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

Open charm and bottom meson-nucleon potentials à la the nuclear force

Yasuhiro Yamaguchi [email protected] Department of Physics, Nagoya University, Nagoya 464-8602, Japan Advanced Science Research Center, Japan Atomic Energy Agency (JAEA), Tokai 319-1195, Japan    Shigehiro Yasui [email protected] Research and Education Center for Natural Sciences, Keio University, Hiyoshi 4-1-1, Yokohama, Kanagawa 223-8521, Japan    Atsushi Hosaka [email protected] Research Center for Nuclear Physics (RCNP), Ibaraki, Osaka 567-0047, Japan Advanced Science Research Center, Japan Atomic Energy Agency, Tokai, Ibaraki 319-1195, Japan Theoretical Research Division, Nishina Center, RIKEN, Hirosawa, Wako, Saitama 351-0198, Japan
Abstract

We discuss the interaction of an open heavy meson (D¯\bar{D} and D¯\bar{D}^{\ast} for charm or BB and BB^{\ast} for bottom) and a nucleon (NN) by considering the π\pi, σ\sigma, ρ\rho, and ω\omega exchange potentials. We construct a potential model by respecting chiral symmetry for light quarks and spin symmetry for heavy quarks. Model parameters are adjusted by referring the phenomenological nuclear (CD-Bonn) potentials reproducing the low-energy NNNN scatterings. We show that the resulting interaction may accommodate D¯N\bar{D}N and BNBN bound states with quantum numbers I(JP)=0(1/2)I(J^{P})=0(1/2^{-}), and 1(1/2)1(1/2^{-}). We find that, in the present potential model, the π\pi exchange potential plays an important role for the isosinglet channel, while the σ\sigma exchange potential does for the isotriplet one.

preprint:

I Introduction

Studies of exotic hadrons, such as XX, YY, ZZ, PcP_{c}, XccX_{cc}, TccT_{cc}, and so on, have revealed novel properties of multiquark systems with heavy flavors of charm and bottom Swanson (2006); Voloshin (2008); Brambilla et al. (2011, 2014); Chen et al. (2016); Hosaka et al. (2016); Lebed et al. (2017); Esposito et al. (2017); Ali et al. (2017); Olsen et al. (2018); Guo et al. (2018); Liu et al. (2019); Brambilla et al. (2020); Yamaguchi et al. (2020); Chen et al. (2022). One of the most important problems in exotic hadrons is the inter-hadron interactions. In the present paper, we focus on the interaction between a nucleon NN and an open-heavy meson, a D¯\bar{D} (D¯\bar{D}^{\ast}) meson or a BB (BB^{\ast}) meson, which is intimately relevant to the formation of pentaquarks. Such an interaction is also relevant for heavy-flavored exotic nuclei as bound states formed by a multiple number of baryons Hosaka et al. (2017). Recently the ALICE collaboration in LHCb has reported the first experimental study of the D¯N\bar{D}N interaction which was measured through the correlation functions from proton-proton collisions Acharya et al. (2022). Further development of studying the interaction between a nucleon NN and an open-heavy meson should be awaited.

One of the efficient theoretical analyses can be performed systematically with the basis on the heavy-quark effective theory. This is an effective theory of QCD, where a charm (bottom) quark is approximately regarded as a particle with an infinitely heavy mass mQm_{Q}\rightarrow\infty. In this limit, there appears the heavy-quark spin (HQS) symmetry, i.e., the SU(2) spin symmetry, as in the nonrelativistic limit. This symmetry stems from the decoupling of the heavy quark from light degrees of freedom with the suppressed magnetic interaction, i.e., the spin-flip interaction. The HQS symmetry puts conditions on the spin structure of interaction vertices not only in the quark-gluon dynamics but also in the hadron dynamics.

The HQS symmetry is seen in the observed approximate degeneracy in masses of D¯\bar{D} and D¯\bar{D}^{\ast} (BB and BB^{\ast}) mesons. Also, the HQS symmetry constrains the structure of the inter-hadron interaction in the channel-coupled D¯N\bar{D}N and D¯N\bar{D}^{\ast}N (BNBN and BNB^{\ast}N) systems. For example, it was shown that the approximate degeneracy in D¯\bar{D} and D¯\bar{D}^{\ast} mesons increases the attractive interaction strength between a nucleon and a D¯\bar{D} meson through the box diagram D¯ND¯ND¯N\bar{D}N\rightarrow\bar{D}^{\ast}N\rightarrow\bar{D}N in the second-order perturbative process Haidenbauer et al. (2007). This mechanism is different from the conventional approach based on the SU(4) flavor symmetry Lutz and Korpa (2006); Hofmann and Lutz (2005) and the quark-meson coupling model Haidenbauer et al. (2007); Fontoura et al. (2013); Carames and Valcarce (2012); Caramés and Valcarce (2017). The role of the HQS symmetry is shown to be important by including all the coupled channels of D¯N\bar{D}N and D¯N\bar{D}^{\ast}N (BNBN and BNB^{\ast}N). Hereafter we will introduce the short notations PP and PP^{\ast} corresponding to D¯\bar{D} and D¯\bar{D}^{\ast} (BB and BB^{\ast}), respectively. We employ P()P^{(\ast)} to denote either PP or PP^{\ast}. In such a framework, we consider the coupled channels of PNPN and PNP^{\ast}N and study the interaction between a P()P^{(\ast)} meson and a nucleon, denoted by PNPN-PNP^{\ast}N.

In the literature, the PNPN-PNP^{\ast}N interactions were introduced by the one-pion exchange potential (OPEP) with the constraint conditions induced by the HQS symmetry Cohen et al. (2005); Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012); Yasui et al. (2013); Yamaguchi et al. (2015). The analysis of the PNPN-PNP^{\ast}N systems showed the possible existence of composite states: bound states below the D¯N\bar{D}N (BNBN) threshold Cohen et al. (2005); Yasui and Sudoh (2009), and Feshbach resonant states in the continuum region slightly below the D¯N\bar{D}^{\ast}N (BNB^{\ast}N) threshold Yamaguchi et al. (2011, 2012). In the heavy quark limit, these two states are regarded as the doublet states in the mixed bases by PNPN and PNP^{\ast}N in terms of the HQS symmetry. Such HQS multiplets have been studied for negative parity Yasui and Sudoh (2009); Yamaguchi et al. (2011) and positive parity Yamaguchi et al. (2012) states.

In HQS symmetry, it is important to realize that the PNPN-PNP^{\ast}N interactions can be provided by the interaction between the nucleon and the light quark qq in P()P^{(\ast)}. Thus, the PNPN-PNP^{\ast}N interactions can be regarded effectively as the “qNqN” interaction. This would be a generalization of the conventional nuclear force (NNNN) to the force between a light quark and a nucleon with different baryon numbers. Such an idea enables us to construct the PNPN-PNP^{\ast}N potential from the qNqN potential, with reference to the NNNN potential in detail. It was shown that the PNPN-PNP^{\ast}N interaction can be expressed by the qNqN interaction by applying the unitary transformations Yasui et al. (2013); Yamaguchi et al. (2015).

In the present work, we reconstruct the PNPN-PNP^{\ast}N potential, where we refer to the phenomenological nuclear potential, the CD-Bonn potential Machleidt (2001). In the framework of the CD-Bonn potential, the nuclear force is described by the π\pi, ρ\rho, ω\omega, and σ\sigma exchanges. It is known that the σ\sigma exchange is important to reproduce the phase shifts in NNNN scatterings for isospin singlet and triplet channels simultaneously. In fact, the π\pi, ρ\rho, and ω\omega-exchange potentials are not enough for the fitting to the observed data of NNNN scatterings. In reference to the CD-Bonn potential, thus we also introduce the middle-range force by the σ\sigma exchange potential in addition to the π\pi, ρ\rho, and ω\omega potentials in the PNPN-PNP^{\ast}N interaction which were discussed by the previous studies Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012). As introduced in the CD-Bonn potential, the parameters of the σ\sigma exchange have different values between the isosinglet and isotriplet channels. Considering D¯N\bar{D}N-D¯N\bar{D}^{\ast}N and BNBN-BNB^{\ast}N systems with the reconstructed potentials, we discuss the possible existence of bound states, as discussed in Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012).

The paper is organized as the followings. In Sec. II.1, we introduce the potentials for PNPN and PNP^{\ast}N in terms of the π\pi, σ\sigma, ρ\rho, and ω\omega exchanges. We give an analysis for the σ\sigma exchange potential which is newly introduced in the present study. We present in details the calculation process of the derivation of the potential, because we include some corrections for the potential forms derived in our previous works. In Sec. III, we present the numerical results for the scattering lengths in the PNPN and PNP^{\ast}N potentials and the binding energies for the bound states. The final section is devoted to our conclusion and prospects for future studies.

II Formalism

II.1 Construction of PNPN and PNP^{\ast}N potentials

II.1.1 OPEP

Let us consider the PNPN-PNP^{\ast}N states of JP=1/2J^{P}=1/2^{-} with a total angular momentum JJ and parity PP. PNPN and PNP^{\ast}N components in those states are represented by

PN(2S1/2),PN(2S1/2),PN(4D1/2).\displaystyle PN(^{2}{S}_{1/2}),P^{\ast}N(^{2}{S}_{1/2}),P^{\ast}N(^{4}{D}_{1/2}). (1)

Here the notation LJ2S+1{}^{2S+1}L_{J} in the parentheses stands for the combination of the total spin SS and the relative angular momentum LL for a given JJ. In view of the HQS symmetry, the wave functions given above are decomposed into the product of a heavy antiquark Q¯\bar{Q} and a light component “ll”. Here “ll” is nonperturbatively composed of the light quarks (qq) and gluons (gg) inside the PNPN-PNP^{\ast}N state. Such a light component may be schematically denoted by qqqqqqqq, because it should be a composite state of the light quark qq in PP or PP^{\ast} and the three quarks qqqqqq in the nucleon NN. This is the special case of the so-called brown muck which was introduced in the early days when the heavy quark effective theory (HQET) was constructed.111In the present setting, the brown muck is regarded to have the special component qNqN in qqqqqqqq.

The idea of the light composite state leads to the mass degeneracy of the PNPN-PNP^{\ast}N states with different JPJ^{P}, such as JP=1/2J^{P}=1/2^{-} and 3/23/2^{-} by taking the heavy quark limit, because the spin-dependent interaction between the heavy antiquark (Q¯\bar{Q}) and the brown muck (qqqqqqqq) is suppressed by 1/mQ1/m_{Q} with the heavy quark mass mQm_{Q}. The mass degeneracy of the PNPN-PNP^{\ast}N states have been studied in Refs. Yasui et al. (2013); Yamaguchi et al. (2015); Hosaka et al. (2017).

For the interaction in the PNPN-PNP^{\ast}N systems, we adopt the meson-exchange potential between P()P^{(\ast)} and NN. We consider the one-pion exchange potential (OPEP) as the long-range force. We also consider the σ\sigma-meson exchange potentials and the ρ\rho and ω\omega-meson exchange potentials as the middle-range force.

Let us first explain the derivation of the OPEP in details as an illustration. In constructing the OPEP, we need the information of the interaction vertices of π\pi and P()P^{(\ast)} and those of π\pi and NN. For the πPP\pi PP^{\ast} and πPP\pi P^{\ast}P^{\ast} vertices, we employ the heavy meson effective theory (HMET) satisfying the HQS as well as chiral symmetry Manohar and Wise (2000); Casalbuoni et al. (1997). Notice the absence of the πPP\pi PP vertex due to the parity conservation.

For heavy mesons PP and PP^{\ast}, we define the effective field HaH_{a} being a superposition of a heavy pseudoscalar meson and a vector meson as

Hα=(Pαμγμ+Pαγ5)1/v2,\displaystyle H_{\alpha}=\bigl{(}P_{\alpha}^{\ast\mu}\gamma_{\mu}+P_{\alpha}\gamma_{5}\bigr{)}\frac{1-\ooalign{\hfil/\hfil\crcr$v$}}{2}, (4)

where the subscripts α=±1/2\alpha=\pm 1/2 represent the isospin components (up and down) in the light quark components. PαP_{\alpha} and PαμP_{\alpha}^{\ast\mu} denote the pseudoscalar and vector meson fields, respectively. The relative phase of PαμP_{\alpha}^{\ast\mu} and PαP_{\alpha} is arbitrary, and the present choice is adopted for the convenience in representing the PNPN-PNP^{\ast}N potential as it will be shown later. Here vμv^{\mu} (μ=0,1,2,3\mu=0,1,2,3) is the four velocity of the heavy meson (heavy antiquark) satisfying vμvμ=1v_{\mu}v^{\mu}=1 and v0>0v^{0}>0. We notice that (1/v)/2(1-\ooalign{\hfil/\hfil\crcr$v$})/2 is the operator for projecting out the positive-energy component in the heavy antiquark Q¯\bar{Q} and discarding the negative-energy component. The complex conjugate of HαH_{\alpha} is defined by H¯α=γ0Hαγ0\bar{H}_{\alpha}=\gamma_{0}H_{\alpha}^{\dagger}\gamma_{0}. The effective field HαH_{\alpha} transforms as HαUαβHβSH_{\alpha}\rightarrow U_{\alpha\beta}H_{\beta}S^{\dagger} under the heavy-quark spin and chiral symmetries. Here SSU(2)spinS\in\mathrm{SU}(2)_{\mathrm{spin}} represents the transformation operator for the heavy-quark spin and Uαβ=Uαβ(L,R)U_{\alpha\beta}=U_{\alpha\beta}(L,R) is a function in the nonlinear representation of chiral symmetry with LSU(2)LL\in\mathrm{SU}(2)_{\mathrm{L}} and RSU(2)RR\in\mathrm{SU}(2)_{\mathrm{R}} for light up and down flavors.

In terms of HαH_{\alpha} defined by Eq. (4), the interaction Lagrangian for the πP()P()\pi P^{(\ast)}P^{(\ast)} vertex is given by

πHH=igπtr(HαH¯βγμγ5Aβαμ),\displaystyle{\cal L}_{\pi HH}=ig_{\pi}\mbox{tr}\bigl{(}H_{\alpha}\bar{H}_{\beta}\gamma_{\mu}\gamma_{5}A^{\mu}_{\beta\alpha}\bigr{)}, (5)

where the axial current AβαμA^{\mu}_{\beta\alpha} by pions is defined by Aμ=(ξμξξμξ)/2A^{\mu}=\bigl{(}\xi^{{\dagger}}\partial^{\mu}\xi-\xi\partial^{\mu}\xi^{{\dagger}}\bigr{)}/2 with the nonlinear representation

ξ=exp(i𝝉𝝅2fπ),\displaystyle\xi=\exp\biggl{(}i\frac{\boldsymbol{\tau}\!\cdot\!\boldsymbol{\pi}}{2f_{\pi}}\biggr{)}, (6)

with the pion decay constant fπ=94f_{\pi}=94 MeV. The pion field is defined by 𝝅=(π1,π2,π3)\boldsymbol{\pi}=(\pi_{1},\pi_{2},\pi_{3}) with π±=(π1iπ2)/2\pi^{\pm}=(\pi_{1}\mp i\pi_{2})/\sqrt{2} for charged pions and π0=π3\pi^{0}=\pi_{3} for a neutral pion. Notice that the matrix AμA^{\mu} is transformed by AμUAμUA^{\mu}\rightarrow UA^{\mu}U^{{\dagger}} in the nonlinear representation of chiral symmetry. Thus we confirm that the interaction Lagrangian (5) is invariant under both the HQS and chiral symmetries. The coupling constant gπ=0.59g_{\pi}=0.59 in Eq. (5) is determined from the decay width of DDπ0D^{\ast-}\rightarrow D^{-}\pi^{0} observed by experiments Zyla et al. (2020). We note that gπg_{\pi} is nothing but the quark axial coupling gAqg_{A}^{q} whose value looks smaller than what is naively expected, gAq=1g_{A}^{q}=1 Weinberg (1990). The small value is understood by considering corrections due to quark’s relativistic motion inside hadrons as discussed in detail for baryon decays Arifi et al. (2022). There are uncertainties for choosing the signs of the coupling constants in D¯\bar{D} (D¯\bar{D}^{\ast}) and BB (BB^{\ast}). In the present study, we assume that the π\pi, σ\sigma, ρ\rho, and ω\omega mesons couple to the light constituent quarks in the heavy mesons as well as in the nucleons. In this scheme, we can consider that the signs of these meson couplings for the light quarks in the heavy meson are the same as for the nucleon, because both have the same light (up and down) constituent quarks according to the conventional quark model.

Below we consider the frame in which the heavy meson is at rest and set vμ=(1,𝟎)v^{\mu}=(1,\boldsymbol{0}) in Eq. (5). Thus we obtain the πP()P()\pi P^{(\ast)}P^{(\ast)} vertices:

πPP\displaystyle{\cal L}_{\pi P^{\ast}P^{\ast}} =igπfπενρμσvνPβρ(𝝉μ𝝅)βαPασ,\displaystyle=\frac{ig_{\pi}}{f_{\pi}}\varepsilon_{\nu\rho\mu\sigma}v^{\nu}P_{\beta}^{\ast\rho{\dagger}}\bigl{(}\boldsymbol{\tau}\!\cdot\!\partial^{\mu}\boldsymbol{\pi}\bigr{)}_{\beta\alpha}P_{\alpha}^{\ast\sigma}, (7)
πPP\displaystyle{\cal L}_{\pi P^{\ast}P} =iigπfπPβμ(𝝉μ𝝅)βαPα,\displaystyle=i\frac{ig_{\pi}}{f_{\pi}}P_{\beta\mu}^{\ast{\dagger}}\bigl{(}\boldsymbol{\tau}\!\cdot\!\partial^{\mu}\boldsymbol{\pi}\bigr{)}_{\beta\alpha}P_{\alpha}, (8)
πPP\displaystyle{\cal L}_{\pi PP^{\ast}} =iigπfπPβ(𝝉μ𝝅)βαPαμ.\displaystyle=i\frac{ig_{\pi}}{f_{\pi}}P_{\beta}^{{\dagger}}\bigl{(}\boldsymbol{\tau}\!\cdot\!\partial^{\mu}\boldsymbol{\pi}\bigr{)}_{\beta\alpha}P_{\alpha\mu}^{\ast}. (9)

We introduce the interaction Lagrangian of a pion and a nucleon in the axial-vector coupling

πNN\displaystyle{\cal L}_{\pi NN} =gAN2fπψ¯γμγ5𝝉μ𝝅ψ.\displaystyle=\frac{g_{A}^{N}}{2f_{\pi}}\bar{\psi}\gamma_{\mu}\gamma_{5}\boldsymbol{\tau}\!\cdot\!\partial^{\mu}\boldsymbol{\pi}\psi. (10)

Here ψ=(ψ+1/2,ψ1/2)T\psi=(\psi_{+1/2},\psi_{-1/2})^{T} with the isospin components ψ+1/2\psi_{+1/2} and ψ1/2\psi_{-1/2} for a proton and a neutron, respectively. The value of gANg_{A}^{N} is given by the Goldberger-Treiman relation

gANfπ=gπNNmN,\displaystyle\frac{g_{A}^{N}}{f_{\pi}}=\frac{g_{\pi NN}}{m_{N}}, (11)

and gπNN2/4π=13.6g_{\pi NN}^{2}/4\pi=13.6 from the phenomenological nuclear potential in Ref. Machleidt (2001) (see also Ref. Machleidt et al. (1987)). We adopt the values of the coupling constants and the cutoff parameters by referring the parameters in the CD-Bonn potential. The nuclear potentials used in the present study are explained in Appendix A.

With the interaction vertices (5) and (10), we construct the OPEP between P()P^{(\ast)} and NN Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012). We show the demonstration to derive the potential for the simple model in Appendix B. The OPEP includes three channels: PNPNP^{\ast}N\rightarrow P^{\ast}N, PNPNP^{\ast}N\rightarrow PN, and PNPNPN\rightarrow P^{\ast}N. We notice that the PNPNPN\rightarrow PN process is absent as a direct process due to the prohibition of the πPP\pi PP vertex, and that the PNPN-PNPN interaction is indirectly supplied by multi-step process stemming from the mixing of PNPN and PNP^{\ast}N Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012). The OPEPs for PNP^{\ast}N-PNP^{\ast}N, PNP^{\ast}N-PNPN, and PNPN-PNP^{\ast}N are given by

VπPN-PN(𝒓)\displaystyle V_{\pi}^{P^{\ast}N\text{-}P^{\ast}N}(\boldsymbol{r}) =Gπ(T(r;mπ)(3(𝑻𝒓^)(𝝈𝒓^)𝑻𝝈)+C(r;mπ)𝑻𝝈)𝝉H𝝉N,\displaystyle={G_{\pi}}\Bigl{(}T(r;m_{\pi})\bigl{(}3(\boldsymbol{T}\!\cdot\!\hat{\boldsymbol{r}})(\boldsymbol{\sigma}\!\cdot\!\hat{\boldsymbol{r}})-\boldsymbol{T}\!\cdot\!\boldsymbol{\sigma}\bigr{)}+C(r;m_{\pi})\boldsymbol{T}\!\cdot\!\boldsymbol{\sigma}\Bigr{)}\boldsymbol{\tau}^{H}\!\cdot\!\boldsymbol{\tau}^{N}, (12)
VπPN-PN(𝒓)\displaystyle V_{\pi}^{P^{\ast}N\text{-}PN}(\boldsymbol{r}) =Gπ(T(r;mπ)(3(ϵ𝒓^)(𝝈𝒓^)ϵ𝝈)+C(r;mπ)ϵ𝝈)𝝉H𝝉N,\displaystyle=-{G_{\pi}}\Bigl{(}T(r;m_{\pi})\bigl{(}3(\boldsymbol{\epsilon}^{\ast}\!\cdot\!\hat{\boldsymbol{r}})(\boldsymbol{\sigma}\!\cdot\!\hat{\boldsymbol{r}})-\boldsymbol{\epsilon}^{\ast}\!\cdot\!\boldsymbol{\sigma}\bigr{)}+C(r;m_{\pi})\boldsymbol{\epsilon}^{\ast}\!\cdot\!\boldsymbol{\sigma}\Bigr{)}\boldsymbol{\tau}^{H}\!\cdot\!\boldsymbol{\tau}^{N}, (13)
VπPN-PN(𝒓)\displaystyle V_{\pi}^{PN\text{-}P^{\ast}N}(\boldsymbol{r}) =Gπ(T(r;mπ)(3(ϵ𝒓^)(𝝈𝒓^)ϵ𝝈)+C(r;mπ)ϵ𝝈)𝝉H𝝉N,\displaystyle=-{G_{\pi}}\Bigl{(}T(r;m_{\pi})\bigl{(}3(\boldsymbol{\epsilon}\!\cdot\!\hat{\boldsymbol{r}})(\boldsymbol{\sigma}\!\cdot\!\hat{\boldsymbol{r}})-\boldsymbol{\epsilon}\!\cdot\!\boldsymbol{\sigma}\bigr{)}+C(r;m_{\pi})\boldsymbol{\epsilon}\!\cdot\!\boldsymbol{\sigma}\Bigr{)}\boldsymbol{\tau}^{H}\!\cdot\!\boldsymbol{\tau}^{N}, (14)

with the coefficient

Gπ=1312gπNN2mNgπfπ.\displaystyle{G_{\pi}}=\frac{1}{3}{\frac{1}{2}}\frac{g_{\pi NN}}{2m_{N}}\frac{g_{\pi}}{f_{\pi}}. (15)

We notice that the coefficient 1/21/2 is necessary due to the normalization factor of the wave functions, which was missing in Refs. Yasui et al. (2013); Yamaguchi et al. (2015); Hosaka et al. (2017). The derivation of the OPEP is shown in Appendix C in details. The functions C(r;m)C(r;m) and T(r;m)T(r;m) are defined by

C(r;m)\displaystyle C(r;m) =m24π1r\displaystyle=\frac{m^{2}}{4\pi}\frac{1}{r}
×(emr+ΛH2m2ΛN2ΛH2eΛNr+ΛN2m2ΛH2ΛN2eΛHr),\displaystyle\times\biggl{(}e^{-mr}+\frac{\Lambda_{H}^{2}-m^{2}}{\Lambda_{N}^{2}-\Lambda_{H}^{2}}e^{-\Lambda_{N}r}+\frac{\Lambda_{N}^{2}-m^{2}}{\Lambda_{H}^{2}-\Lambda_{N}^{2}}e^{-\Lambda_{H}r}\biggr{)}, (16)
T(r;m)\displaystyle T(r;m) =14π(m2(1r+3mr2+3m2r3)emr\displaystyle=\frac{1}{4\pi}\Biggl{(}m^{2}\biggl{(}\frac{1}{r}+\frac{3}{mr^{2}}+\frac{3}{m^{2}r^{3}}\biggr{)}e^{-mr}
+ΛN2(1r+3ΛNr2+3ΛN2r3)ΛH2m2ΛN2ΛH2eΛNr\displaystyle\hskip 10.00002pt+\Lambda_{N}^{2}\biggl{(}\frac{1}{r}+\frac{3}{\Lambda_{N}r^{2}}+\frac{3}{\Lambda_{N}^{2}r^{3}}\biggr{)}\frac{\Lambda_{H}^{2}-m^{2}}{\Lambda_{N}^{2}-\Lambda_{H}^{2}}e^{-\Lambda_{N}r}
+ΛH2(1r+3ΛHr2+3ΛH2r3)ΛN2m2ΛH2ΛN2eΛHr),\displaystyle\hskip 10.00002pt+\Lambda_{H}^{2}\biggl{(}\frac{1}{r}+\frac{3}{\Lambda_{H}r^{2}}+\frac{3}{\Lambda_{H}^{2}r^{3}}\biggr{)}\frac{\Lambda_{N}^{2}-m^{2}}{\Lambda_{H}^{2}-\Lambda_{N}^{2}}e^{-\Lambda_{H}r}\Biggr{)}, (17)

with m=mπm=m_{\pi}, respectively, as functions of an interdistance r=|𝒓|r=|\boldsymbol{r}| for 𝒓\boldsymbol{r} being the relative coordinate vector between P()P^{(\ast)} and NN. The detailed information to derive the potentials are presented in Appendix C. Notice that the values of the cutoff parameters ΛH\Lambda_{H} (H=D¯,BH=\bar{D},B) and ΛN\Lambda_{N} are dependent on the species of the exchanged light meson, e.g., the π\pi meson. Originally, C(r,m)C(r,m) and V(r,m)V(r,m) are defined by

C(r;m)=d3𝒒(2π)3m2𝒒  2+m2ei𝒒𝒓F(𝒒;m),\displaystyle C(r;m)=\int\frac{d^{3}\boldsymbol{q}}{(2\pi)^{3}}\frac{m^{2}}{\boldsymbol{q}^{\,\,2}+m^{2}}e^{i\boldsymbol{q}\cdot\boldsymbol{r}}F(\boldsymbol{q};m)\,, (18)
S𝒪(𝒓^)T(r;m)=d3𝒒(2π)3𝒒  2𝒒  2+m2S𝒪(𝒒^)ei𝒒𝒓F(𝒒;m),\displaystyle S_{\cal O}(\hat{\boldsymbol{r}})T(r;m)=\int\frac{d^{3}\boldsymbol{q}}{(2\pi)^{3}}\frac{-\boldsymbol{q}^{\,\,2}}{\boldsymbol{q}^{\,\,2}+m^{2}}S_{\cal O}(\hat{\boldsymbol{q}})e^{i\boldsymbol{q}\cdot\boldsymbol{r}}F(\boldsymbol{q};m), (19)

for the central and tensor parts, respectively, with 𝒒^=𝒒/|𝒒|\hat{\boldsymbol{q}}=\boldsymbol{q}/|\boldsymbol{q}|. We note that the contact term in the central part is neglected. The dipole-type form factor is given by

F(𝒒;m)=ΛH2m2ΛH2+|𝒒|2ΛN2m2ΛN2+|𝒒|2,\displaystyle F(\boldsymbol{q};m)=\frac{\Lambda_{H}^{2}-m^{2}}{\Lambda_{H}^{2}+|\boldsymbol{q}|^{2}}\frac{\Lambda_{N}^{2}-m^{2}}{\Lambda_{N}^{2}+|\boldsymbol{q}|^{2}}, (20)

which is normalized at q2=m2q^{2}=m^{2} with a four-momentum qq. The cutoff parameters ΛH\Lambda_{H} and ΛN\Lambda_{N} would correspond to the inverse of the spatial sizes of hadrons. See the derivations in Appendix C for more details. In Eqs. (13) and (14), we define the polarization vectors ϵ(λ)\boldsymbol{\epsilon}^{\,(\lambda)} (ϵ(λ)\boldsymbol{\epsilon}^{\,(\lambda)*}) for the incoming (outgoing) PP^{\ast} meson with the polarization λ=0,±1\lambda=0,\,\pm 1. The explicit forms of ϵ(λ)\boldsymbol{\epsilon}^{\,(\lambda)} can be represented by

ϵ(±)=12(1,i,0),ϵ(0)=(0,0,1),\displaystyle\boldsymbol{\epsilon}^{\,(\pm)}=\frac{1}{\sqrt{2}}\bigl{(}\mp 1,-i,0\bigr{)},\quad\boldsymbol{\epsilon}^{\,(0)}=(0,0,1), (21)

by choosing the positive direction in the zz axis for the helicity λ=0\lambda=0. As for the spin-one operator for the PP^{\ast} meson in Eq. (12), we define 𝑻=(T1,T2,T3)\boldsymbol{T}=(T_{1},T_{2},T_{3}) by (Ti)λλiεijkϵj(λ)ϵk(λ)(T_{i})_{\lambda^{\prime}\lambda}\equiv-i\varepsilon_{ijk}\epsilon_{j}^{(\lambda^{\prime})\ast}\epsilon_{k}^{(\lambda)} (i,j,k=1,2,3i,j,k=1,2,3):

T1\displaystyle T_{1} =12(010101010),T2=12(0i0i0i0i0),\displaystyle=\frac{1}{\sqrt{2}}\left(\begin{array}[]{ccc}0&1&0\\ 1&0&1\\ 0&1&0\end{array}\right),\quad T_{2}=\frac{1}{\sqrt{2}}\left(\begin{array}[]{ccc}0&-i&0\\ i&0&-i\\ 0&i&0\end{array}\right), (28)
T3\displaystyle T_{3} =(100000001),\displaystyle=\left(\begin{array}[]{ccc}1&0&0\\ 0&0&0\\ 0&0&-1\end{array}\right), (32)

satisfying the commutation relation [Ti,Tj]=iεijkTk[T_{i},T_{j}]=i\varepsilon_{ijk}T_{k} as the generators of the spin symmetry. We define the tensor operators Sϵ(𝒓^)S_{\boldsymbol{\epsilon}}(\hat{\boldsymbol{r}}) and S𝑻(𝒓^)S_{\boldsymbol{T}}(\hat{\boldsymbol{r}}) by S𝓞(𝒓^)=3(𝓞𝒓^)(𝝈𝒓^)𝓞𝝈S_{\boldsymbol{\cal O}}(\hat{\boldsymbol{r}})=3(\boldsymbol{\cal O}\!\cdot\!\hat{\boldsymbol{r}})(\boldsymbol{\sigma}\!\cdot\!\hat{\boldsymbol{r}})-\boldsymbol{\cal O}\!\cdot\!\boldsymbol{\sigma} with 𝒓^=𝒓/r\hat{\boldsymbol{r}}=\boldsymbol{r}/r for 𝓞=ϵ\boldsymbol{{\cal O}}=\boldsymbol{\epsilon} and 𝑻\boldsymbol{T}. Here 𝝈\boldsymbol{\sigma} are the Pauli matrices acting on the nucleon spin, and 𝝉β1α1H\boldsymbol{\tau}^{H}_{\beta_{1}\alpha_{1}} and 𝝉β2α2N\boldsymbol{\tau}^{N}_{\beta_{2}\alpha_{2}} with αi,βi=±1/2\alpha_{i},\beta_{i}=\pm 1/2 are the isospin Pauli operators for P()P^{(\ast)} (i=1i=1) and NN (i=2i=2), respectively.

Using the basis of the JP=1/2J^{P}=1/2^{-} channel in Eq. (1), we represent the OPEPs (12), (13), and (14) by the matrix forms,

V1/2π\displaystyle V_{1/2^{-}}^{\pi} =(03Cπ6Tπ3Cπ2Cπ2Tπ6Tπ2TπCπ2Tπ),\displaystyle=\left(\begin{array}[]{ccc}0&\sqrt{3}\,{C}_{\pi}&-\sqrt{6}\,{T}_{\pi}\\ \sqrt{3}\,{C}_{\pi}&-2\,{C}_{\pi}&-\sqrt{2}\,{T}_{\pi}\\ -\sqrt{6}\,{T}_{\pi}&-\sqrt{2}\,{T}_{\pi}&{C}_{\pi}-2\,{T}_{\pi}\end{array}\right), (36)

where we define Cπ=GπC(r;mπ){C}_{\pi}=G_{\pi}C(r;m_{\pi}) and Tπ=GπT(r;mπ){T}_{\pi}=G_{\pi}T(r;m_{\pi}) for short notations. In Eq. (36), we confirm that the mixing between PNPN and PNP^{\ast}N are represented by the off-diagonal parts including the tensor potentials. These tensor potentials induce the strong mixing by different angular momenta, leading to the strong attractions at short-range scales. Thus, the mixing of PNPN and PNP^{\ast}N is important to switch on the strong attraction. This is analogous to the OPEP in the nucleon-nucleon interaction.

II.1.2 σ\sigma exchange potential

The interaction Lagrangian for a σ\sigma meson and a P()P^{(\ast)} meson is given by

σIHH=\displaystyle{\cal L}_{\sigma_{I}HH}=\, gσItr(H¯σIH),\displaystyle-g_{\sigma_{I}}\mathrm{tr}\bigl{(}\bar{H}\sigma_{I}H\bigr{)}, (37)

which leads to the σP()P()\sigma P^{(\ast)}P^{(\ast)} vertices,

σIPP=\displaystyle{\cal L}_{\sigma_{I}PP}=\, 2gσI(PσIP),\displaystyle 2g_{\sigma_{I}}\left(P^{\dagger}\sigma_{I}P\right), (38)
σIPP=\displaystyle{\cal L}_{\sigma_{I}P^{\ast}P^{\ast}}=\, 2gσI(PμσIPμ).\displaystyle-2g_{\sigma_{I}}\left(P^{\ast\mu\dagger}\sigma_{I}P^{\ast}_{\mu}\right). (39)

Here we introduce the channel-dependent σI\sigma_{I} meson for isospin-singlet (I=0)(I=0) and isospin-triplet (I=1)(I=1) channels for the PNPN-PNP^{\ast}N scatterings, as introduced in the CD-Bonn potential Machleidt (2001). The parameter of the σ\sigma exchange potential in the CD-Bonn potential Machleidt (2001) has the different value for each partial waves, i.e., isospin channels. Thus, σI\sigma_{I} in the present work also has an channel-dependent mass (mσIm_{\sigma_{I}}), coupling constant (gσIg_{\sigma_{I}}), and cutoff parameter (ΛσI\Lambda_{\sigma_{I}}). Using the σNN\sigma NN vertices given by

σINN\displaystyle{\cal L}_{\sigma_{I}NN} =gσINNψ¯σIψ,\displaystyle=g_{\sigma_{I}NN}\bar{\psi}\sigma_{I}\psi, (40)

we find that the σ\sigma potentials for PNPN and PNP^{\ast}N are obtained by

VσIPN-PN(r)\displaystyle V_{\sigma_{I}}^{PN\text{-}PN}(r) =gσINNgσImσI2C(r;mσI),\displaystyle=-\frac{g_{\sigma_{I}NN}g_{\sigma_{I}}}{m_{\sigma_{I}}^{2}}C(r;m_{\sigma_{I}}), (41)
VσIPN-PN(r)\displaystyle V_{\sigma_{I}}^{P^{\ast}N\text{-}P^{\ast}N}(r) =gσINNgσImσI2C(r;mσI),\displaystyle=-\frac{g_{\sigma_{I}NN}g_{\sigma_{I}}}{m_{\sigma_{I}}^{2}}C(r;m_{\sigma_{I}}), (42)

where we employ the values of mσIm_{\sigma_{I}} and gσINNg_{\sigma_{I}NN} in the CD-Bonn potential, see Appendix A. Concerning the values of gσIg_{\sigma_{I}}, we choose gσI=gσINN/3g_{\sigma_{I}}=g_{\sigma_{I}NN}/3 by assuming that the coupling of a σ\sigma meson and a hadron h=P()h=P^{(\ast)}, NN is proportional to the number of the light quarks in the hadron hh: one light quark in P()P^{(\ast)} and three light quarks in NN. The σ\sigma-exchange potentials are expressed explicitly by

V1/2σI\displaystyle V_{1/2^{-}}^{\sigma_{I}} =(CσI000CσI000CσI),\displaystyle=\left(\begin{array}[]{ccc}C_{\sigma_{I}}&0&0\\ 0&C_{\sigma_{I}}&0\\ 0&0&C_{\sigma_{I}}\end{array}\right), (46)

for the basis by Eq. (1), where we define the function

CσI=gσINNgσImσI2C(r;mσI)\displaystyle C_{\sigma_{I}}=-\frac{g_{\sigma_{I}NN}g_{\sigma_{I}}}{m_{\sigma_{I}}^{2}}C(r;m_{\sigma_{I}}) (47)

for short notations.

II.1.3 ρ\rho and ω\omega exchanges potential

Finally, we consider the exchange of the vector mesons, ρ\rho and ω\omega, at shorter range. The ρ\rho and ω\omega potentials can be constructed from the vP()P()vP^{(\ast)}P^{(\ast)} vertices for light vector meson vv (v=ρv=\rho, ω\omega). Following the previous papers Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012), we consider the interaction Lagrangian

vHH=\displaystyle{\cal L}_{vHH}=\, iβtr(H¯βvμ(ρμ)βαHα)\displaystyle i\beta\mathrm{tr}\bigl{(}\bar{H}_{\beta}v^{\mu}(\rho_{\mu})_{\beta\alpha}H_{\alpha}\bigr{)}
+iλtr(H¯βσμν(Fμν(ρ))βαHα),\displaystyle+i\lambda\mathrm{tr}\bigl{(}\bar{H}_{\beta}\sigma^{\mu\nu}(F_{\mu\nu}(\rho))_{\beta\alpha}H_{\alpha}\bigr{)}, (48)

by respecting the HQS symmetry. The vector meson field is defined by ρμ=igVρ^μ/2\rho_{\mu}=ig_{V}\hat{\rho}_{\mu}/\sqrt{2} with ρ^μ\hat{\rho}_{\mu},

ρ^μ=(ρ02+ω2ρ+ρρ02+ω2)μ,\displaystyle\hat{\rho}_{\mu}=\begin{pmatrix}\frac{\rho^{0}}{\sqrt{2}}+\frac{\omega}{\sqrt{2}}&\rho^{+}\\ \rho^{-}&-\frac{\rho^{0}}{\sqrt{2}}+\frac{\omega}{\sqrt{2}}\end{pmatrix}_{\mu}, (49)

and gV5.8g_{V}\simeq 5.8 the universal vector-meson coupling. In Eq. (48), the tensor field is given by Fμν(ρ)=μρννρμ+[ρμ,ρν]F_{\mu\nu}(\rho)=\partial_{\mu}\rho_{\nu}-\partial_{\nu}\rho_{\mu}+[\rho_{\mu},\rho_{\nu}]. The coupling constants are given by β=0.9\beta=0.9 and λ=0.56\lambda=0.56 GeV-1 by following Refs. Casalbuoni et al. (1997); Isola et al. (2003). In Ref. Isola et al. (2003), β\beta was determined by the vector-meson dominance, and λ\lambda was evaluated by the long distance charming penguin diagrams in the BB meson decay process. The vP()P()vP^{(\ast)}P^{(\ast)} vertices are obtained by the Lagrangians (48) as

vPP=\displaystyle{\cal L}_{vP^{\ast}P^{\ast}}=\, βgVvμPβν(𝝉𝝆μ)βαPαν\displaystyle-\beta g_{V}v_{\mu}P^{\ast\dagger}_{\beta\nu}(\boldsymbol{\tau}\cdot\boldsymbol{\rho}\,^{\mu})_{\beta\alpha}P^{\ast\nu}_{\alpha}
+2iλgV(Pβν(𝝉μ𝝆ν)βαPαμ\displaystyle+2i\lambda g_{V}\left(P^{\ast\nu\dagger}_{\beta}(\boldsymbol{\tau}\cdot\partial_{\mu}\boldsymbol{\rho}_{\nu})_{\beta\alpha}P^{\ast\mu}_{\alpha}\right.
Pβμ(𝝉μ𝝆ν)βαPαν),\displaystyle\left.\quad-P^{\ast\mu\dagger}_{\beta}(\boldsymbol{\tau}\cdot\partial_{\mu}\boldsymbol{\rho}_{\nu})_{\beta\alpha}P^{\ast\nu}_{\alpha}\right), (50)
vPP=\displaystyle{\cal L}_{vP^{\ast}P}=\, 2λgVϵσρμνvσPβρ(𝝉μ𝝆ν)βαPα,\displaystyle 2\lambda g_{V}\epsilon_{\sigma\rho\mu\nu}v^{\sigma}P^{\ast\rho\dagger}_{\beta}(\boldsymbol{\tau}\cdot\partial^{\mu}\boldsymbol{\rho}^{\nu})_{\beta\alpha}P_{\alpha}, (51)
vPP=\displaystyle{\cal L}_{vPP^{\ast}}=\, 2λgVϵσρμνvσPβ(𝝉μ𝝆ν)βαPαρ,\displaystyle 2\lambda g_{V}\epsilon_{\sigma\rho\mu\nu}v^{\sigma}P_{\beta}^{\dagger}(\boldsymbol{\tau}\cdot\partial^{\mu}\boldsymbol{\rho}^{\nu})_{\beta\alpha}P^{\ast\rho}_{\alpha}, (52)
vPP=\displaystyle{\cal L}_{vPP}=\, βgVvμPβ(𝝉𝝆μ)βαPα.\displaystyle\beta g_{V}v_{\mu}P^{\dagger}_{\beta}(\boldsymbol{\tau}\cdot\boldsymbol{\rho}^{\mu})_{\beta\alpha}P_{\alpha}. (53)

For the vNNvNN vertex, we use the interaction Lagrangian

vNN\displaystyle{\cal L}_{vNN} =gρNNψ¯γμ𝝉𝝆μψ+fρNN2mNψ¯σμν𝝉μ𝝆νψ\displaystyle=g_{\rho NN}\bar{\psi}\gamma_{\mu}\boldsymbol{\tau}\!\cdot\!\boldsymbol{\rho}^{\,\mu}\psi+\frac{f_{\rho NN}}{2m_{N}}\bar{\psi}\sigma_{\mu\nu}\boldsymbol{\tau}\!\cdot\!\partial^{\mu}\boldsymbol{\rho}^{\,\nu}\psi
+gωNNψ¯γμωμψ+fωNN2mNψ¯σμνμωνψ,\displaystyle\quad+g_{\omega NN}\bar{\psi}\gamma_{\mu}\omega^{\mu}\psi+\frac{f_{\omega NN}}{2m_{N}}\bar{\psi}\sigma_{\mu\nu}\partial^{\mu}\omega^{\nu}\psi, (54)

for 𝝆μ=(ρ1μ,ρ2μ,ρ3μ)\boldsymbol{\rho}^{\,\mu}=(\rho^{\mu}_{1},\rho^{\mu}_{2},\rho^{\mu}_{3}) with ρ±μ=(ρ1μiρ2μ)/2\rho_{\pm}^{\mu}=(\rho_{1}^{\mu}\mp i\rho_{2}^{\mu})/\sqrt{2} and ρ0μ=ρ3μ\rho_{0}^{\mu}=\rho_{3}^{\mu}. The coupling constants are given by gρNN2/4π=0.84g_{\rho NN}^{2}/4\pi=0.84, gωNN2/4π=20.0g_{\omega NN}^{2}/4\pi=20.0, fρNN/gρNN=6.1f_{\rho NN}/g_{\rho NN}=6.1, and fωNN/gωNN=0.0f_{\omega NN}/g_{\omega NN}=0.0 Machleidt (2001) (see also Ref. Machleidt et al. (1987)). We leave a comment that the coupling strengths in Eqs. (48) and (54) reflect the number of constituent quarks inside the hadrons. This can be easily checked by the nonrelativistic quark model. We should notice, however, that the tensor parts, λ\lambda and fvNNf_{vNN} (v=ρv=\rho, ω\omega), could be different by some factors from the naive expectations, which would be understood from the composite structures of the constituent quarks.

From Eqs. (48) and (54), the one-boson exchange potentials are obtained as

V1/2v\displaystyle V_{1/2^{-}}^{v} =(Cv23Cv6Tv23CvCv4Cv2Tv6Tv2TvCv+2Cv+2Tv),\displaystyle=\begin{pmatrix}C_{v}^{\prime}&2\sqrt{3}C_{v}&\sqrt{6}T_{v}\\ 2\sqrt{3}C_{v}&C_{v}^{\prime}-4C_{v}&\sqrt{2}T_{v}\\ \sqrt{6}T_{v}&\sqrt{2}T_{v}&C_{v}^{\prime}+2C_{v}+2T_{v}\end{pmatrix}, (55)

with v=ρv=\rho, ω\omega for the 1/21/2^{-} state in Eq. (1). The functions CvC_{v}^{\prime}, CvC_{v}, and TvT_{v} are defined by

Cρ\displaystyle C_{\rho}^{\prime} =gVgρNNβ2mρ2C(r;mρ)𝝉H𝝉N,\displaystyle=\frac{g_{V}g_{\rho NN}\beta}{2m_{\rho}^{2}}C(r;m_{\rho})\boldsymbol{\tau}^{H}\!\cdot\!\boldsymbol{\tau}^{N}, (56)
Cρ\displaystyle C_{\rho} =gV(gρNN+fρNN)λ2mN13T(r;mρ)𝝉H𝝉N,\displaystyle=\frac{g_{V}(g_{\rho NN}+f_{\rho NN})\lambda}{2m_{N}}\frac{1}{3}T(r;m_{\rho})\boldsymbol{\tau}^{H}\!\cdot\!\boldsymbol{\tau}^{N}, (57)
Tρ\displaystyle T_{\rho} =gV(gρNN+fρNN)λ2mN13T(r;mρ)𝝉H𝝉N,\displaystyle=\frac{g_{V}(g_{\rho NN}+f_{\rho NN})\lambda}{2m_{N}}\frac{1}{3}T(r;m_{\rho})\boldsymbol{\tau}^{H}\!\cdot\!\boldsymbol{\tau}^{N}, (58)
Cω\displaystyle C_{\omega}^{\prime} =gVgωNNβ2mω2C(r;mω),\displaystyle=\frac{g_{V}g_{\omega NN}\beta}{2m_{\omega}^{2}}C(r;m_{\omega}), (59)
Cω\displaystyle C_{\omega} =gV(gωNN+fωNN)λ2mN13C(r;mω),\displaystyle=\frac{g_{V}(g_{\omega NN}+f_{\omega NN})\lambda}{2m_{N}}\frac{1}{3}C(r;m_{\omega}), (60)
Tω\displaystyle T_{\omega} =gV(gωNN+fωNN)λ2mN13T(r;mω),\displaystyle=\frac{g_{V}(g_{\omega NN}+f_{\omega NN})\lambda}{2m_{N}}\frac{1}{3}T(r;m_{\omega}), (61)

with 𝝉H\boldsymbol{\tau}^{H} and 𝝉N\boldsymbol{\tau}^{N} being the abbreviations of 𝝉β1α1H\boldsymbol{\tau}^{H}_{\beta_{1}\alpha_{1}} and 𝝉β2α2N\boldsymbol{\tau}^{N}_{\beta_{2}\alpha_{2}} for the isospin Pauli operators acting on P()P^{(\ast)} and NN, respectively.

II.2 Total Hamiltonian

The total Hamiltonian for the P()NP^{(\ast)}N states is given as a sum of the kinetic term and the π\pi, σ\sigma, ρ\rho, and ω\omega potentials as

HIJP=KJP+VJPπ+VJPσI+VJPρ+VJPω.\displaystyle H_{IJ^{P}}=K_{J^{P}}+V^{\pi}_{J^{P}}+V^{\sigma_{I}}_{J^{P}}+V^{\rho}_{J^{P}}+V^{\omega}_{J^{P}}. (62)

Here KJPK_{J^{P}} is the diagonal matrix for the kinetic terms given by

K1/2\displaystyle K_{1/2^{-}} =diag(K0,K0,K2),\displaystyle=\mathrm{diag}\bigl{(}K_{0},K_{0}^{\ast},K_{2}^{\ast}\bigr{)}, (63)

where each component is defined by

KL\displaystyle K_{L} =12μ(2r2+2rrL(L+1)r2),\displaystyle=-\frac{1}{2\mu}\left(\frac{\partial^{2}}{\partial r^{2}}+\frac{2}{r}\frac{\partial}{\partial r}-\frac{L(L+1)}{r^{2}}\right), (64)
KL\displaystyle K_{L}^{\ast} =12μ(2r2+2rrL(L+1)r2),\displaystyle=-\frac{1}{2\mu^{\ast}}\left(\frac{\partial^{2}}{\partial r^{2}}+\frac{2}{r}\frac{\partial}{\partial r}-\frac{L(L+1)}{r^{2}}\right), (65)

for angular momenta L=0L=0 and L=2L=2. The reduced masses μ=mNmP/(mN+mP)\mu=m_{N}m_{P}/(m_{N}+m_{P}) and μ=mNmP/(mN+mP)\mu^{\ast}=m_{N}m_{P^{\ast}}/(m_{N}+m_{P^{\ast}}) are defined with mPm_{P} and mPm_{P^{\ast}} being the masses of PP and PP^{\ast} mesons, respectively.

Concerning the cutoff parameters in the potentials, we consider ΛH\Lambda_{H} in Eq. (20) to be expressed by ΛH=κHNΛN\Lambda_{H}=\kappa_{HN}\Lambda_{N} where κHN\kappa_{HN} is the ratio stemming from inverse hadron size. In Refs. Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012), we obtained κD¯N=1.35\kappa_{\bar{D}N}=1.35 for the D¯()N\bar{D}^{(\ast)}N potential and κBN=1.29\kappa_{BN}=1.29 for the B()NB^{(\ast)}N potential. The same ratios were adopted for the ρ\rho and ω\omega exchange potentials, and can be applied also to the σ\sigma exchange potential. In the present study, however, we regard κHN\kappa_{HN} as a free parameter in order to investigate the dependence of the results on the choice of κHN\kappa_{HN} within a range around κD¯N=1.35\kappa_{\bar{D}N}=1.35 and κBN=1.29\kappa_{BN}=1.29. The value of ΛN\Lambda_{N} is determined by modifying the cutoffs in the CD-Bonn potential by another scale parameter κI\kappa_{I} (I=0,1)(I=0,1) for each isospin channels. The scale parameter is determined by reproducing the scattering lengths of the NNNN scatterings for I=1I=1 and the binding energy of a deuteron for I=0I=0, where we employ the simplified nuclear potential neglecting the massive scalar meson, non-local effects and so on in the CD-Bonn potential, see Appendix A in details. The obtained cutoffs are summarized in Table 1.

Table 1: Parameters of the meson exchange potentials. The meson masses are given as the isospin-averaged values. gπg_{\pi}, β\beta, λ\lambda, and gσIg_{\sigma_{I}} are the coupling constants of heavy mesons (see text in details), while gαNNg_{\alpha NN} and fαNNf_{\alpha NN} are those of a nucleon taken from the CD-Bonn potential Machleidt (2001). The cutoffs ΛD¯\Lambda_{\bar{D}} and ΛB\Lambda_{B} are shown as typical values for ΛD¯=1.35ΛN\Lambda_{\bar{D}}=1.35\Lambda_{N} and ΛB=1.29ΛN\Lambda_{B}=1.29\Lambda_{N}, where ΛN\Lambda_{N} is the nucleon cutoff which is scaled by the parameter κI\kappa_{I} (κ0=0.804\kappa_{0}=0.804 and κ1=0.773\kappa_{1}=0.773) from the CD-Bonn potential (see Appendix A in details).
Mesons (α)(\alpha) Masses [MeV] gπg_{\pi} β\beta λ\lambda [GeV-1] gσIg_{\sigma_{I}} gαNN24π\frac{g^{2}_{\alpha NN}}{4\pi} fαNNgαNN\frac{f_{\alpha NN}}{g_{\alpha NN}} ΛD¯\Lambda_{\bar{D}} [MeV] ΛB\Lambda_{B} [MeV] ΛN\Lambda_{N} [MeV]
I=0I=0 I=1I=1 I=0I=0 I=1I=1 I=0I=0 I=1I=1
π\pi 138.04 0.59 13.6 1868 1795 1785 1715 1384 1330
ρ\rho 769.68 0.9 0.56 0.84 6.1 1359 1306 1423 1367 1054 1013
ω\omega 781.94 0.9 0.56 20 0.0 1629 1565 1557 1496 1207 1159
σ0\sigma_{0} 350 0.849406 0.51673 2715 2594 2011
σ1\sigma_{1} 452 2.35276 3.96451 2609 2493 1932

III Numerical results

(a) D¯N(I=0)\bar{D}N\,(I=0) (b) D¯N(I=1)\bar{D}N\,(I=1)
Refer to caption Refer to caption
(c) BN(I=0)BN\,(I=0) (d) BN(I=1)BN\,(I=1)
Refer to caption Refer to caption
Figure 1: The phase shifts of D¯N\bar{D}N [(a) and (b)] and BNBN [(c) and (d)] as functions of the scattering energy. Panels (a) and (c) are for I=0I=0, and panels (b) and (d) are for I=1I=1.
Table 2: Binding energies (B.E.) and mixing ratios of the D¯()N\bar{D}^{(\ast)}N and B()NB^{(\ast)}N states with I(JP)I(J^{P}) quantum numbers. The binding energies are measured from the mass thresholds of D¯N\bar{D}N or BNBN.
D¯N\bar{D}N B.E. [MeV]         Mixing ratio [%]
0(1/2)0(1/2^{-}) 1.381.38
D¯N(2S1/2)\bar{D}N(^{2}{S}_{1/2}) 96.196.1
D¯N(2S1/2)\bar{D}^{\ast}N(^{2}{S}_{1/2}) 1.941.94
D¯N(4D1/2)\bar{D}^{\ast}N(^{4}{D}_{1/2}) 1.931.93
1(1/2)1(1/2^{-}) 5.995.99
D¯N(2S1/2)\bar{D}N(^{2}S_{1/2}) 88.9
D¯N(2S1/2)\bar{D}^{\ast}N(^{2}S_{1/2}) 10.9
D¯N(4D1/2)\bar{D}^{\ast}N(^{4}D_{1/2}) 0.11
BNBN B.E. [MeV]         Mixing ratio [%]
0(1/2)0(1/2^{-}) 29.729.7
BN(2S1/2)BN(^{2}{S}_{1/2}) 76.476.4
BN(2S1/2)B^{\ast}N(^{2}{S}_{1/2}) 14.114.1
BN(4D1/2)B^{\ast}N(^{4}{D}_{1/2}) 9.469.46
1(1/2)1(1/2^{-}) 66.0
BN(2S1/2)BN(^{2}{S}_{1/2}) 38.538.5
BN(2S1/2)B^{\ast}N(^{2}{S}_{1/2}) 61.561.5
BN(4D1/2)B^{\ast}N(^{4}{D}_{1/2}) 1.82×1021.82\times 10^{-2}

First let us show the phase shifts for D¯()N\bar{D}^{(\ast)}N and B()NB^{(\ast)}N scatterings with I=0I=0 and I=1I=1 in Fig. 1. In the case of D¯N\bar{D}N, the I=0I=0 channel has a bound state below the D¯N\bar{D}N mass threshold as the phase shift starts at δ=π\delta=\pi and it decreases to zero as the scattering energy increases [Fig. 1(a)]. We notice that the D¯N\bar{D}^{\ast}N component feels repulsion due to the existence of the shallow bound state. At first sight, if we look at the phase shift of the D¯N\bar{D}N component in the I=1I=1 channel, then we may notice that the interaction is repulsive and therefore no bound state exists. However, if we turn our attention to the phase shift of the D¯N(2S1/2)\bar{D}^{\ast}N(^{2}S_{1/2}) channel, it starts at δ=π\delta=\pi, indicating the presence of a bound state [Fig. 1(b)]. As a result, we find a bound state that is formed below the D¯N\bar{D}N threshold. In the bottom case, the BNBN interaction in the I=0I=0 channel has a bound state below the BNBN mass threshold, and the BNB^{\ast}N component feels repulsion due to this bound state [Fig. 1(c)]. For I=1I=1, the BN(2S1/2)B^{\ast}N(^{2}S_{1/2}) phase shift also starts as δ=π\delta=\pi [Fig. 1(d)], as well as the D¯N(2S1/2)\bar{D}^{\ast}N(^{2}S_{1/2}) one, indicating that there is a bound state driven by the BNB^{\ast}N component.

In Table 2, we summarize the binding energies and the mixing ratios of PNPN and PNP^{\ast}N components. The bound D¯N\bar{D}N state in I=0I=0 has the binding energy 1.38 MeV. The state is almost dominated by D¯N(2S1/2)\bar{D}N(^{2}{S}_{1/2}) with a small mixture of D¯N(2S1/2)\bar{D}^{\ast}N(^{2}{S}_{1/2}) and D¯N(4D1/2)\bar{D}^{\ast}N(^{4}{D}_{1/2}). Even when the amount of DD-wave component is small, it plays an important role to provide attraction by the tensor interaction in the OPEP as emphasized in our previous papers Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012). In I=1I=1, we also obtain the bound state with the binding energy 5.995.99 MeV. In contrast to the isosinglet state, the bound state has a few amount of the D¯N(4D1/2)\bar{D}^{\ast}N(^{4}{D}_{1/2}) component. This suggests that the D¯N\bar{D}N bound state with I=1I=1 is generated mainly not by the OPEP but by the other potentials. In the present model setting, in fact, the σ\sigma exchange potential provides a strong attraction in the P()NP^{(\ast)}N systems as the σ\sigma exchange potential is strongly attractive for the NNNN system with I=1I=1 in the CD-Bonn potential. In the bottom case, the BNBN states with I=0I=0 and I=1I=1 give deeply bound states with the binding energies 29.7 and 66.0 MeV, respectively. In I=0I=0, the main component is BN(2S1/2)BN(^{2}{S}_{1/2}) with a small amount of BN(2S1/2)B^{\ast}N(^{2}{S}_{1/2}) and BN(4D1/2)B^{\ast}N(^{4}{D}_{1/2}) components. The existence of the DD-wave component indicates again the importance of the OPEP. In I=1I=1, the DD-wave component is negligible as seen in the D¯N\bar{D}N bound state. Interestingly, the BN(2S1/2)B^{\ast}N(^{2}S_{1/2}) channel dominates in the isotriplet bound state, which will be discussed in Sec. IV. The scattering lengths in each state are summarized in Table 3.

The phase shifts for I=1I=1 in Figs. 1(b) and 1(d), starting at δ=π\delta=\pi, imply the existence of the D¯N\bar{D}^{\ast}N and BNB^{\ast}N bound states. In order to confirm this idea, we have performed a bound-state analysis considering only the D¯N\bar{D}^{\ast}N (BNB^{\ast}N) channels, when the D¯N\bar{D}N (BNBN) channel is switched off. As a result, we find D¯N\bar{D}^{\ast}N and BNB^{\ast}N bound states with the binding energies 29.129.1 and 51.051.0 MeV, respectively, measured from the D¯N\bar{D}^{\ast}N (BNB^{\ast}N) threshold.

We investigate the parameter dependence of the attraction in P()NP^{(\ast)}N, where the values of these parameters have some ambiguity in the present model setting. In Fig. 2, we show the dependence of the scattering lengths on the cutoff-ratio parameters, κD¯N\kappa_{\bar{D}N} and κBN\kappa_{BN}. In the D¯N\bar{D}N case, we find that the attraction in I=0I=0 is provided for κD¯N>1.1\kappa_{\bar{D}N}\mathrel{\hbox to0.0pt{\raise 2.20013pt\hbox{$>$}\hss}{\lower 2.20013pt\hbox{$\sim$}}}1.1 whose values are consistent with the one estimated by the ratio of the different hadron sizes of a D¯\bar{D} meson and a nucleon, as previously discussed in Refs. Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012). The strength of attraction in I=1I=1 is not so dependent on the choice of κD¯N\kappa_{\bar{D}N}. In the BNBN case, the attraction in I=0I=0 has only weak dependence on the choice of κBN\kappa_{BN} in the range of κBN>1.0\kappa_{BN}\mathrel{\hbox to0.0pt{\raise 2.20013pt\hbox{$>$}\hss}{\lower 2.20013pt\hbox{$\sim$}}}1.0. This result would tell us a confidence for the existence of the BNBN bound state in I=0I=0. In comparison with I=0I=0, the attraction in I=1I=1 is more sensitive to choice of the value of κBN\kappa_{BN}. Thus the deeply BNBN bound state in I=1I=1 needs to be carefully considered in terms of its model dependence.

Uncertainty in the current model is also brought by the sigma coupling. In general, the coupling constants of the meson exchange potential are fixed by the experimental data, such as the nucleon-nucleon scattering data and heavy meson decays. However, the sigma coupling to the heavy meson is difficult to be determined uniquely only by the currently existing experimental data. In our present calculation framework, we have adopted gσ=gσNN/3g_{\sigma}=g_{\sigma NN}/3 (see Sec. II A2). In order to investigate the uncertainty from the ambiguity of the sigma coupling value, we estimate the dependence of binding energies and scattering lengths on the gσg_{\sigma} coupling constant as shown in Fig 3. Here we show (a) the binding energies, (b) the scattering lengths for PNPN, and (c) the scattering length for PNP^{\ast}N, respectively, for D¯N(I=0)\bar{D}N(I=0), D¯N(I=1)\bar{D}N(I=1), BN(I=0)BN(I=0), and BN(I=1)BN(I=1). For I=0I=0, the binding energies and the scattering lengths are not sensitive to gσg_{\sigma} value indicating that the sigma exchange force is not dominant in I=0I=0; the pion exchange is the most dominant. For I=1I=1, however, the results indicate the sensitiveness to gσg_{\sigma}, i.e., that the sigma exchange force is dominant rather than the other meson exchanges.

The existence of the D¯N\bar{D}N and BNBN bound states in I=0I=0 is consistent with the result in our previous works Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012). However, we should note the difference between the present analysis and the previous one. In the previous case, the π\pi exchange potential was almost dominant among the π\pi, ρ\rho, and ω\omega exchanges. However, the coupling strengths of the meson exchange potentials were incorrectly overestimated by a factor of 2 due to the incorrect normalization of wave functions in Refs. Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012). In the present analysis for I=0I=0, we have also found that similar bound states exist by reconstructing the PNPN interaction model newly including the σ\sigma exchange. Again, the π\pi exchange potential plays the dominate role to produce the attraction. In contrast, the bound states in I=1I=1 have been obtained in the PNPN states, where the main attraction is provided by the σ\sigma potential whose strength in I=1I=1 is set to be larger than that in I=0I=0.

(a) D¯N(I=0)\bar{D}N\,(I=0) (b) D¯N(I=1)\bar{D}N\,(I=1)
Refer to caption Refer to caption
(c) BN(I=0)BN\,(I=0) (d) BN(I=1)BN\,(I=1)
Refer to caption Refer to caption
Figure 2: The scattering lengths of D¯N\bar{D}N [(a) and (b)] and BNBN [(c) and (d)] as functions of the cutoff ratio κD¯N\kappa_{\bar{D}N} and κBN\kappa_{BN}. Panels (a) and (c) are for I=0I=0, and panels (b) and (d) are for I=1I=1.
(a) Binding energy (b) Scattering length for PNPN (c) Scattering length for PNP^{\ast}N
Refer to caption Refer to caption Refer to caption
Figure 3: The dependence of the binding energies [panel (a)] and scattering lengths of the PNPN and PNP^{\ast}N states [panels (b) and (c)] on the sigma coupling strengths. The scale parameter κσ\kappa_{\sigma} is introduced as gσ=κσgσg^{\prime}_{\sigma}=\kappa_{\sigma}g_{\sigma}, i.e., changing gσg_{\sigma} to gσg^{\prime}_{\sigma} as a free coupling value.

IV Discussion

We discuss the internal spin structures of the bound D¯N\bar{D}N and BNBN states in a view of the HQS symmetry. As already discussed in detail in Ref. Yamaguchi et al. (2015), the P()NP^{(\ast)}N state can be decomposed into product states of the heavy antiquark Q¯\bar{Q} and the light quarks qqqqqqqq in the heavy quark limit. The latter component is called the light spin complex, instead of the brown muck, because it makes a specific structure composed of qq and NN which is denoted by [qN]j𝒫[qN]_{j^{\cal P}} with total spin jj and parity 𝒫{\cal P} of the light quark components. These are a conserved quantities due to the spin decoupling from the heavy quark. The important property in the heavy quark limit is that the ratio of the fractions of the amount of PN(2S+1LJ)PN(^{2S+1}L_{J}) and PN(2S+1LJ)P^{\ast}N(^{2S^{\prime}+1}L^{\prime}_{J}) wave functions is determined uniquely. Here SS^{\prime} and LL^{\prime} can be different from SS and LL, respectively, in general. As shown explicitly in Ref. Yamaguchi et al. (2015), we obtain the fractions

PN(2S1/2):PN(2S1/2)=1:3,\displaystyle PN(^{2}{S}_{1/2}):P^{\ast}N(^{2}{S}_{1/2})=1:3, (66)

for j𝒫=0+j^{\cal P}=0^{+} and

PN(2S1/2):PN(2S1/2)=3:1,\displaystyle PN(^{2}{S}_{1/2}):P^{\ast}N(^{2}{S}_{1/2})=3:1, (67)

for j𝒫=1+j^{\cal P}=1^{+}, which hold irrespectively of the choice of the PNPN-PNP^{\ast}N potential. Although these ratios are exact only in the heavy quark limit, they provide us with a guideline to understand the internal spin structures of the obtained D¯N\bar{D}N and BNBN bound states.

In Table 2, for example, we show that the mixing ratios of BN(2S1/2)BN(^{2}{S}_{1/2}) and BN(2S1/2)B^{\ast}N(^{2}{S}_{1/2}) in I=0I=0 are 76.4 % and 14.4%, respectively, which are close to the ratio in Eq. (67) rather than that in Eq. (66). Thus, it is suggested that the BNBN bound state in I=0I=0 is dominated by the light spin complex with j𝒫=1+j^{\cal P}=1^{+}. In contrast, the mixing ratios BN(2S1/2)BN(^{2}{S}_{1/2}) and BN(2S1/2)B^{\ast}N(^{2}{S}_{1/2}) in I=1I=1 are 38.5 % and 61.5 %, respectively, are close to the ratio in Eq. (66) rather than that in Eq. (67). Thus, it is suggested that the BNBN bound state in I=1I=1 includes the light spin complex with j𝒫=0+j^{\cal P}=0^{+} as a major component.

One may wonder that the ratios in bottom sector are not the same as the ratios in Eqs. (66) and (67) in spite of the sufficient heaviness of the bottom quark mass. This would be simply due to the violation of the heavy quark spin symmetry stemming from the difference of the BB meson mass and the BB^{\ast} meson mass, as noted in Ref. Yamaguchi et al. (2015).

We should notice that the existence of the j𝒫=0+j^{\cal P}=0^{+} state is new because only the j𝒫=1+j^{\cal P}=1^{+} state was reported for the π\pi, ρ\rho, and ω\omega potentials in Ref. Yamaguchi et al. (2015). We can understand this new result in terms of the fact that the j𝒫=0+j^{\cal P}=0^{+} state is provided mainly by the σ\sigma potential because of the sufficient attraction in the σ1\sigma_{1} exchange stemming from the characteristic property of the CD-Bonn potential (see Table 4 in Appendix A).

Table 3: SS-wave scattering lengths (aa) of the D¯()N\bar{D}^{(\ast)}N and B()NB^{(\ast)}N states. An attractive scattering length is given by the negative sign (a<0a<0), and a repulsive scattering length and the scattering length for a bound state are given by the positive sign (a>0a>0).
D¯N\bar{D}N                aa [fm]
0(1/2)0(1/2^{-})
D¯N(2S1/2)\bar{D}N(^{2}{S}_{1/2}) 5.21
D¯N(2S1/2)\bar{D}^{\ast}N(^{2}{S}_{1/2}) 0.868i3.72×1020.868-i3.72\times 10^{-2}
1(1/2)1(1/2^{-})
D¯N(2S1/2)\bar{D}N(^{2}{S}_{1/2}) 2.60
D¯N(2S1/2)\bar{D}^{\ast}N(^{2}{S}_{1/2}) 0.944i0.7220.944-i0.722
BNBN                aa [fm]
0(1/2)0(1/2^{-})
BN(2S1/2)BN(^{2}{S}_{1/2}) 1.25
BN(2S1/2)B^{\ast}N(^{2}{S}_{1/2}) 1.03i1.07×1021.03-i1.07\times 10^{-2}
1(1/2)1(1/2^{-})
BN(2S1/2)BN(^{2}{S}_{1/2}) 3.84×1023.84\times 10^{-2}
BN(2S1/2)B^{\ast}N(^{2}{S}_{1/2}) 0.263i0.5850.263-i0.585

V Conclusion

We have discussed the D¯()N\bar{D}^{(\ast)}N and B()NB^{(\ast)}N bound states in terms of the π\pi, σ\sigma, ρ\rho, and ω\omega meson-exchange potentials by considering the heavy-quark spin symmetry and the chiral symmetry. By referring the CD-Bonn potential for the nuclear force, we have constructed the PNPN-PNP^{\ast}N potential with the σ\sigma exchanges as new degrees of freedom at middle-range interaction. We have carefully calculated the potentials with appropriate factors stemming from the normalization of the wave function which were underestimated in our previous studies Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012). As results, we have found that the interaction is largely attractive to hold the D¯N\bar{D}N bound state and the BNBN bound state below the lowest mass threshold for each in I(JP)=0(1/2)I(J^{P})=0(1/2^{-}) channel. Their binding energies are close to the values that were obtained by our previous works. With the present potential including σ\sigma exchange, interestingly, we have found that the σ\sigma exchange as well as the π\pi exchange still plays an important role. We also have found the D¯N\bar{D}N and BNBN bound states in I(JP)=1(1/2)I(J^{P})=1(1/2^{-}) as a new state which has not been discussed so far. It is expected that those states are relevant to the DpD^{-}p interaction researched in LHCb Acharya et al. (2022).

The attraction in PNPN-PNP^{\ast}N systems would open a new way to understand the inter-hadron interaction in heavy flavors. It is important that these systems are made of genuinely five-quark components due to the absence of the annihilation channels. It may help us to understand the new channels of exotic hadrons. Furthermore, the many-body dynamics would be an interesting subject, because the PNPN-PNP^{\ast}N attraction suggests the formation of heavy-flavored nuclei as many-body states having the impurity particles in nuclei Hosaka et al. (2017). Few-body systems such as D¯NN\bar{D}NN (BNNBNNYamaguchi et al. (2014) and D¯α\bar{D}\alpha (D¯{}^{\bar{D}}He) (BαB\alpha (BHe)) are also interesting, which can be accessed through the relativistic heavy ion collisions in LHC and RHIC Cho et al. (2011a, b, 2017). The nuclear structure of charm and bottom nuclei has been studied theoretically for some possible exotic light nuclei  Yamaguchi and Yasui (2017). Experiments at J-PARC, GSI-FAIR, NICA, and so on would also be interesting. In theoretical study, the cross sections for producing such exotic nuclei have been discussed Yamagata-Sekihara et al. (2016). As one of the advanced topics related to heavy-flavored nuclei, the isospin Kondo effect is interesting as it exhibits the “confinement” of isospin charge Yasui and Sudoh (2013); Yasui (2016); Yasui and Sudoh (2017); Yasui (2017); Yasui and Miyamoto (2019). Many subjects are awaiting to be discussed in the future.

Acknowledgment

This work is in part supported by Grants-in-Aid for Scientific Research under Grant Numbers JP20K14478 (Y.Y.) and 21H04478(A) (A.H.). A.H. was also supported by Grant-in-Aid for Scientific Research on Innovative Areas (No. 18H05407).

Appendix A The NNNN potential

We construct the nuclear potential by considering the π\pi, σ\sigma, ρ\rho, and ω\omega exchanges. Their interaction Lagrangians for the vertices with a nucleon are given by

πNN=\displaystyle{\cal L}_{\pi NN}=\, gπNNψ¯iγ5𝝉𝝅ψ,\displaystyle g_{\pi NN}\bar{\psi}i\gamma_{5}\boldsymbol{\tau}\!\cdot\boldsymbol{\pi}\psi, (68)
σINN=\displaystyle{\cal L}_{\sigma_{I}NN}=\, gσINNψ¯σIψ,\displaystyle g_{\sigma_{I}NN}\bar{\psi}\sigma_{I}\psi, (69)
ρNN=\displaystyle{\cal L}_{\rho NN}=\, gρNNψ¯γμ𝝉𝝆μψ\displaystyle g_{\rho NN}\bar{\psi}\gamma_{\mu}\boldsymbol{\tau}\!\cdot\!\boldsymbol{\rho}^{\mu}\psi
+fρNN4mNψ¯σμν𝝉(μ𝝆νν𝝆μ)ψ,\displaystyle+\frac{f_{\rho NN}}{4m_{N}}\bar{\psi}\sigma_{\mu\nu}\boldsymbol{\tau}\!\cdot\!\bigl{(}\partial^{\mu}\boldsymbol{\rho}^{\nu}-\partial^{\nu}\boldsymbol{\rho}^{\mu}\bigr{)}\psi, (70)
ωNN=\displaystyle{\cal L}_{\omega NN}=\, gωNNψ¯γμωμψ,\displaystyle g_{\omega NN}\bar{\psi}\gamma_{\mu}\omega^{\mu}\psi, (71)

with the appropriate coupling constants. We use different σ\sigma mesons: the σ0\sigma_{0} meson for the isosinglet (I=0I=0) NNNN scatterings and the σ1\sigma_{1} meson for the isotriplet (I=1I=1) NNNN scatterings. Their difference appears not only in the coupling constants but also in their masses. We sometimes omit the underscript II if unnecessary. From the Lagrangians (68)-(71), we obtain the NNNN potentials:

Vπ(r)=\displaystyle V_{\pi}(r)=\, (gπNN2mN)213(𝝈1𝝈2Cπ(r)+S12(𝒓^)Tπ(r))𝝉1𝝉2\displaystyle\biggl{(}\frac{g_{\pi NN}}{2m_{N}}\biggr{)}^{2}\frac{1}{3}\Bigl{(}\boldsymbol{\sigma}_{1}\!\cdot\!\boldsymbol{\sigma}_{2}C_{\pi}(r)+S_{12}(\hat{\boldsymbol{r}})T_{\pi}(r)\Bigr{)}\boldsymbol{\tau}_{1}\!\cdot\!\boldsymbol{\tau}_{2}
\displaystyle\equiv\, (𝝈1𝝈2CπNN(r)+S12(𝒓^)TπNN(r))𝝉1𝝉2,\displaystyle\Bigl{(}\boldsymbol{\sigma}_{1}\!\cdot\!\boldsymbol{\sigma}_{2}C_{\pi}^{NN}(r)+S_{12}(\hat{\boldsymbol{r}})T_{\pi}^{NN}(r)\Bigr{)}\boldsymbol{\tau}_{1}\!\cdot\!\boldsymbol{\tau}_{2}, (72)
Vv(r)=\displaystyle V_{v}(r)=\, gvNN2(1mv2+1+fvNN/gvNN2mN2)Cv(r)\displaystyle g_{vNN}^{2}\biggl{(}\frac{1}{m_{v}^{2}}+\frac{1+f_{vNN}/g_{vNN}}{2m_{N}^{2}}\biggr{)}C_{v}(r)
+gvNN2(1+fvNN/gvNN2mN)2\displaystyle+g_{vNN}^{2}\biggl{(}\frac{1+f_{vNN}/g_{vNN}}{2m_{N}}\biggr{)}^{2}
×13(2𝝈1𝝈2Cv(r)S12(𝒓^)Tv(r))\displaystyle\times\frac{1}{3}\Bigl{(}2\boldsymbol{\sigma}_{1}\!\cdot\!\boldsymbol{\sigma}_{2}C_{v}(r)-S_{12}(\hat{\boldsymbol{r}})T_{v}(r)\Bigr{)}
\displaystyle\equiv\, CvNN(r)+2𝝈1𝝈2CvNN(r)S12(𝒓^)TvNN(r),\displaystyle C_{v}^{\prime NN}(r)+2\boldsymbol{\sigma}_{1}\!\cdot\!\boldsymbol{\sigma}_{2}C_{v}^{NN}(r)-S_{12}(\hat{\boldsymbol{r}})T_{v}^{NN}(r), (73)
VσI(r)=\displaystyle V_{\sigma_{I}}(r)=\, (gσINN2mN)2((2mNmσI)21)CσI(r)\displaystyle-\biggl{(}\frac{g_{\sigma_{I}NN}}{2m_{N}}\biggr{)}^{2}\Biggl{(}\biggl{(}\frac{2m_{N}}{m_{\sigma_{I}}}\biggr{)}^{2}-1\Biggr{)}C_{\sigma_{I}}(r)
\displaystyle\equiv\, CσINN(r),\displaystyle-C_{\sigma_{I}}^{NN}(r), (74)

with v=ρv=\rho, ω\omega, where the functions CπNNC_{\pi}^{NN}, TπNNT_{\pi}^{NN}, CσINNC_{\sigma_{I}}^{NN}, CvNNC_{v}^{\prime NN}, CvNNC_{v}^{NN}, and TvNNT_{v}^{NN} are defined as above. More concretely, the NNNN potentials are expressed by

VS13NN(r)\displaystyle V_{{}^{3}{S}_{1}}^{NN}(r) =V¯πNN(r)+V¯σ0NN(r)+V¯ρNN(r)+V¯ωNN(r),\displaystyle=\bar{V}_{\pi}^{NN}(r)+\bar{V}_{\sigma_{0}}^{NN}(r)+\bar{V}_{\rho}^{NN}(r)+\bar{V}_{\omega}^{NN}(r), (75)

with

V¯πNN(r)\displaystyle\bar{V}_{\pi}^{NN}(r) =(3CπNN62TπNN62TπNN3CπNN+6TπNN),\displaystyle=\left(\begin{array}[]{cc}-3C_{\pi}^{NN}&-6\sqrt{2}T_{\pi}^{NN}\\ -6\sqrt{2}T_{\pi}^{NN}&-3C_{\pi}^{NN}+6T_{\pi}^{NN}\end{array}\right), (78)
V¯vNN(r)\displaystyle\bar{V}_{v}^{NN}(r) =(CvNN+2CvNN22TπNN22TπNNCvNN+2CvNN+2TvNN),\displaystyle=\left(\begin{array}[]{cc}C_{v}^{\prime NN}+2C_{v}^{NN}&-2\sqrt{2}T_{\pi}^{NN}\\ -2\sqrt{2}T_{\pi}^{NN}&C_{v}^{\prime NN}+2C_{v}^{NN}+2T_{v}^{NN}\end{array}\right), (81)
V¯σ0NN\displaystyle\bar{V}_{\sigma_{0}}^{NN} =(Cσ0NN00Cσ0NN),\displaystyle=\left(\begin{array}[]{cc}-C_{\sigma_{0}}^{NN}&0\\ 0&-C_{\sigma_{0}}^{NN}\end{array}\right), (84)

in the S13{}^{3}{S}_{1} channel, where the S13{}^{3}S_{1} and D13{}^{3}D_{1} components are coupled, and

VS01NN(r)\displaystyle V_{{}^{1}{S}_{0}}^{NN}(r) =VπNN(r)+Vσ1NN(r)+VρNN(r)+VωNN(r),\displaystyle=V_{\pi}^{NN}(r)+V_{\sigma_{1}}^{NN}(r)+V_{\rho}^{NN}(r)+V_{\omega}^{NN}(r), (85)

with

VπNN(r)\displaystyle V_{\pi}^{NN}(r) =3CπNN(r),\displaystyle=-3C_{\pi}^{NN}(r), (86)
VvNN(r)\displaystyle V_{v}^{NN}(r) =CvNN(r)6CvNN,\displaystyle=C_{v}^{\prime NN}(r)-6C_{v}^{NN}, (87)
Vσ1NN\displaystyle V_{\sigma_{1}}^{NN} =Cσ1NN,\displaystyle=-C_{\sigma_{1}}^{NN}, (88)

in the S01{}^{1}{S}_{0} channel. Notice that the tensor potentials are switched on due to the spin-1 property in the I=0I=0 channel.

We choose the values of the coupling constants to be the same values as those in the CD-Bonn potential Machleidt (2001) as summarized in Table 4. We notice that the CD-Bonn potential originally includes the nonlocal potentials in the π\pi, σ\sigma, ρ\rho, and ω\omega exchanges, and contact terms stemming from the short-range part in the meson exchange. In the present study, however, we neglect the nonlocal potentials, the contact terms and massive σ\sigma mesons, and so on, because we are interested only in the low-energy parts in the NNNN scatterings.

In order to compensate the difference from the CD-Bonn potential, we rescale the cutoff parameter by introducing κI\kappa_{I} providing the new cutoffs ΛN=κIΛNCD-Bonn\Lambda_{N}=\kappa_{I}\Lambda_{N}^{\text{CD-Bonn}}. Here ΛNCD-Bonn\Lambda_{N}^{\text{CD-Bonn}} is the original cutoff parameter in the CD-Bonn potential Machleidt (2001), whose values depend on the exchanged mesons, π\pi, σ\sigma, ρ\rho, and ω\omega. κI\kappa_{I} (I=0I=0 and I=1I=1) are the scale parameter, introduced newly for the adjustment to reproduce the low-energy NNNN scatterings in the present simple model of nuclear force. Notice the values of κI\kappa_{I} are dependent only on the isospin channels I=0I=0 and I=1I=1, while they are common to the π\pi, σ\sigma, ρ\rho, and ω\omega exchanges. We use the values in proton-neutron channel in I=1I=1 in the CD-Bonn potential, because the electric Coulomb force is not included in our potential. We determine the values of κI\kappa_{I} to reproduce the binding energy of a deuteron BdB_{\text{d}} in the S13{}^{3}{S}_{1} (I=0I=0) channel as well as the NNNN scattering length in the S01{}^{1}{S}_{0} (I=1I=1) channel. As the best fitting, we obtain κ0=0.804\kappa_{0}=0.804 for I=0I=0 and κ1=0.773\kappa_{1}=0.773 for I=1I=1. Roughly, we consider that those values would represent the “effective” cutoff parameters when the higher-energy dynamics is renormalized at lower energy near thresholds. Similar values are obtained also when the NNNN scattering length in the S13{}^{3}{S}_{1} (I=0I=0) channel is chosen instead of BdB_{\text{d}}. As shown in Table 5, the obtained values of the scattering lengths and the effective ranges are well consistent with those obtained from the original CD-Bonn potential, a(3S1)=5.419±0.007a(^{3}{S}_{1})=5.419\pm 0.007 fm, re(3S1)=1.753±0.008r_{\text{e}}(^{3}{S}_{1})=1.753\pm 0.008 fm, a(1S0)=23.740±0.020a(^{1}{S}_{0})=-23.740\pm 0.020 fm, re(1S0)=2.77±0.05r_{\text{e}}(^{1}{S}_{0})=2.77\pm 0.05 fm, and Bd=2.225B_{\text{d}}=2.225 MeV, see Ref. Machleidt (2001) for details.

Table 4: Parameters of the local NNNN potentials. σI\sigma_{I} is the σ\sigma meson considered in the NNNN scatterings for the isosinglet (I=0)(I=0) and isotriplet (I=1)(I=1) channels. The meson masses are given as the isospin-averaged values. The coupling constants are taken from Ref. Machleidt (2001). The cutoff parameters ΛN\Lambda_{N} are obtained by scaling the original cutoffs in the CD-Bonn potential Machleidt (2001) by the parameter κI\kappa_{I}, where κ0=0.804\kappa_{0}=0.804 and κ1=0.773\kappa_{1}=0.773 (cf. Table 5), see details in text.
Mesons (α\alpha) Masses [MeV] gαNN24π\frac{g_{\alpha NN}^{2}}{4\pi} fαNNgαNN\frac{f_{\alpha NN}}{g_{\alpha NN}} ΛN\Lambda_{N} [MeV]
I=0I=0 I=1I=1
π\pi 138.04 13.6 1384 1330
ρ\rho 769.68 0.84 6.1 1054 1013
ω\omega 781.94 20 0.0 1207 1159
σ0\sigma_{0} 350 0.51673 2011
σ1\sigma_{1} 452 3.96451 1932
Table 5: The scale parameters κI\kappa_{I} (I=0I=0 and I=1I=1) and the observables in the NNNN scatterings. aa and rer_{e} are the scattering length and the effective range, respectively. BdB_{\text{d}} is the binding energy of a deuteron in I=0I=0. The values with * indicate the input values.
Channel κI\kappa_{I} (I=0,1)(I=0,1) aa [fm] rer_{\text{e}} [fm] BdB_{\text{d}} [MeV]
S13{}^{3}{S}_{1} (I=0I=0) 0.804 5.296 1.562 2.225*
S01{}^{1}{S}_{0} (I=1I=1) 0.773 23.740-23.740* 2.337

Appendix B Potential in a simple model

As an illustration of deriving a potential, we consider a simple model where a potential is provided by the boson exchange interaction (ϕ\phi) between two heavy particles (Φ\Phi). We consider the Lagrangian

[ϕ,Φ]=\displaystyle{\cal L}[\phi,\Phi]=\, 12(μϕμϕm2ϕ2)gϕΦΦ\displaystyle\frac{1}{2}\bigl{(}\partial_{\mu}\phi\,\partial^{\mu}\phi-m^{2}\phi^{2}\bigr{)}-g\phi\Phi^{{\dagger}}\Phi
+μΦμΦM2ΦΦ,\displaystyle+\partial_{\mu}\Phi^{{\dagger}}\,\partial^{\mu}\Phi-M^{2}\Phi^{{\dagger}}\Phi, (89)

with the masses mm and MM for ϕ\phi and Φ\Phi, respectively. From the equation of motion for ϕ\phi, (2+m2)ϕ=gΦΦ(\partial^{2}+m^{2})\phi=-g\Phi^{{\dagger}}\Phi, we obtain the solution

ϕ(x)=gd4yx|(12+m2)xy|yΦ(y)Φ(y),\displaystyle\phi(x)=g\int\mathrm{d}^{4}y\,\langle x|\biggl{(}\frac{-1}{\partial^{2}+m^{2}}\biggr{)}_{xy}|y\rangle\Phi^{{\dagger}}(y)\Phi(y), (90)

for given Φ(y)\Phi(y). As a nonrelativistic limit, making the approximation 2=0222\partial^{2}=\partial_{0}^{2}-\boldsymbol{\partial}^{2}\approx-\boldsymbol{\partial}^{2}, we find that the solution is expressed by

ϕ(𝒙)=gd3𝒚𝒙|(12m2)𝒙𝒚|𝒚Φ(𝒚)Φ(𝒚),\displaystyle\phi(\boldsymbol{x})=g\int\mathrm{d}^{3}\boldsymbol{y}\,\langle\boldsymbol{x}|\biggl{(}\frac{1}{\boldsymbol{\partial}^{2}-m^{2}}\biggr{)}_{\boldsymbol{x}\boldsymbol{y}}|\boldsymbol{y}\rangle\Phi^{{\dagger}}(\boldsymbol{y})\Phi(\boldsymbol{y}), (91)

by dropping the temporal dependence in xμ=(x0,𝒙)x^{\mu}=(x_{0},\boldsymbol{x}) and yμ=(y0,𝒚)y^{\mu}=(y_{0},\boldsymbol{y}). The states |x|x\rangle and |y|y\rangle are also changed to |𝒙|\boldsymbol{x}\rangle and |𝒚|\boldsymbol{y}\rangle, respectively. Hereafter, we omit x0x_{0} and y0y_{0} unless required for specification.

From the Lagrangian (89), we obtain the interaction Hamiltonian Hint=d4xint(x)\displaystyle H_{\mathrm{int}}=\int\mathrm{d}^{4}x\,{\cal H}_{\mathrm{int}}(x) with int(x)=gϕ(x)Φ(x)Φ(x){\cal H}_{\mathrm{int}}(x)=g\phi(x)\Phi^{{\dagger}}(x)\Phi(x). In the following discussion, we express this term by int(𝒙)=gϕ(𝒙)Φ(𝒙)Φ(𝒙){\cal H}_{\mathrm{int}}(\boldsymbol{x})=g\phi(\boldsymbol{x})\Phi^{{\dagger}}(\boldsymbol{x})\Phi(\boldsymbol{x}) because the temporal dependence is dropped in the nonrelativistic approximation. The expectation value of int(𝒙){\cal H}_{\mathrm{int}}(\boldsymbol{x}) leads to the energy shift of the system:

ΔE\displaystyle\Delta E 1,2|d3𝒙int(𝒙)|1,2,\displaystyle\equiv\langle 1,2|\int\mathrm{d}^{3}\boldsymbol{x}\,{\cal H}_{\mathrm{int}}(\boldsymbol{x})|1,2\rangle, (92)

with |1,2=|1|2|1,2\rangle=|1\rangle\otimes|2\rangle where |1|1\rangle and |2|2\rangle denote the heavy-particle states at the position 1 and 2, respectively, at the equal time. By using Eq. (91), we rewrite ΔE\Delta E in the following form:

ΔE\displaystyle\Delta E =g2d3𝒙d3𝒚1,2|Φ(𝒙)Φ(𝒙)𝒙|d3𝒑(2π)3|𝒑𝒑|12m2d3𝒒(2π)3|𝒒𝒒|𝒚Φ(𝒚)Φ(𝒚)|1,2\displaystyle=g^{2}\int\mathrm{d}^{3}\boldsymbol{x}\int\mathrm{d}^{3}\boldsymbol{y}\,\langle 1,2|\Phi^{{\dagger}}(\boldsymbol{x})\Phi(\boldsymbol{x})\langle\boldsymbol{x}|\int\frac{\mathrm{d}^{3}\boldsymbol{p}}{(2\pi)^{3}}|\boldsymbol{p}\rangle\langle\boldsymbol{p}|\frac{1}{\boldsymbol{\partial}^{2}-m^{2}}\int\frac{\mathrm{d}^{3}\boldsymbol{q}}{(2\pi)^{3}}|\boldsymbol{q}\rangle\langle\boldsymbol{q}|\boldsymbol{y}\rangle\Phi^{{\dagger}}(\boldsymbol{y})\Phi(\boldsymbol{y})|1,2\rangle
=d3𝒙d3𝒚1|Φ(𝒙)|00|Φ(𝒙)|1V~ϕ(𝒙,𝒚)2|Φ(𝒚)|00|Φ(𝒚)|2.\displaystyle=\int\mathrm{d}^{3}\boldsymbol{x}\int\mathrm{d}^{3}\boldsymbol{y}\,\langle 1|\Phi^{{\dagger}}(\boldsymbol{x})|0\rangle\langle 0|\Phi(\boldsymbol{x})|1\rangle\tilde{V}_{\phi}(\boldsymbol{x},\boldsymbol{y})\langle 2|\Phi^{{\dagger}}(\boldsymbol{y})|0\rangle\langle 0|\Phi(\boldsymbol{y})|2\rangle. (93)

In the last equation, we have inserted the vacuum state denoted by |0|0\rangle normalized by 0|0=1\langle 0|0\rangle=1. We have used 𝒙|𝒑=ei𝒑𝒙\langle\boldsymbol{x}|\boldsymbol{p}\rangle=e^{i\boldsymbol{p}\cdot\boldsymbol{x}} for the plane wave, and defined the potential by

V~ϕ(𝒙,𝒚)g2d3𝒑(2π)31𝒑2+m2ei𝒑(𝒙𝒚),\displaystyle\tilde{V}_{\phi}(\boldsymbol{x},\boldsymbol{y})\equiv g^{2}\int\frac{\mathrm{d}^{3}\boldsymbol{p}}{(2\pi)^{3}}\frac{-1}{\boldsymbol{p}^{2}+m^{2}}e^{-i\boldsymbol{p}\cdot(\boldsymbol{x}-\boldsymbol{y})}, (94)

between 𝒙\boldsymbol{x} and 𝒚\boldsymbol{y}.

Let us consider the scattering process 𝒑1+𝒑2𝒑1+𝒑2\boldsymbol{p}_{1}+\boldsymbol{p}_{2}\rightarrow\boldsymbol{p}_{1}^{\prime}+\boldsymbol{p}_{2}^{\prime} of two Φ\Phi particles, where the states |1|1\rangle and |2|2\rangle (1|\langle 1| and 2|\langle 2|) have the three-dimensional momenta 𝒑1\boldsymbol{p}_{1} and 𝒑2\boldsymbol{p}_{2} (𝒑1\boldsymbol{p}_{1}^{\prime} and 𝒑2\boldsymbol{p}_{2}^{\prime}), respectively. Here we need to evaluate the wave functions, 0|Φ(𝒙)|1\langle 0|\Phi(\boldsymbol{x})|1\rangle, 0|Φ(𝒚)|2\langle 0|\Phi(\boldsymbol{y})|2\rangle, 1|Φ(𝒙)|0\langle 1|\Phi^{{\dagger}}(\boldsymbol{x})|0\rangle, and 2|Φ(𝒚)|0\langle 2|\Phi^{{\dagger}}(\boldsymbol{y})|0\rangle, in the plane waves with momentum 𝒑1\boldsymbol{p}_{1}, 𝒑2\boldsymbol{p}_{2}, 𝒑1\boldsymbol{p}_{1}^{\prime}, and 𝒑2\boldsymbol{p}_{2}^{\prime}. For this purpose, we expand Φ(𝒙)\Phi(\boldsymbol{x}) by

Φ(𝒙)=d3𝒑(2π)312E𝒑(a𝒑ei𝒑𝒙+b𝒑ei𝒑𝒙),\displaystyle\Phi(\boldsymbol{x})=\int\frac{\mathrm{d}^{3}\boldsymbol{p}}{(2\pi)^{3}}\frac{1}{\sqrt{2E_{\boldsymbol{p}}}}\bigl{(}a_{\boldsymbol{p}}e^{i\boldsymbol{p}\cdot\boldsymbol{x}}+b_{\boldsymbol{p}}^{{\dagger}}e^{-i\boldsymbol{p}\cdot\boldsymbol{x}}\bigr{)}, (95)

according to the conventional forms, where E𝒑=𝒑2+M2E_{\boldsymbol{p}}=\sqrt{\boldsymbol{p}^{2}+M^{2}} is the energy of the heavy particle, and a𝒑a_{\boldsymbol{p}} and b𝒑b_{\boldsymbol{p}} (a𝒑a_{\boldsymbol{p}}^{{\dagger}} and b𝒑b_{\boldsymbol{p}}^{{\dagger}}) are the annihilation (creation) operators for the particle and antiparticle states with three-dimensional momentum 𝒑\boldsymbol{p}. The commutation relations for a𝒑a_{\boldsymbol{p}} and a𝒑a_{\boldsymbol{p}}^{{\dagger}} (b𝒑b_{\boldsymbol{p}} and b𝒑b_{\boldsymbol{p}}^{{\dagger}}) are given by [a𝒑,a𝒑]=[b𝒑,b𝒑]=(2π)3δ(3)(𝒑𝒑)[a_{\boldsymbol{p}},a_{\boldsymbol{p}^{\prime}}^{{\dagger}}]=[b_{\boldsymbol{p}},b_{\boldsymbol{p}^{\prime}}^{{\dagger}}]=(2\pi)^{3}\delta^{(3)}(\boldsymbol{p}-\boldsymbol{p}^{\prime}). In the followings, we consider only the particle state described by a𝒑a_{\boldsymbol{p}} and a𝒑a_{\boldsymbol{p}}^{{\dagger}} by neglecting the antiparticle states.

We consider the state given by |𝒑=2E𝒑a𝒑|0|\boldsymbol{p}\rangle=\sqrt{2E_{\boldsymbol{p}}}a_{\boldsymbol{p}}^{{\dagger}}|0\rangle. The normalization of |𝒑|\boldsymbol{p}\rangle is given by

𝒑|𝒑=2E𝒑(2π)3δ3(0)=2E𝒑V,\displaystyle\langle\boldsymbol{p}|\boldsymbol{p}\rangle=2E_{\boldsymbol{p}}(2\pi)^{3}\delta^{3}(0)=2E_{\boldsymbol{p}}V, (96)

which has the factor 2E𝒑V2E_{\boldsymbol{p}}V, where VV is a volume of the whole space. This indicates that the number of the particle in the wave function is 2E𝒑V2E_{\boldsymbol{p}}V. In Eq. (93), we calculate 0|Φ(𝒙)|1\langle 0|\Phi(\boldsymbol{x})|1\rangle, 0|Φ(𝒚)|2\langle 0|\Phi(\boldsymbol{y})|2\rangle, 1|Φ(𝒙)|0\langle 1|\Phi^{{\dagger}}(\boldsymbol{x})|0\rangle, and 2|Φ(𝒚)|0\langle 2|\Phi^{{\dagger}}(\boldsymbol{y})|0\rangle.

We represent the states by |𝒑1|\boldsymbol{p}_{1}\rangle, |𝒑2|\boldsymbol{p}_{2}\rangle, 𝒑1|\langle\boldsymbol{p}_{1}^{\prime}|, and 𝒑2|\langle\boldsymbol{p}_{2}^{\prime}|, and consider 0|Φ(𝒙)|𝒑1\langle 0|\Phi(\boldsymbol{x})|\boldsymbol{p}_{1}\rangle, 0|Φ(𝒚)|𝒑2\langle 0|\Phi(\boldsymbol{y})|\boldsymbol{p}_{2}\rangle, 𝒑1|Φ(𝒙)|0\langle\boldsymbol{p}_{1}^{\prime}|\Phi^{{\dagger}}(\boldsymbol{x})|0\rangle, and 𝒑2|Φ(𝒚)|0\langle\boldsymbol{p}_{2}^{\prime}|\Phi^{{\dagger}}(\boldsymbol{y})|0\rangle. Using Eq. (95), we obtain

0|Φ(𝒙)|𝒑1\displaystyle\langle 0|\Phi(\boldsymbol{x})|\boldsymbol{p}_{1}\rangle =ei𝒑1𝒙,\displaystyle=e^{i\boldsymbol{p}_{1}\cdot\boldsymbol{x}}, (97)
0|Φ(𝒚)|𝒑2\displaystyle\langle 0|\Phi(\boldsymbol{y})|\boldsymbol{p}_{2}\rangle =ei𝒑2𝒚,\displaystyle=e^{i\boldsymbol{p}_{2}\cdot\boldsymbol{y}}, (98)
𝒑1|Φ(𝒙)|0\displaystyle\langle\boldsymbol{p}_{1}^{\prime}|\Phi^{{\dagger}}(\boldsymbol{x})|0\rangle =ei𝒑1𝒙,\displaystyle=e^{-i\boldsymbol{p}_{1}^{\prime}\cdot\boldsymbol{x}}, (99)
𝒑2|Φ(𝒚)|0\displaystyle\langle\boldsymbol{p}_{2}^{\prime}|\Phi^{{\dagger}}(\boldsymbol{y})|0\rangle =ei𝒑2𝒚.\displaystyle=e^{-i\boldsymbol{p}_{2}^{\prime}\cdot\boldsymbol{y}}. (100)

Then, we find that ΔE\Delta E, which stems from ΔE\Delta E in the relativistic version of the states, is expressed by

ΔE\displaystyle\Delta E =d3𝒙d3𝒚V~ϕ(𝒙,𝒚)ei(𝒑1𝒑1)𝒙ei(𝒑2𝒑2)𝒚.\displaystyle=\int\mathrm{d}^{3}\boldsymbol{x}\int\mathrm{d}^{3}\boldsymbol{y}\,\tilde{V}_{\phi}(\boldsymbol{x},\boldsymbol{y})e^{i(\boldsymbol{p}_{1}-\boldsymbol{p}_{1}^{\prime})\cdot\boldsymbol{x}}e^{i(\boldsymbol{p}_{2}-\boldsymbol{p}_{2}^{\prime})\cdot\boldsymbol{y}}. (101)

When we consider the limit of 𝒑1,𝒑2,𝒑1,𝒑20\boldsymbol{p}_{1},\boldsymbol{p}_{2},\boldsymbol{p}_{1}^{\prime},\boldsymbol{p}_{2}^{\prime}\rightarrow 0 in the static approximation, we express ΔE\Delta E by

ΔE\displaystyle\Delta E d3𝒙d3𝒚V~ϕ(𝒙,𝒚).\displaystyle\approx\int\mathrm{d}^{3}\boldsymbol{x}\int\mathrm{d}^{3}\boldsymbol{y}\,\tilde{V}_{\phi}(\boldsymbol{x},\boldsymbol{y}). (102)

From Eq. (96), we remember that the states |𝒑1|\boldsymbol{p}_{1}\rangle, |𝒑2|\boldsymbol{p}_{2}\rangle, 𝒑1|\langle\boldsymbol{p}_{1}^{\prime}|, and 𝒑2|\langle\boldsymbol{p}_{2}^{\prime}| are normalized to have 2E𝒑1V2E_{\boldsymbol{p}_{1}}V, 2E𝒑2V2E_{\boldsymbol{p}_{2}}V, 2E𝒑1V2E_{\boldsymbol{p}_{1}^{\prime}}V, 2E𝒑2V2MV2E_{\boldsymbol{p}_{2}^{\prime}}V\approx 2MV particles in the nonrelativistic limit. Then, we should regard the quantity ΔE/(2MV)2\Delta E/(2MV)^{2} as the potential energy for a pair of particles. Thus, the energy per a pair of particles is given by

Vϕ(𝒙,𝒚)\displaystyle V_{\phi}(\boldsymbol{x},\boldsymbol{y}) 1(2M)2V~ϕ(𝒙,𝒚),\displaystyle\equiv\frac{1}{(2M)^{2}}\tilde{V}_{\phi}(\boldsymbol{x},\boldsymbol{y}), (103)

with V~ϕ(𝒙,𝒚)\tilde{V}_{\phi}(\boldsymbol{x},\boldsymbol{y}) in Eq. (94). As a conclusion, VϕV_{\phi} is the potential between two Φ\Phi’s used in the non-relativistic quantum mechanics.

Appendix C Derivation of OPEP for a P()P^{(\ast)} meson and a nucleon

From Eqs. (5) and (10), we obtain the Lagrangian including π\pi, NN, and HH (=P,P)(=P,P^{\ast})

πHN=\displaystyle{\cal L}_{\pi HN}=\, 12(μπaμπam2πa2)\displaystyle\frac{1}{2}\bigl{(}\partial_{\mu}\pi_{a}\partial^{\mu}\pi_{a}-m^{2}\pi_{a}^{2}\bigr{)}
+igπfπενρμσvνPβρ(𝝉μ𝝅)βαPασ\displaystyle+\frac{ig_{\pi}}{f_{\pi}}\varepsilon_{\nu\rho\mu\sigma}v^{\nu}P_{\beta}^{\ast\rho{\dagger}}\bigl{(}\boldsymbol{\tau}\!\cdot\!\partial^{\mu}\boldsymbol{\pi}\bigr{)}_{\beta\alpha}P_{\alpha}^{\ast\sigma}
+iigπfπPβμ(𝝉μ𝝅)βαPα\displaystyle+i\frac{ig_{\pi}}{f_{\pi}}P_{\beta\mu}^{\ast{\dagger}}\bigl{(}\boldsymbol{\tau}\!\cdot\!\partial^{\mu}\boldsymbol{\pi}\bigr{)}_{\beta\alpha}P_{\alpha}
+iigπfπPβ(𝝉μ𝝅)βαPαμ\displaystyle+i\frac{ig_{\pi}}{f_{\pi}}P_{\beta}^{{\dagger}}\bigl{(}\boldsymbol{\tau}\!\cdot\!\partial^{\mu}\boldsymbol{\pi}\bigr{)}_{\beta\alpha}P_{\alpha\mu}^{\ast}
+gπNN2mNψ¯βγμγ5(𝝉μ𝝅)βαψα,\displaystyle+\frac{g_{\pi NN}}{2m_{N}}\bar{\psi}_{\beta}\gamma_{\mu}\gamma_{5}(\boldsymbol{\tau}\!\cdot\!\partial^{\mu}\boldsymbol{\pi})_{\beta\alpha}\psi_{\alpha}, (104)

where the kinetic terms of HH and NN are not shown. The equation of motion for π\pi is

(2+m2)πa=igπfπμ(ενρμσvνPβρ(τa)βαPασ+iPβμ(τa)βαPα+iPβ(τa)βαPαμ)gπNN2mNμ(ψ¯βγμγ5(τa)βαψα).\displaystyle(\partial^{2}+m^{2})\pi_{a}=-\frac{ig_{\pi}}{f_{\pi}}\partial^{\mu}\Bigl{(}\varepsilon_{\nu\rho\mu\sigma}v^{\nu}P_{\beta}^{\ast\rho{\dagger}}(\tau_{a})_{\beta\alpha}P_{\alpha}^{\ast\sigma}+iP_{\beta\mu}^{\ast{\dagger}}(\tau_{a})_{\beta\alpha}P_{\alpha}{+i}P_{\beta}^{{\dagger}}(\tau_{a})_{\beta\alpha}P_{\alpha\mu}^{\ast}\Bigr{)}{-}\frac{g_{\pi NN}}{2m_{N}}\partial_{\mu}\bigl{(}\bar{\psi}_{\beta}\gamma^{\mu}\gamma_{5}(\tau_{a})_{\beta\alpha}\psi_{\alpha}\bigr{)}. (105)

When we consider only the spatial dependence in the fields, we express the solution by

πa(𝒙)=\displaystyle\pi_{a}(\boldsymbol{x})= igπfπd3𝒚𝒙|12+m2|𝒚yj(ετχjωvτPδχ(𝒚)(τa)δγPγω(𝒚)+iPδj(𝒚)(τa)δγPγ(𝒚)+iPδ(𝒚)(τa)δγPγj(𝒚))\displaystyle-\frac{ig_{\pi}}{f_{\pi}}\int\mathrm{d}^{3}\boldsymbol{y}\,\langle\boldsymbol{x}|\frac{1}{-\boldsymbol{\partial}^{2}+m^{2}}|\boldsymbol{y}\rangle\partial_{y}^{j}\Bigl{(}\varepsilon_{\tau\chi j\omega}v^{\tau}P_{\delta}^{\ast\chi{\dagger}}(\boldsymbol{y})(\tau_{a})_{\delta\gamma}P_{\gamma}^{\ast\omega}(\boldsymbol{y}){+i}P_{\delta j}^{\ast{\dagger}}(\boldsymbol{y})(\tau_{a})_{\delta\gamma}P_{\gamma}(\boldsymbol{y}){+i}P_{\delta}^{{\dagger}}(\boldsymbol{y})(\tau_{a})_{\delta\gamma}P_{\gamma j}^{\ast}(\boldsymbol{y})\Bigr{)}
gπNN2mNd3𝒚𝒙|12+m2|𝒚yj(ψ¯β(𝒚)γjγ5(τa)βαψα(𝒚)),\displaystyle{-}\frac{g_{\pi NN}}{2m_{N}}\int\mathrm{d}^{3}\boldsymbol{y}\,\langle\boldsymbol{x}|\frac{1}{-\boldsymbol{\partial}^{2}+m^{2}}|\boldsymbol{y}\rangle\partial_{yj}\bigl{(}\bar{\psi}_{\beta}(\boldsymbol{y})\gamma^{j}\gamma_{5}(\tau_{a})_{\beta\alpha}\psi_{\alpha}(\boldsymbol{y})\bigr{)}, (106)

with i,j=1,2,3i,j=1,2,3 for given ψ\psi, PP, and PP^{\ast}. Then, the interaction energy between P()P^{(\ast)} and NN is given by

ΔEHN\displaystyle\Delta E^{HN}\equiv\, 1,2|d3𝒙intπHN(𝒙)|1,2\displaystyle\langle 1,2|\int\mathrm{d}^{3}\boldsymbol{x}\,{\cal H}_{\mathrm{int}}^{\pi HN}(\boldsymbol{x})|1,2\rangle
=\displaystyle=\, d3𝒙d3𝒚1|(εiklPβk(𝒙)(τa)βαPαl(𝒙)+iPβi(𝒙)(τa)βαPα(𝒙)+iPβ(𝒙)(τa)βαPαi(𝒙))|1\displaystyle\int\mathrm{d}^{3}\boldsymbol{x}\int\mathrm{d}^{3}\boldsymbol{y}\,\langle 1|\Bigl{(}-\varepsilon_{ikl}P_{\beta}^{\ast k{\dagger}}(\boldsymbol{x})(\tau_{a})_{\beta\alpha}P_{\alpha}^{\ast l}(\boldsymbol{x}){+i}P_{\beta i}^{\ast{\dagger}}(\boldsymbol{x})(\tau_{a})_{\beta\alpha}P_{\alpha}(\boldsymbol{x}){+i}P_{\beta}^{{\dagger}}(\boldsymbol{x})(\tau_{a})_{\beta\alpha}P_{\alpha i}^{\ast}(\boldsymbol{x})\Bigr{)}|1\rangle
×V~πijHN(𝒙,𝒚)2|ψ¯β(𝒚)γjγ5(τa)βαψα(𝒚)|2,\displaystyle\times\tilde{V}_{\pi\,ij}^{HN}(\boldsymbol{x},\boldsymbol{y})\langle 2|\bar{\psi}_{\beta^{\prime}}(\boldsymbol{y})\gamma^{j}\gamma_{5}(\tau_{a})_{\beta^{\prime}\alpha^{\prime}}\psi_{\alpha^{\prime}}(\boldsymbol{y})|2\rangle, (107)

where intπHN{\cal H}_{\mathrm{int}}^{\pi HN} represents the interaction Hamiltonian stemming from Eq. (104), and |1|1\rangle and |2|2\rangle represent a P()P^{(\ast)} meson and a nucleon, respectively. For brevity we have defined V~πijHN(𝒙,𝒚)\tilde{V}_{\pi\,ij}^{HN}(\boldsymbol{x},\boldsymbol{y}) by

V~πijHN(𝒙,𝒚)\displaystyle\tilde{V}_{\pi\,ij}^{HN}(\boldsymbol{x},\boldsymbol{y}) gπNN2mNigπfπd3𝒑(2π)3pipj𝒑2+m2ei𝒑(𝒙𝒚).\displaystyle\equiv{-}\frac{g_{\pi NN}}{2m_{N}}\frac{ig_{\pi}}{f_{\pi}}\int\frac{\mathrm{d}^{3}\boldsymbol{p}}{(2\pi)^{3}}\frac{p_{i}p_{j}}{\boldsymbol{p}^{2}+m^{2}}e^{-i\boldsymbol{p}\cdot(\boldsymbol{x}-\boldsymbol{y})}. (108)

We consider the matrix element by using the basis states |1=|Pα1(𝒑1,λ1)|1\rangle=|P_{\alpha_{1}}^{\ast}(\boldsymbol{p}_{1},\lambda_{1})\rangle or |Pα1(𝒑1)|P_{\alpha_{1}}(\boldsymbol{p}_{1})\rangle and 1|=Pβ1(𝒑1,λ1)|\langle 1|=\langle P_{\beta_{1}}^{\ast}(\boldsymbol{p}_{1}^{\prime},\lambda_{1}^{\prime})| or Pβ1(𝒑1)|\langle P_{\beta_{1}}(\boldsymbol{p}_{1}^{\prime})|. Here 𝒑1\boldsymbol{p}_{1} (𝒑1\boldsymbol{p}_{1}^{\prime}) is the three-dimensional momentum of the P()P^{(\ast)} meson and λ1\lambda_{1} (λ1\lambda_{1}^{\prime}) is the helicity of the PP^{\ast} meson (λ1,λ1=0,±\lambda_{1},\lambda_{1}^{\prime}=0,\pm). α1,β1=±1/2\alpha_{1},\beta_{1}=\pm 1/2 are the isospin components. Adopting the following channels,

{1|,|1}=\displaystyle\bigl{\{}\langle 1|,|1\rangle\bigr{\}}= {Pβ1(𝒑1,λ1)|,|Pα1(𝒑1,λ1)},\displaystyle\bigl{\{}\langle P_{\beta_{1}}^{\ast}(\boldsymbol{p}_{1}^{\prime},\lambda_{1}^{\prime})|,|P_{\alpha_{1}}^{\ast}(\boldsymbol{p}_{1},\lambda_{1})\rangle\bigr{\}},
{Pβ1(𝒑1,λ1)|,|Pα1(𝒑1)},\displaystyle\bigl{\{}\langle P_{\beta_{1}}^{\ast}(\boldsymbol{p}_{1}^{\prime},\lambda_{1}^{\prime})|,|P_{\alpha_{1}}(\boldsymbol{p}_{1})\rangle\bigr{\}},
{Pβ1(𝒑1)|,|Pα1(𝒑1,λ1)},\displaystyle\bigl{\{}\langle P_{\beta_{1}}(\boldsymbol{p}_{1}^{\prime})|,|P_{\alpha_{1}}^{\ast}(\boldsymbol{p}_{1},\lambda_{1})\rangle\bigr{\}}, (109)

and

{2|,|2}=\displaystyle\bigl{\{}\langle 2|,|2\rangle\bigr{\}}= {Nβ2(𝒑2,s2)|,|Nα2(𝒑2,s2)},\displaystyle\bigl{\{}\langle N_{\beta_{2}}(\boldsymbol{p}_{2}^{\prime},s_{2}^{\prime})|,|N_{\alpha_{2}}(\boldsymbol{p}_{2},s_{2})\rangle\bigr{\}}, (110)

we obtain the potential energy in each channel:

ΔEPN-PN=\displaystyle\Delta E_{P^{\ast}N\text{-}P^{\ast}N}= d3𝒙d3𝒚D¯β1(𝒑1,λ1)|(εiklPβk(𝒙)(τa)βαPαl(𝒙))|D¯α1(𝒑1,λ1)V~πijHN(𝒙,𝒚)\displaystyle\int\mathrm{d}^{3}\boldsymbol{x}\int\mathrm{d}^{3}\boldsymbol{y}\,\langle\bar{D}_{\beta_{1}}^{\ast}(\boldsymbol{p}_{1}^{\prime},\lambda_{1}^{\prime})|\Bigl{(}-\varepsilon_{ikl}P_{\beta}^{\ast k{\dagger}}(\boldsymbol{x})(\tau_{a})_{\beta\alpha}P_{\alpha}^{\ast l}(\boldsymbol{x})\Bigr{)}|\bar{D}_{\alpha_{1}}^{\ast}(\boldsymbol{p}_{1},\lambda_{1})\rangle\tilde{V}_{\pi ij}^{HN}(\boldsymbol{x},\boldsymbol{y})
×Nβ2(𝒑2,s2)|ψ¯β(𝒚)γjγ5(τa)βαψα(𝒚)|Nα2(𝒑2,s2),\displaystyle\times\langle N_{\beta_{2}}(\boldsymbol{p}_{2}^{\prime},s_{2}^{\prime})|\bar{\psi}_{\beta^{\prime}}(\boldsymbol{y})\gamma^{j}\gamma_{5}(\tau_{a})_{\beta^{\prime}\alpha^{\prime}}\psi_{\alpha^{\prime}}(\boldsymbol{y})|N_{\alpha_{2}}(\boldsymbol{p}_{2},s_{2})\rangle, (111)
ΔEPN-PN=\displaystyle\Delta E_{P^{\ast}N\text{-}PN}= d3𝒙d3𝒚D¯β1(𝒑1,λ1)|(iPβi(𝒙)(τa)βαPα(𝒙))|D¯α1(𝒑1)V~πijHN(𝒙,𝒚)\displaystyle\int\mathrm{d}^{3}\boldsymbol{x}\int\mathrm{d}^{3}\boldsymbol{y}\,\langle\bar{D}_{\beta_{1}}^{\ast}(\boldsymbol{p}_{1}^{\prime},\lambda_{1}^{\prime})|\Bigl{(}iP_{\beta i}^{\ast{\dagger}}(\boldsymbol{x})(\tau_{a})_{\beta\alpha}P_{\alpha}(\boldsymbol{x})\Bigr{)}|\bar{D}_{\alpha_{1}}(\boldsymbol{p}_{1})\rangle\tilde{V}_{\pi ij}^{HN}(\boldsymbol{x},\boldsymbol{y})
×Nβ2(𝒑2,s2)|ψ¯β(𝒚)γjγ5(τa)βαψα(𝒚)|Nα2(𝒑2,s2),\displaystyle\times\langle N_{\beta_{2}}(\boldsymbol{p}_{2}^{\prime},s_{2}^{\prime})|\bar{\psi}_{\beta^{\prime}}(\boldsymbol{y})\gamma^{j}\gamma_{5}(\tau_{a})_{\beta^{\prime}\alpha^{\prime}}\psi_{\alpha^{\prime}}(\boldsymbol{y})|N_{\alpha_{2}}(\boldsymbol{p}_{2},s_{2})\rangle, (112)
ΔEPN-PN=\displaystyle\Delta E_{PN\text{-}P^{\ast}N}= d3𝒙d3𝒚D¯β1(𝒑1)|(iPβ(𝒙)(τa)βαPαi(𝒙))|D¯α1(𝒑1,λ1)V~πijHN(𝒙,𝒚)\displaystyle\int\mathrm{d}^{3}\boldsymbol{x}\int\mathrm{d}^{3}\boldsymbol{y}\,\langle\bar{D}_{\beta_{1}}(\boldsymbol{p}_{1}^{\prime})|\Bigl{(}iP_{\beta}^{{\dagger}}(\boldsymbol{x})(\tau_{a})_{\beta\alpha}P_{\alpha i}^{\ast}(\boldsymbol{x})\Bigr{)}|\bar{D}_{\alpha_{1}}^{\ast}(\boldsymbol{p}_{1},\lambda_{1})\rangle\tilde{V}_{\pi ij}^{HN}(\boldsymbol{x},\boldsymbol{y})
×Nβ2(𝒑2,s2)|ψ¯β(𝒚)γjγ5(τa)βαψα(𝒚)|Nα2(𝒑2,s2).\displaystyle\times\langle N_{\beta_{2}}(\boldsymbol{p}_{2}^{\prime},s_{2}^{\prime})|\bar{\psi}_{\beta^{\prime}}(\boldsymbol{y})\gamma^{j}\gamma_{5}(\tau_{a})_{\beta^{\prime}\alpha^{\prime}}\psi_{\alpha^{\prime}}(\boldsymbol{y})|N_{\alpha_{2}}(\boldsymbol{p}_{2},s_{2})\rangle. (113)

Here 𝒑2\boldsymbol{p}_{2} and 𝒑2\boldsymbol{p}_{2}^{\prime} are the three-dimensional momenta of the nucleon, s2,s2=±1/2s_{2},s_{2}^{\prime}=\pm 1/2 are the spin components, and α2,β2=±1/2\alpha_{2},\beta_{2}=\pm 1/2 are the isospin components.

In order to calculate the matrix elements, we expand Pαi(𝒙)P_{\alpha}^{\ast i}(\boldsymbol{x}) and Pα(𝒙)P_{\alpha}(\boldsymbol{x}) by plane waves. This is obtained by considering multiplying the mass scale MM to Eq. (95) and taking the large MM limit. The results are

Pαi(𝒙)\displaystyle P^{\ast i}_{\alpha}(\boldsymbol{x}) =d3𝒑(2π)312(a𝒑αiei𝒑𝒙+b𝒑αiei𝒑𝒙),\displaystyle=\int\frac{\mathrm{d}^{3}\boldsymbol{p}}{(2\pi)^{3}}\frac{1}{\sqrt{2}}\bigl{(}a_{\boldsymbol{p}\alpha}^{i}e^{i\boldsymbol{p}\cdot\boldsymbol{x}}+b_{\boldsymbol{p}\alpha}^{i{\dagger}}e^{-i\boldsymbol{p}\cdot\boldsymbol{x}}\bigr{)}, (114)
Pα(𝒙)\displaystyle P_{\alpha}(\boldsymbol{x}) =d3𝒑(2π)312(a𝒑αei𝒑𝒙+b𝒑αei𝒑𝒙),\displaystyle=\int\frac{\mathrm{d}^{3}\boldsymbol{p}}{(2\pi)^{3}}\frac{1}{\sqrt{2}}\bigl{(}a_{\boldsymbol{p}\alpha}e^{i\boldsymbol{p}\cdot\boldsymbol{x}}+b_{\boldsymbol{p}\alpha}^{{\dagger}}e^{-i\boldsymbol{p}\cdot\boldsymbol{x}}\bigr{)}, (115)

where the factor 1/21/\sqrt{2} stems from 1/2E𝒑1/\sqrt{2E_{\boldsymbol{p}}} in the conventional representation multiplied by the factor M\sqrt{M} and taking the large MM limit. Here a𝒑αia_{\boldsymbol{p}\alpha}^{i} and b𝒑αib_{\boldsymbol{p}\alpha}^{i} (a𝒑αa_{\boldsymbol{p}\alpha} and b𝒑αb_{\boldsymbol{p}\alpha}) satisfy the commutation relations, [a𝒑i,a𝒑j]=[b𝒑i,b𝒑j]=(2π)3δijδ(3)(𝒑𝒑)[a_{\boldsymbol{p}}^{i},a_{\boldsymbol{p}^{\prime}}^{j{\dagger}}]=[b_{\boldsymbol{p}}^{i},b_{\boldsymbol{p}^{\prime}}^{j{\dagger}}]=(2\pi)^{3}\delta^{ij}\delta^{(3)}(\boldsymbol{p}-\boldsymbol{p}^{\prime}) and [a𝒑,a𝒑]=[b𝒑,b𝒑]=(2π)3δ(3)(𝒑𝒑)[a_{\boldsymbol{p}},a_{\boldsymbol{p}^{\prime}}^{{\dagger}}]=[b_{\boldsymbol{p}},b_{\boldsymbol{p}^{\prime}}^{{\dagger}}]=(2\pi)^{3}\delta^{(3)}(\boldsymbol{p}-\boldsymbol{p}^{\prime}). Let us consider the large MM limit and leave only the leading terms of MM. Because the particle states |Pα(𝒑,λ)|P_{\alpha}^{\ast}(\boldsymbol{p},\lambda)\rangle and |Pα(𝒑)|P_{\alpha}(\boldsymbol{p})\rangle are defined by

|Pα(𝒑,λ)\displaystyle|P_{\alpha}^{\ast}(\boldsymbol{p},\lambda)\rangle 2ϵi(λ)(a𝒑αi)|0,\displaystyle\equiv\sqrt{2}\epsilon_{i}^{(\lambda)}(a_{\boldsymbol{p}\alpha}^{i})^{{\dagger}}|0\rangle, (116)
|Pα(𝒑)\displaystyle|P_{\alpha}(\boldsymbol{p})\rangle 2(a𝒑αi)|0,\displaystyle\equiv\sqrt{2}(a_{\boldsymbol{p}\alpha}^{i})^{{\dagger}}|0\rangle, (117)

which indicate that the states |Pα(𝒑,λ)|P_{\alpha}^{\ast}(\boldsymbol{p},\lambda)\rangle and |Pα(𝒑)|P_{\alpha}(\boldsymbol{p})\rangle include 2V2V particles in the volume VV. The polarization vectors for the PP^{\ast} meson are given by Eq. (21). We also use TiT_{i} (i=1,2,3i=1,2,3) in Eq. (32). As for the nucleon part, we consider the expansion for ψ(𝒙)\psi(\boldsymbol{x}) given by

ψ(𝒙)\displaystyle\psi(\boldsymbol{x}) =d3𝒑(2π)312E𝒑\displaystyle=\int\frac{\mathrm{d}^{3}\boldsymbol{p}}{(2\pi)^{3}}\frac{1}{\sqrt{2E_{\boldsymbol{p}}}}
×s=±1/2α=±1/2(a𝒑sαus(𝒑)ei𝒑𝒙+b𝒑sαvs(𝒑)ei𝒑𝒙),\displaystyle\times\sum_{s=\pm 1/2}\sum_{\alpha=\pm 1/2}\Bigl{(}a_{\boldsymbol{p}s\alpha}u_{s}(\boldsymbol{p})e^{i\boldsymbol{p}\cdot\boldsymbol{x}}+b_{\boldsymbol{p}s\alpha}v_{s}(\boldsymbol{p})e^{-i\boldsymbol{p}\cdot\boldsymbol{x}}\Bigr{)}, (118)

with the commutation relations

[a𝒑rβ,a𝒒sα]=[b𝒑rβ,b𝒒sα]=(2π)3δ(3)(𝒑𝒒)δrsδβα,\displaystyle[a_{\boldsymbol{p}r\beta},a_{\boldsymbol{q}s\alpha}^{{\dagger}}]=[b_{\boldsymbol{p}r\beta},b_{\boldsymbol{q}s\alpha}^{{\dagger}}]=(2\pi)^{3}\delta^{(3)}(\boldsymbol{p}-\boldsymbol{q})\delta_{rs}\delta_{\beta\alpha}, (119)

for spin s,rs,r and isospin α,β\alpha,\beta. The normalizations for uu and vv by ur(𝒑)us(𝒑)=vr(𝒑)vs(𝒑)=2E𝒑ξrξsu_{r}(\boldsymbol{p})^{{\dagger}}u_{s}(\boldsymbol{p})=v_{r}(\boldsymbol{p})^{{\dagger}}v_{s}(\boldsymbol{p})=2E_{\boldsymbol{p}}\xi_{r}^{{\dagger}}\xi_{s} and u¯r(𝒑)us(𝒑)=v¯r(𝒑)vs(𝒑)=2mNξrξs\bar{u}_{r}(\boldsymbol{p})u_{s}(\boldsymbol{p})=-\bar{v}_{r}(\boldsymbol{p})v_{s}(\boldsymbol{p})=2m_{N}\xi_{r}^{{\dagger}}\xi_{s} with E𝒑=𝒑2+mN2E_{\boldsymbol{p}}=\sqrt{\boldsymbol{p}^{2}+m_{N}^{2}} for the nucleon mass mNm_{N}. The concrete forms of us(𝒑)u_{s}(\boldsymbol{p}) and vs(𝒑)v_{s}(\boldsymbol{p}) are given by

us(𝒑)\displaystyle u_{s}(\boldsymbol{p}) =E𝒑+mN(ξs𝒑𝝈E𝒑+mNξs),\displaystyle=\sqrt{E_{\boldsymbol{p}}+m_{N}}\left(\begin{array}[]{c}\xi_{s}\\ \dfrac{\boldsymbol{p}\!\cdot\!\boldsymbol{\sigma}}{E_{\boldsymbol{p}}+m_{N}}\xi_{s}\end{array}\right), (122)
vs(𝒑)\displaystyle v_{s}(\boldsymbol{p}) =E𝒑+mN(𝒑𝝈E𝒑+mNζsζs),\displaystyle=\sqrt{E_{\boldsymbol{p}}+m_{N}}\left(\begin{array}[]{c}\dfrac{\boldsymbol{p}\!\cdot\!\boldsymbol{\sigma}}{E_{\boldsymbol{p}}+m_{N}}\zeta_{s}\\ \zeta_{s}\end{array}\right), (125)

for the standard representation of the Dirac matrices, where us(𝒑)us(𝒑)=vs(𝒑)vs(𝒑)=2E𝒑u_{s}(\boldsymbol{p})^{{\dagger}}u_{s}(\boldsymbol{p})=v_{s}(\boldsymbol{p})^{{\dagger}}v_{s}(\boldsymbol{p})=2E_{\boldsymbol{p}} holds for the normalization |ξs|2=|ζs|2=1|\xi_{s}|^{2}=|\zeta_{s}|^{2}=1. We consider the scattering process for the nucleons, (𝒑1,s1,α1)+(𝒑2,s2,α2)(𝒑1,s1,α1)+(𝒑2,s2,α2)(\boldsymbol{p}_{1},s_{1},\alpha_{1})+(\boldsymbol{p}_{2},s_{2},\alpha_{2})\rightarrow(\boldsymbol{p}_{1}^{\prime},s_{1}^{\prime},\alpha_{1}^{\prime})+(\boldsymbol{p}_{2}^{\prime},s_{2}^{\prime},\alpha_{2}^{\prime}), with 𝒑i\boldsymbol{p}_{i} (𝒑i\boldsymbol{p}_{i}^{\prime}) for the initial (final) momentum, sis_{i} and αi\alpha_{i} (sis_{i}^{\prime} and αi\alpha_{i}^{\prime}) for the initial (final) spin and isospin for the nucleon i=1,2i=1,2. The wave functions are denoted by |1=|𝒑1s1α1|1\rangle=|\boldsymbol{p}_{1}s_{1}\alpha_{1}\rangle, |2=|𝒑2s2α2|2\rangle=|\boldsymbol{p}_{2}s_{2}\alpha_{2}\rangle, 1|=𝒑1s1α1|\langle 1|=\langle\boldsymbol{p}_{1}^{\prime}s_{1}^{\prime}\alpha_{1}^{\prime}|, and 2|=𝒑2s2α2|\langle 2|=\langle\boldsymbol{p}_{2}^{\prime}s_{2}^{\prime}\alpha_{2}^{\prime}|. We define the plane-wave state by

|Nsα(𝒑)2E𝒑a𝒑sα|0,\displaystyle|N_{s\alpha}(\boldsymbol{p})\rangle\equiv\sqrt{2E_{\boldsymbol{p}}}\,a_{\boldsymbol{p}s\alpha}^{{\dagger}}|0\rangle, (126)

for the vacuum state |0|0\rangle with the normalization 0|0=1\langle 0|0\rangle=1. The normalization for |𝒑sα|\boldsymbol{p}s\alpha\rangle is given by 𝒑sα|𝒑sα=2E𝒑(2π)3δ(3)(𝒑𝒑)δssδαα\langle\boldsymbol{p}^{\prime}s^{\prime}\alpha^{\prime}|\boldsymbol{p}s\alpha\rangle=2E_{\boldsymbol{p}}(2\pi)^{3}\delta^{(3)}(\boldsymbol{p}^{\prime}-\boldsymbol{p})\delta_{s^{\prime}s}\delta_{\alpha^{\prime}\alpha}.

From Eqs. (111)-(113), we obtain the potentials

VπPN-PN(𝒙,𝒚)\displaystyle V_{\pi}^{P^{\ast}N\text{-}P^{\ast}N}(\boldsymbol{x},\boldsymbol{y}) =12gπNN2mNgπfπd3𝒑(2π)3pipj𝒑2+m2ei𝒑(𝒙𝒚)(Ti)λ1λ1(σj)s2s2(τa)β1α1(τa)β2α2,\displaystyle={-}{\frac{1}{2}}\frac{g_{\pi NN}}{2m_{N}}\frac{g_{\pi}}{f_{\pi}}\int\frac{\mathrm{d}^{3}\boldsymbol{p}}{(2\pi)^{3}}\frac{p_{i}p_{j}}{\boldsymbol{p}^{2}+m^{2}}e^{-i\boldsymbol{p}\cdot(\boldsymbol{x}-\boldsymbol{y})}(T_{i})_{\lambda_{1}^{\prime}\lambda_{1}}(\sigma_{j})_{s_{2}^{\prime}s_{2}}(\tau_{a})_{\beta_{1}\alpha_{1}}(\tau_{a})_{\beta_{2}\alpha_{2}}, (127)
VπPN-PN(𝒙,𝒚)\displaystyle V_{\pi}^{P^{\ast}N\text{-}PN}(\boldsymbol{x},\boldsymbol{y}) =12gπNN2mNgπfπd3𝒑(2π)3pipj𝒑2+m2ei𝒑(𝒙𝒚)ϵi(λ1)(σj)s2s2(τa)β1α1(τa)β2α2,\displaystyle=\frac{1}{2}\frac{g_{\pi NN}}{2m_{N}}\frac{g_{\pi}}{f_{\pi}}\int\frac{\mathrm{d}^{3}\boldsymbol{p}}{(2\pi)^{3}}\frac{p_{i}p_{j}}{\boldsymbol{p}^{2}+m^{2}}e^{-i\boldsymbol{p}\cdot(\boldsymbol{x}-\boldsymbol{y})}\epsilon_{i}^{(\lambda_{1}^{\prime})*}(\sigma_{j})_{s_{2}^{\prime}s_{2}}(\tau_{a})_{\beta_{1}\alpha_{1}}(\tau_{a})_{\beta_{2}\alpha_{2}}, (128)
VπPN-PN(𝒙,𝒚)\displaystyle V_{\pi}^{PN\text{-}P^{\ast}N}(\boldsymbol{x},\boldsymbol{y}) =12gπNN2mNgπfπd3𝒑(2π)3pipj𝒑2+m2ei𝒑(𝒙𝒚)ϵi(λ1)(σj)s2s2(τa)β1α1(τa)β2α2,\displaystyle=\frac{1}{2}\frac{g_{\pi NN}}{2m_{N}}\frac{g_{\pi}}{f_{\pi}}\int\frac{\mathrm{d}^{3}\boldsymbol{p}}{(2\pi)^{3}}\frac{p_{i}p_{j}}{\boldsymbol{p}^{2}+m^{2}}e^{-i\boldsymbol{p}\cdot(\boldsymbol{x}-\boldsymbol{y})}\epsilon_{i}^{(\lambda_{1})}(\sigma_{j})_{s_{2}^{\prime}s_{2}}(\tau_{a})_{\beta_{1}\alpha_{1}}(\tau_{a})_{\beta_{2}\alpha_{2}}, (129)

to be transformed to Eqs. (12), (13), and (14) in the end. Notice that the factor 1/21/2 in the coefficients have been missed in the previous studies by the authors Yasui and Sudoh (2009); Yamaguchi et al. (2011, 2012). The calculation of the momentum integrations is easily performed by introducing the form factor (20) in the integrands. In the calculations, it is useful to adopt the formula of the plane-wave expansion

ei𝒑𝒓=4πl,lz(i)ljl(pr)Yllz(𝒑^)Yllz(𝒓^),\displaystyle e^{-i\boldsymbol{p}\cdot\boldsymbol{r}}=4\pi\sum_{l,l_{z}}(-i)^{l}j_{l}(pr)Y_{ll_{z}}^{\ast}(\hat{\boldsymbol{p}})Y_{ll_{z}}(\hat{\boldsymbol{r}}), (130)

with l=0,1,2,l=0,1,2,\dots and lz=l,l+1,,l1,ll_{z}=-l,-l+1,\dots,l-1,l. Here jl(x)j_{l}(x) is the spherical Bessel function and Yllz(𝒙^)Y_{ll_{z}}(\hat{\boldsymbol{x}}) is the spherical harmonic function. As a result, we obtain the explicit forms of the central potential C(r;m)C(r;m) and the tensor potential T(r;m)T(r;m) in Eqs. (16) and (17), respectively. In the calculation of the tensor potential, we have used the relationship

aibjSij(𝒑^)\displaystyle a_{i}b_{j}S_{ij}(\hat{\boldsymbol{p}}) =24π5μ=22(1)μ(𝒂×𝒃)μ(2)Y2μ(𝒑^),\displaystyle=\sqrt{\frac{24\pi}{5}}\sum_{\mu=-2}^{2}(-1)^{\mu}(\boldsymbol{a}\times\boldsymbol{b})^{(2)}_{\mu}Y_{2\mu}(\hat{\boldsymbol{p}}), (131)

where (𝒂×𝒃)μ(2)(\boldsymbol{a}\times\boldsymbol{b})^{(2)}_{\mu} is the rank-2 tensor composed of 𝒂=(a1,a2,a3)\boldsymbol{a}=(a_{1},a_{2},a_{3}) and 𝒃=(b1,b2,b3)\boldsymbol{b}=(b_{1},b_{2},b_{3}).

References