.

Here, 𝒪r=1r2rr2r\mathcal{O}_{r}=\frac{1}{r^{2}}\frac{\partial}{\partial r}r^{2}\frac{\partial}{\partial r} and 𝒫r=rr1rr\mathcal{P}_{r}=r\frac{\partial}{\partial r}\frac{1}{r}\frac{\partial}{\partial r}. Additionally, we also define several variables, which include A=lSgSA=l_{S}g_{S}, B=g1g/fπ2B=g_{1}g/f_{\pi}^{2}, C=βSβgV2C=\beta_{S}\beta g_{V}^{2}, and D=λSλgV2D=\lambda_{S}\lambda g_{V}^{2}. The function YiY_{i} can be defined as

Yi=emireΛir4πrΛi2mi28πΛieΛir.\displaystyle Y_{i}=\dfrac{e^{-m_{i}r}-e^{-\Lambda_{i}r}}{4\pi r}-\dfrac{\Lambda_{i}^{2}-m_{i}^{2}}{8\pi\Lambda_{i}}e^{-\Lambda_{i}r}. (1.11)

Here, mi=m2qi2m_{i}=\sqrt{m^{2}-q_{i}^{2}} and Λi=Λ2qi2\Lambda_{i}=\sqrt{\Lambda^{2}-q_{i}^{2}}. Variables qi(i=1,, 6)q_{i}\,(i=1\,,...,\,6) are defined as q1=0.04q_{1}=0.04 GeV, q2=0.06q_{2}=0.06 GeV, q3=0.02q_{3}=0.02 GeV, q4=0.10q_{4}=0.10 GeV, q5=0.06q_{5}=0.06 GeV, and q6=0.04q_{6}=0.04 GeV.

In the above effective potentials, we also introduce several operators, i.e.,

𝒜1\displaystyle\mathcal{A}_{1} =\displaystyle= a,b,m,nC12a,1b32,a+bC12m,1n32,m+nχ3a(ϵ3bϵ1n)χ1m,\displaystyle\sum_{a,b,m,n}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}C^{\frac{3}{2},m+n}_{\frac{1}{2}m,1n}\chi^{\dagger a}_{3}\left({\bm{\epsilon}^{\dagger b}_{3}}\cdot{\bm{\epsilon}^{n}_{1}}\right)\chi^{m}_{1},
𝒜2\displaystyle\mathcal{A}_{2} =\displaystyle= χ3(ϵ4ϵ2)χ1,\displaystyle\chi^{\dagger}_{3}\left({\bm{\epsilon}^{\dagger}_{4}}\cdot{\bm{\epsilon}_{2}}\right)\chi_{1},
𝒜3\displaystyle\mathcal{A}_{3} =\displaystyle= χ3[𝝈(iϵ2×ϵ4)]χ1,\displaystyle\chi^{\dagger}_{3}\left[{\bm{\sigma}}\cdot\left(i{\bm{\epsilon}_{2}}\times{\bm{\epsilon}^{\dagger}_{4}}\right)\right]\chi_{1},
𝒜4\displaystyle\mathcal{A}_{4} =\displaystyle= χ3T(𝝈,iϵ2×ϵ4)χ1,\displaystyle\chi^{\dagger}_{3}T({\bm{\sigma}},i{\bm{\epsilon}_{2}}\times{\bm{\epsilon}^{\dagger}_{4}})\chi_{1},
𝒜5\displaystyle\mathcal{A}_{5} =\displaystyle= a,b,m,nC12a,1b32,a+bC12m,1n32,m+nχ3a(ϵ1nϵ3b)(ϵ2ϵ4)χ1m,\displaystyle\sum_{a,b,m,n}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}C^{\frac{3}{2},m+n}_{\frac{1}{2}m,1n}\chi^{\dagger a}_{3}\left({\bm{\epsilon}^{n}_{1}}\cdot{\bm{\epsilon}^{\dagger b}_{3}}\right)\left({\bm{\epsilon}_{2}}\cdot{\bm{\epsilon}^{\dagger}_{4}}\right)\chi^{m}_{1},
𝒜6\displaystyle\mathcal{A}_{6} =\displaystyle= a,b,m,nC12a,1b32,a+bC12m,1n32,m+nχ3a(ϵ1n×ϵ3b)(ϵ2×ϵ4)χ1m,\displaystyle\sum_{a,b,m,n}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}C^{\frac{3}{2},m+n}_{\frac{1}{2}m,1n}\chi^{\dagger a}_{3}\left({\bm{\epsilon}^{n}_{1}}\times{\bm{\epsilon}^{\dagger b}_{3}}\right)\cdot\left({\bm{\epsilon}_{2}}\times{\bm{\epsilon}^{\dagger}_{4}}\right)\chi^{m}_{1},
𝒜7\displaystyle\mathcal{A}_{7} =\displaystyle= a,b,m,nC12a,1b32,a+bC12m,1n32,m+nχ3aT(ϵ1n×ϵ3b,ϵ2×ϵ4)χ1m,\displaystyle\sum_{a,b,m,n}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}C^{\frac{3}{2},m+n}_{\frac{1}{2}m,1n}\chi^{\dagger a}_{3}T({\bm{\epsilon}^{n}_{1}}\times{\bm{\epsilon}^{\dagger b}_{3}},{\bm{\epsilon}_{2}}\times{\bm{\epsilon}^{\dagger}_{4}})\chi^{m}_{1},
𝒜8\displaystyle\mathcal{A}_{8} =\displaystyle= a,bC12a,1b32,a+bχ3a(ϵ3b𝝈)χ1,\displaystyle\sum_{a,b}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}\chi^{\dagger a}_{3}\left({\bm{\epsilon}^{\dagger b}_{3}}\cdot{\bm{\sigma}}\right)\chi_{1},
𝒜9\displaystyle\mathcal{A}_{9} =\displaystyle= χ3(𝝈ϵ4)χ1,\displaystyle\chi^{\dagger}_{3}\left({\bm{\sigma}}\cdot{\bm{\epsilon}^{\dagger}_{4}}\right)\chi_{1},
𝒜10\displaystyle\mathcal{A}_{10} =\displaystyle= χ3T(𝝈,ϵ4)χ1,\displaystyle\chi^{\dagger}_{3}T({\bm{\sigma}},{\bm{\epsilon}^{\dagger}_{4}})\chi_{1},
𝒜11\displaystyle\mathcal{A}_{11} =\displaystyle= a,bC12a,1b32,a+bχ3a[ϵ4(i𝝈×ϵ3b)]χ1,\displaystyle\sum_{a,b}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}\chi^{\dagger a}_{3}\left[{\bm{\epsilon}^{\dagger}_{4}}\cdot\left(i{\bm{\sigma}}\times{\bm{\epsilon}^{\dagger b}_{3}}\right)\right]\chi_{1},
𝒜12\displaystyle\mathcal{A}_{12} =\displaystyle= a,bC12a,1b32,a+bχ3aT(ϵ4,i𝝈×ϵ3b)χ1,\displaystyle\sum_{a,b}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}\chi^{\dagger a}_{3}T({\bm{\epsilon}^{\dagger}_{4}},i{\bm{\sigma}}\times{\bm{\epsilon}^{\dagger b}_{3}})\chi_{1},
𝒜13\displaystyle\mathcal{A}_{13} =\displaystyle= a,bC12a,1b32,a+bχ3[ϵ4(i𝝈×ϵ1b)]χ1a,\displaystyle\sum_{a,b}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}\chi^{\dagger}_{3}\left[{\bm{\epsilon}^{\dagger}_{4}}\cdot\left(i{\bm{\sigma}}\times{\bm{\epsilon}^{b}_{1}}\right)\right]\chi^{a}_{1},
𝒜14\displaystyle\mathcal{A}_{14} =\displaystyle= a,bC12a,1b32,a+bχ3T(ϵ4,i𝝈×ϵ1b)χ1a,\displaystyle\sum_{a,b}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}\chi^{\dagger}_{3}T({\bm{\epsilon}^{\dagger}_{4}},i{\bm{\sigma}}\times{\bm{\epsilon}^{b}_{1}})\chi^{a}_{1},
𝒜15\displaystyle\mathcal{A}_{15} =\displaystyle= a,b,m,nC12a,1b32,a+bC12m,1n32,m+nχ3a[ϵ4(iϵ1n×ϵ3b)]χ1m,\displaystyle\sum_{a,b,m,n}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}C^{\frac{3}{2},m+n}_{\frac{1}{2}m,1n}\chi^{\dagger a}_{3}\left[{\bm{\epsilon}^{\dagger}_{4}}\cdot\left(i{\bm{\epsilon}^{n}_{1}}\times{\bm{\epsilon}^{\dagger b}_{3}}\right)\right]\chi^{m}_{1},
𝒜16\displaystyle\mathcal{A}_{16} =\displaystyle= a,b,m,nC12a,1b32,a+bC12m,1n32,m+nχ3aT(ϵ4,iϵ1n×ϵ3b)χ1m,\displaystyle\sum_{a,b,m,n}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}C^{\frac{3}{2},m+n}_{\frac{1}{2}m,1n}\chi^{\dagger a}_{3}T({\bm{\epsilon}^{\dagger}_{4}},i{\bm{\epsilon}^{n}_{1}}\times{\bm{\epsilon}^{\dagger b}_{3}})\chi^{m}_{1},
𝒜17\displaystyle\mathcal{A}_{17} =\displaystyle= a,bC12a,1b32,a+bχ3a(𝝈ϵ3b)(ϵ2ϵ4)χ1,\displaystyle\sum_{a,b}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}\chi^{\dagger a}_{3}\left({\bm{\sigma}}\cdot{\bm{\epsilon}^{\dagger b}_{3}}\right)\left({\bm{\epsilon}_{2}}\cdot{\bm{\epsilon}^{\dagger}_{4}}\right)\chi_{1},
𝒜18\displaystyle\mathcal{A}_{18} =\displaystyle= a,bC12a,1b32,a+bχ3a(𝝈×ϵ3b)(ϵ2×ϵ4)χ1,\displaystyle\sum_{a,b}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}\chi^{\dagger a}_{3}\left({\bm{\sigma}}\times{\bm{\epsilon}^{\dagger b}_{3}}\right)\cdot\left({\bm{\epsilon}_{2}}\times{\bm{\epsilon}^{\dagger}_{4}}\right)\chi_{1},
𝒜19\displaystyle\mathcal{A}_{19} =\displaystyle= a,bC12a,1b32,a+bχ3aT(𝝈×ϵ3b,ϵ2×ϵ4)χ1.\displaystyle\sum_{a,b}C^{\frac{3}{2},a+b}_{\frac{1}{2}a,1b}\chi^{\dagger a}_{3}T({\bm{\sigma}}\times{\bm{\epsilon}^{\dagger b}_{3}},{\bm{\epsilon}_{2}}\times{\bm{\epsilon}^{\dagger}_{4}})\chi_{1}. (1.12)

Here, T(𝒙,𝒚)=3(𝒓^𝒙)(𝒓^𝒚)𝒙𝒚T({\bm{x}},{\bm{y}})=3\left(\hat{\bm{r}}\cdot{\bm{x}}\right)\left(\hat{\bm{r}}\cdot{\bm{y}}\right)-{\bm{x}}\cdot{\bm{y}} is the tensor force operator. In Table 7, we collect the numerical matrix elements f|𝒜k|i(k=1,, 7)\langle f|\mathcal{A}_{k}|i\rangle\,(k=1\,,...,\,7) with the SS-DD wave mixing effect analysis. Of course, the relevant numerical matrix elements f|𝒜k|i(k=8,, 19)\langle f|\mathcal{A}_{k}|i\rangle\,(k=8\,,...,\,19) will be involved in the coupled channel analysis. For the coupled channel analysis with J=1/2J=1/2, we have 𝒜9=3\mathcal{A}_{9}=\sqrt{3}, 𝒜11=2\mathcal{A}_{11}=\sqrt{2}, 𝒜18=2/3\mathcal{A}_{18}=-\sqrt{{2}/{3}}, and 𝒜k=0(k=10, 12, 17, 19)\mathcal{A}_{k}=0\,(k=10,\,12,\,17,\,19). And, there exists 𝒜13=1\mathcal{A}_{13}=1, 𝒜15=5/3\mathcal{A}_{15}=\sqrt{{5}/{3}}, 𝒜18=5/3\mathcal{A}_{18}=-\sqrt{{5}/{3}}, and 𝒜k=0(k=14, 16, 17, 19)\mathcal{A}_{k}=0\,(k=14,\,16,\,17,\,19) for the coupled channel analysis with J=3/2J=3/2.

Table 7: The numerical matrix elements f|𝒜k|i(k=1,, 7)\langle f|\mathcal{A}_{k}|i\rangle\,(k=1\,,...,\,7) with the SS-DD wave mixing effect analysis.
Matrix elements J=1/2J=1/2 J=3/2J=3/2 J=5/2J=5/2
ΩcD¯s|𝒜1|ΩcD¯s\langle\Omega_{c}^{*}\bar{D}_{s}|\mathcal{A}_{1}|\Omega_{c}^{*}\bar{D}_{s}\rangle // diag(1,1) //
ΩcD¯s|𝒜2|ΩcD¯s\langle\Omega_{c}\bar{D}_{s}^{*}|\mathcal{A}_{2}|\Omega_{c}\bar{D}_{s}^{*}\rangle diag(1,1) diag(1,1,1) //
ΩcD¯s|𝒜3|ΩcD¯s\langle\Omega_{c}\bar{D}_{s}^{*}|\mathcal{A}_{3}|\Omega_{c}\bar{D}_{s}^{*}\rangle diag(2-2,11) diag(11,2-2,11) //
ΩcD¯s|𝒜4|ΩcD¯s\langle\Omega_{c}\bar{D}_{s}^{*}|\mathcal{A}_{4}|\Omega_{c}\bar{D}_{s}^{*}\rangle (0222)\left(\begin{array}[]{cc}0&-\sqrt{2}\\ -\sqrt{2}&-2\end{array}\right) (012101210)\left(\begin{array}[]{ccc}0&1&2\\ 1&0&-1\\ 2&-1&0\end{array}\right) //
ΩcD¯s|𝒜5|ΩcD¯s\langle\Omega_{c}^{*}\bar{D}_{s}^{*}|\mathcal{A}_{5}|\Omega_{c}^{*}\bar{D}_{s}^{*}\rangle diag(1,1,1) diag(1,1,1,1) diag(1,1,1,1)
ΩcD¯s|𝒜6|ΩcD¯s\langle\Omega_{c}^{*}\bar{D}_{s}^{*}|\mathcal{A}_{6}|\Omega_{c}^{*}\bar{D}_{s}^{*}\rangle diag(53\frac{5}{3},23\frac{2}{3},1-1) diag(23\frac{2}{3},53\frac{5}{3},23\frac{2}{3},1-1) diag(1-1,53\frac{5}{3},23\frac{2}{3},1-1)
ΩcD¯s|𝒜7|ΩcD¯s\langle\Omega_{c}^{*}\bar{D}_{s}^{*}|\mathcal{A}_{7}|\Omega_{c}^{*}\bar{D}_{s}^{*}\rangle (073525735161515251585)\left(\begin{array}[]{ccc}0&-\frac{7}{3\sqrt{5}}&\frac{2}{\sqrt{5}}\\ -\frac{7}{3\sqrt{5}}&\frac{16}{15}&-\frac{1}{5}\\ \frac{2}{\sqrt{5}}&-\frac{1}{5}&\frac{8}{5}\end{array}\right) (07310161575273100731023516157310011475223511447)\left(\begin{array}[]{cccc}0&\frac{7}{3\sqrt{10}}&-\frac{16}{15}&-\frac{\sqrt{7}}{5\sqrt{2}}\\ \frac{7}{3\sqrt{10}}&0&-\frac{7}{3\sqrt{10}}&-\frac{2}{\sqrt{35}}\\ -\frac{16}{15}&-\frac{7}{3\sqrt{10}}&0&-\frac{1}{\sqrt{14}}\\ -\frac{\sqrt{7}}{5\sqrt{2}}&-\frac{2}{\sqrt{35}}&-\frac{1}{\sqrt{14}}&\frac{4}{7}\end{array}\right) (02157532145215073542105753735162127321454210527347)\left(\begin{array}[]{cccc}0&\frac{2}{\sqrt{15}}&\frac{\sqrt{7}}{5\sqrt{3}}&-\frac{2\sqrt{14}}{5}\\ \frac{2}{\sqrt{15}}&0&\frac{\sqrt{7}}{3\sqrt{5}}&-\frac{4\sqrt{2}}{\sqrt{105}}\\ \frac{\sqrt{7}}{5\sqrt{3}}&\frac{\sqrt{7}}{3\sqrt{5}}&-\frac{16}{21}&-\frac{\sqrt{2}}{7\sqrt{3}}\\ -\frac{2\sqrt{14}}{5}&-\frac{4\sqrt{2}}{\sqrt{105}}&-\frac{\sqrt{2}}{7\sqrt{3}}&-\frac{4}{7}\end{array}\right)

References

  • (1) S. K. Choi et al. (Belle Collaboration), Observation of a Narrow Charmonium-Like State in Exclusive B±K±π+πJ/ψB^{\pm}\to K^{\pm}\pi^{+}\pi^{-}J/\psi Decays, Phys. Rev. Lett.  91, 262001 (2003).
  • (2) H. X. Chen, W. Chen, X. Liu, and S. L. Zhu, The hidden-charm pentaquark and tetraquark states, Phys. Rep.  639, 1 (2016).
  • (3) Y. R. Liu, H. X. Chen, W. Chen, X. Liu, and S. L. Zhu, Pentaquark and tetraquark states, Prog. Part. Nucl. Phys.  107, 237 (2019).
  • (4) S. L. Olsen, T. Skwarnicki, and D. Zieminska, Nonstandard heavy mesons and baryons: Experimental evidence, Rev. Mod. Phys.  90, 015003 (2018).
  • (5) F. K. Guo, C. Hanhart, U. G. Meiß\ssner, Q. Wang, Q. Zhao, and B. S. Zou, Hadronic molecules, Rev. Mod. Phys.  90, 015004 (2018).
  • (6) X. Liu, An overview of XYZXYZ new particles, Chin. Sci. Bull.  59, 3815 (2014).
  • (7) A. Hosaka, T. Iijima, K. Miyabayashi, Y. Sakai, and S. Yasui, Exotic hadrons with heavy flavors: XX, YY, ZZ, and related states, Prog. Theor. Exp. Phys. 2016, 062C01 (2016).
  • (8) N. Brambilla, S. Eidelman, C. Hanhart, A. Nefediev, C. P. Shen, C. E. Thomas, A. Vairo, and C. Z. Yuan, The XYZXYZ states: Experimental and theoretical status and perspectives, Phys. Rep. 873, 1 (2020).
  • (9) M. Gell-Mann, A schematic model of baryons and mesons, Phys. Lett.  8, 214 (1964).
  • (10) G. Zweig, An SU(3) model for strong interaction symmetry and its breaking. Version 1, CERN, Report No. CERN-TH-401, 1964.
  • (11) J. J. Wu, R. Molina, E. Oset, and B. S. Zou, Prediction of Narrow NN^{*} and Λ\Lambda^{*} Resonances with Hidden Charm Above 4 GeV, Phys. Rev. Lett.  105, 232001 (2010).
  • (12) W. L. Wang, F. Huang, Z. Y. Zhang, and B. S. Zou, ΣcD¯\Sigma_{c}\bar{D} and ΛcD¯\Lambda_{c}\bar{D} states in a chiral quark model, Phys. Rev. C 84, 015203 (2011).
  • (13) Z. C. Yang, Z. F. Sun, J. He, X. Liu, and S. L. Zhu, The possible hidden-charm molecular baryons composed of anticharmed meson and charmed baryon, Chin. Phys. C 36, 6 (2012).
  • (14) J. J. Wu, T.-S. H. Lee, and B. S. Zou, Nucleon resonances with hidden charm in coupled-channel Models, Phys. Rev. C 85, 044002 (2012).
  • (15) X. Q. Li and X. Liu, A possible global group structure for exotic states, Eur. Phys. J. C 74, 3198 (2014).
  • (16) R. Chen, X. Liu, X. Q. Li, and S. L. Zhu, Identifying Exotic Hidden-Charm Pentaquarks, Phys. Rev. Lett.  115, 132002 (2015).
  • (17) M. Karliner and J. L. Rosner, New Exotic Meson and Baryon Resonances from Doubly-Heavy Hadronic Molecules, Phys. Rev. Lett.  115, 122001 (2015).
  • (18) R. Aaij et al. (LHCb Collaboration), Observation of J/ψJ/\psi Resonances Consistent with Pentaquark States in Λb0J/ψKp\Lambda_{b}^{0}\rightarrow J/\psi K^{-}p Decays, Phys. Rev. Lett.  115, 072001 (2015).
  • (19) R. Aaij et al. (LHCb Collaboration), Observation of a Narrow Pentaquark State, Pc(4312)+P_{c}(4312)^{+}, and of Two-Peak Structure of the Pc(4450)+P_{c}(4450)^{+}, Phys. Rev. Lett.  122, 222001 (2019).
  • (20) R. Aaij et al. (LHCb Collaboration), Evidence of a J/ψΛJ/\psi\varLambda structure and observation of excited Ξ\Xi^{-} states in the ΞbJ/ψΛK\Xi_{b}^{-}\to J/\psi\varLambda K^{-} decay, arXiv:2012.10380.
  • (21) J. Hofmann and M. F. M. Lutz, Coupled-channel study of crypto-exotic baryons with charm, Nucl. Phys. A 763, 90 (2005).
  • (22) J. J. Wu, R. Molina, E. Oset and B. S. Zou, Dynamically generated NN^{*} and Λ\Lambda^{*} resonances in the hidden charm sector around 4.3 GeV, Phys. Rev. C 84, 015202 (2011).
  • (23) V. V. Anisovich, M. A. Matveev, J. Nyiri, A. V. Sarantsev, and A. N. Semenova, Nonstrange and strange pentaquarks with hidden charm, Int. J. Mod. Phys. A 30, 1550190 (2015).
  • (24) Z. G. Wang, Analysis of the 12±{\frac{1}{2}}^{\pm} pentaquark states in the diquark-diquark-antiquark model with QCD sum rules, Eur. Phys. J. C 76, 142 (2016).
  • (25) A. Feijoo, V. K. Magas, A. Ramos, and E. Oset, A hidden-charm S=1S=-1 pentaquark from the decay of Λb\Lambda_{b} into J/ψηΛJ/\psi\eta\Lambda states, Eur. Phys. J. C 76, no. 8, 446 (2016).
  • (26) J. X. Lu, E. Wang, J. J. Xie, L. S. Geng, and E. Oset, The ΛbJ/ψK0Λ\Lambda_{b}\rightarrow J/\psi K^{0}\Lambda reaction and a hidden-charm pentaquark state with strangeness, Phys. Rev. D 93, 094009 (2016).
  • (27) H. X. Chen, L. S. Geng, W. H. Liang, E. Oset, E. Wang, and J. J. Xie, Looking for a hidden-charm pentaquark state with strangeness S=1S=-1 from Ξb\Xi^{-}_{b} decay into J/ψKΛJ/\psi K^{-}\Lambda, Phys. Rev. C 93, 065203 (2016).
  • (28) R. Chen, J. He, and X. Liu, Possible strange hidden-charm pentaquarks from Σc()D¯s\Sigma_{c}^{(*)}\bar{D}_{s}^{*} and Ξc(,)D¯\Xi^{(^{\prime},*)}_{c}\bar{D}^{*} interactions, Chin. Phys. C 41, 103105 (2017).
  • (29) C. W. Xiao, J. Nieves, and E. Oset, Prediction of hidden charm strange molecular baryon states with heavy quark spin symmetry, Phys. Lett. B 799, 135051 (2019).
  • (30) C. W. Shen, H. J. Jing, F. K. Guo, and J. J. Wu, Exploring possible triangle singularities in the ΞbKJ/ψΛ\Xi^{-}_{b}\to K^{-}J/\psi\Lambda decay, Symmetry 12, 1611 (2020).
  • (31) B. Wang, L. Meng, and S. L. Zhu, Spectrum of the strange hidden charm molecular pentaquarks in chiral effective field theory, Phys. Rev. D 101, 034018 (2020).
  • (32) Q. Zhang, B. R. He, and J. L. Ping, Pentaquarks with the qqsQ¯Qqqs\bar{Q}Q configuration in the Chiral Quark Model, arXiv:2006.01042.
  • (33) H. X. Chen, W. Chen, X. Liu, and X. H. Liu, Establishing the first hidden-charm pentaquark with strangeness, arXiv:2011.01079.
  • (34) F. Z. Peng, M. J. Yan, M. Sánchez Sánchez, and M. P. Valderrama, The Pcs(4459)P_{cs}(4459) pentaquark from a combined effective field theory and phenomenological perspectives, arXiv:2011.01915.
  • (35) R. Chen, Can the newly Pcs(4459)P_{cs}(4459) be a strange hidden-charm ΞcD¯\Xi_{c}\bar{D}^{*} molecular pentaquarks?, arXiv:2011.07214.
  • (36) H. X. Chen, Hidden-charm pentaquark states through the current algebra: From their productions to decays, arXiv:2011.07187.
  • (37) M. Z. Liu, Y. W. Pan, and L. S. Geng, Can discovery of hidden charm strange pentaquark states help determine the spins of Pc(4440)P_{c}(4440) and Pc(4457)P_{c}(4457) ?, Phys. Rev. D 103, 034003 (2021).
  • (38) X. K. Dong, F. K. Guo, and B. S. Zou, A survey of heavy-antiheavy hadronic molecules, arXiv:2101.01021.
  • (39) F. L. Wang, R. Chen, and X. Liu, Prediction of hidden-charm pentaquarks with double strangeness, Phys. Rev. D 103, 034014 (2021).
  • (40) F. L. Wang, R. Chen, Z. W. Liu, and X. Liu, Probing new types of PcP_{c} states inspired by the interaction between SS-wave charmed baryon and anti-charmed meson in a T¯\bar{T} doublet, Phys. Rev. C 101, 025201 (2020).
  • (41) T. Barnes and E. S. Swanson, A Diagrammatic approach to meson meson scattering in the nonrelativistic quark potential model, Phys. Rev. D 46, 131 (1992).
  • (42) T. Barnes, N. Black, D. J. Dean, and E. S. Swanson, BB intermeson potentials in the quark model, Phys. Rev. C 60, 045202 (1999).
  • (43) T. Barnes, N. Black, and E. S. Swanson, Meson meson scattering in the quark model: Spin dependence and exotic channels, Phys. Rev. C 63, 025204 (2001).
  • (44) J. P. Hilbert, N. Black, T. Barnes, and E. S. Swanson, Charmonium-nucleon dissociation cross sections in the quark model, Phys. Rev. C 75, 064907 (2007).
  • (45) G. J. Wang, X. H. Liu, L. Ma, X. Liu, X. L. Chen, W. Z. Deng, and S. L. Zhu, The strong decay patterns of ZcZ_{c} and ZbZ_{b} states in the relativized quark model, Eur. Phys. J. C 79, 567 (2019).
  • (46) G. J. Wang, L. Y. Xiao, R. Chen, X. H. Liu, X. Liu, and S. L. Zhu, Probing hidden-charm decay properties of PcP_{c} states in a molecular scenario, Phys. Rev. D 102, 036012 (2020).
  • (47) L. Y. Xiao, G. J. Wang, and S. L. Zhu, Hidden-charm strong decays of the ZcZ_{c} states, Phys. Rev. D 101, 054001 (2020).
  • (48) G. J. Wang, L. Meng, L. Y. Xiao, M. Oka, and S. L. Zhu, Mass spectrum and strong decays of tetraquark c¯s¯qq\bar{c}\bar{s}qq states, arXiv:2010.09395.
  • (49) R. Aaij et al. (LHCb Collaboration), Physics case for an LHCb Upgrade II-Opportunities in flavour physics, and beyond, in the HL-LHC era, arXiv:1808.08865.
  • (50) M. B. Wise, Chiral perturbation theory for hadrons containing a heavy quark, Phys. Rev. D 45, R2188 (1992).
  • (51) R. Chen, A. Hosaka, and X. Liu, Searching for possible Ωc\Omega_{c}-like molecular states from meson-baryon interaction, Phys. Rev. D 97, 036016 (2018).
  • (52) G. J. Ding, Are Y(4260)Y(4260) and Z2+Z_{2}^{+}(4250) D1D{\rm D_{1}D} or D0D{\rm D_{0}D^{*}} hadronic molecules? Phys. Rev. D 79, 014001 (2009).
  • (53) R. Casalbuoni, A. Deandrea, N. Di Bartolomeo, R. Gatto, F. Feruglio, and G. Nardulli, Light vector resonances in the effective chiral Lagrangian for heavy mesons, Phys. Lett. B 292, 371 (1992).
  • (54) R. Casalbuoni, A. Deandrea, N. Di Bartolomeo, R. Gatto, F. Feruglio, and G. Nardulli, Phenomenology of heavy meson chiral Lagrangians, Phys. Rep.  281, 145 (1997).
  • (55) T. M. Yan, H. Y. Cheng, C. Y. Cheung, G. L. Lin, Y. C. Lin, and H. L. Yu, Heavy quark symmetry and chiral dynamics, Phys. Rev. D 46, 1148 (1992); [Phys. Rev. D 55, 5851E (1997)].
  • (56) M. Bando, T. Kugo, and K. Yamawaki, Nonlinear realization and hidden local symmetries, Phys. Rep.  164, 217 (1988).
  • (57) M. Harada and K. Yamawaki, Hidden local symmetry at loop: A new perspective of composite gauge boson and chiral phase transition, Phys. Rep.  381, 1 (2003).
  • (58) D. O. Riska and G. E. Brown, Nucleon resonance transition couplings to vector mesons, Nucl. Phys. A 679, 577 (2001).
  • (59) R. Chen, Z. F. Sun, X. Liu, and S. L. Zhu, Strong LHCb evidence supporting the existence of the hidden-charm molecular pentaquarks, Phys. Rev. D 100, 011502 (2019).
  • (60) J. He, Study of Pc(4457)P_{c}(4457), Pc(4440)P_{c}(4440), and Pc(4312)P_{c}(4312) in a quasipotential Bethe-Salpeter equation approach, Eur. Phys. J. C 79, 393 (2019).
  • (61) J. He, D¯Σc\bar{D}\Sigma^{*}_{c} and D¯Σc\bar{D}^{*}\Sigma_{c} interactions and the LHCb hidden-charmed pentaquarks, Phys. Lett. B 753, 547 (2016).
  • (62) R. Chen, F. L. Wang, A. Hosaka, and X. Liu, Exotic triple-charm deuteronlike hexaquarks, Phys. Rev. D 97, 114011 (2018).
  • (63) F. L. Wang, R. Chen, Z. W. Liu, and X. Liu, Possible triple-charm molecular pentaquarks from ΞccD1/ΞccD2\Xi_{cc}D_{1}/\Xi_{cc}D_{2}^{*} interactions, Phys. Rev. D 99, 054021 (2019).
  • (64) F. L. Wang and X. Liu, Exotic double-charm molecular states with hidden or open strangeness and around 4.54.74.5\sim 4.7 GeV, Phys. Rev. D 102, 094006 (2020).
  • (65) G. Breit, The effect of retardation on the interaction of two electrons, Phys. Rev.  34, 553 (1929).
  • (66) G. Breit, The fine structure of HE as a test of the spin interactions of two electrons, Phys. Rev.  36, 383 (1930).
  • (67) N. A. Tornqvist, From the deuteron to deusons, an analysis of deuteronlike meson-meson bound states, Z. Phys. C 61, 525 (1994).
  • (68) N. A. Tornqvist, On deusons or deuteron-like meson-meson bound states, Nuovo Cimento Soc. Ital. Fis. 107A, 2471 (1994).
  • (69) P. A. Zyla et al. (Particle Data Group), Review of particle physics, PTEP 2020, no.8, 083C01 (2020).
  • (70) R. Chen, A. Hosaka, and X. Liu, Prediction of triple-charm molecular pentaquarks, Phys. Rev. D 96, 114030 (2017).
  • (71) C. Y. Wong, E. S. Swanson, and T. Barnes, Heavy quarkonium dissociation cross-sections in relativistic heavy ion collisions, Phys. Rev. C 65, 014903 (2002) Erratum: [Phys. Rev. C 66, 029901 (2002)].
  • (72) S. Godfrey and N. Isgur, Mesons in a relativized quark model with chromodynamics, Phys. Rev. D 32, 189 (1985).
  • (73) S. Weinberg, Elementary particle theory of composite particles, Phys. Rev.  130, 776 (1963).
  • (74) S. Weinberg, Quasiparticles and the Born series, Phys. Rev.  131, 440 (1963).
  • (75) Y. H. Lin, C. W. Shen, F. K. Guo, and B. S. Zou, Decay behaviors of the PcP_{c} hadronic molecules, Phys. Rev. D 95, 114017 (2017).
  • (76) Y. H. Lin, C. W. Shen, and B. S. Zou, Decay behavior of the strange and beauty partners of PcP_{c} hadronic molecules, Nucl. Phys. A980, 21-31 (2018).
  • (77) Y. H. Lin and B. S. Zou, Hadronic molecular assignment for the newly observed Ω\Omega^{*} state, Phys. Rev. D 98, 056013 (2018).
  • (78) C. W. Shen, J. J. Wu, and B. S. Zou, Decay behaviors of possible Λcc¯\Lambda_{c\bar{c}} states in hadronic molecule pictures, Phys. Rev. D 100, 056006 (2019).
  • (79) Y. H. Lin and B. S. Zou, Strong decays of the latest LHCb pentaquark candidates in hadronic molecule pictures, Phys. Rev. D 100, 056005 (2019).
  • (80) Y. H. Lin, F. Wang, and B. S. Zou, Reanalysis of the newly observed Ω\Omega^{*} state in a hadronic molecule model, Phys. Rev. D 102, 074025 (2020).
  • (81) X. K. Dong, Y. H. Lin, and B. S. Zou, Prediction of an exotic state around 4240 MeV with JPC=1+J^{PC}=1^{-+} as CC-parity partner of Y(4260)Y(4260) in molecular picture, Phys. Rev. D 101, 076003 (2020).
  • (82) X. K. Dong and B. S. Zou, Prediction of possible DK1DK_{1} bound states, arXiv:2009.11619.
  • (83) D. Y. Chen, C. J. Xiao, and J. He, Hidden-charm decays of Y(4390)Y(4390) in a hadronic molecular scenario, Phys. Rev. D 96, 054017 (2017).
  • (84) C. J. Xiao and D. Y. Chen, Possible B()K¯B^{(\ast)}\bar{K} hadronic molecule state, Eur. Phys. J. A 53, 127 (2017).
  • (85) C. J. Xiao, Y. Huang, Y. B. Dong, L. S. Geng, and D. Y. Chen, Exploring the molecular scenario of Pc(4312)P_{c}(4312) , Pc(4440)P_{c}(4440), and Pc(4457)P_{c}(4457), Phys. Rev. D 100, 014022 (2019).
  • (86) Q. Wu, D. Y. Chen, and F. K. Guo, Production of the Zb()Z_{b}^{(\prime)} states from the Υ(5S,6S)\Upsilon(5S,6S) decays, Phys. Rev. D 99, 034022 (2019).