exchange in the one-boson exchange model involving the ground state octet baryons
Abstract
Based on the one-boson-exchange framework that the meson serves as an effective parameterization for the correlated scalar-isoscalar interaction, we calculate the coupling constants of the to the ground state light baryon octet by matching the amplitude of to that of . The former is calculated using a dispersion relation, supplemented with chiral perturbation theory results for the couplings and the Muskhelishvili-Omnès representation for the rescattering. Explicitly, the coupling constants are obtained as , , , and . These coupling constants can be used in the one-boson-exchange model calculations of the interaction of light baryons with other hadrons.
I Introduction
In the past few decades, the observation of exotic hadronic states, which cannot be accounted for by the conventional quark model, has propelled the study of exotic states to the forefront of hadron physics; see Refs. [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14] for recent reviews on the experimental and theoretical status. Hadronic molecules [7], one of the most promising candidates for exotic states, are loosely bound states of hadrons and a natural extension of the atomic nuclei (such as the deuteron as a proton-neutron bound state) and offer an explanation of the many experimentally observed near-threshold structures, in particular in the heavy-flavor hadron mass region [15].
As a generalization of the one-pion-exchange potential [16], the one-boson-exchange (OBE) model has played a crucial role in studying composite systems of hadrons [17, 18, 19, 20, 21, 22, 23, 24, 25, 10, 11]. Taking the deuteron as an example, it is widely accepted in the OBE model that its formation involves the long-range interaction from the one-pion exchange and the middle-range interaction from the -meson exchange; see Ref. [18] for a detailed review. However, unlike narrow width particles that are associated with clear resonance peaks or dips observed in experiments, the scalar-isoscalar meson, which plays a crucial role in nuclear and hadron physics, had remained a subject of considerable debates for several decades until it was established as the lowest-lying hadronic resonance in quantum chromodynamics (QCD) in the past twenty years based on rigorous dispersive analyses of scattering [26, 27, 28] (see, e.g., Refs. [29, 30] for reviews). The dispersive techniques have recently been applied to determine the nature of the at unphysical pion masses [31, 32].
The meson in the OBE model can be considered as approximating the correlated -wave isoscalar exchange in a few hundred MeV range [33, 18, 34, 35, 36, 37, 38], and some modifications to its properties have been made in order to improve the accuracy of the approximation [33, 18]. However, the effective coupling constants between the and various hadrons remain highly uncertain. One example is the widely used nucleon-nucleon- coupling , which ranges roughly from 8 to 14 [33, 18, 36]. For its couplings to other ground state octet baryons, , , and , there are rare systematic discussions and error analyses. Most of them are estimated either by the quenched quark model or the SU(3) symmetry model assuming the to be a certain member of the light-flavor multiplet [24, 23]. The use of the one- exchange instead of the correlated exchange may raise some questions: Is this approximation reasonable? How good is the approximation? In the present work, we try to address these questions by considering the scattering of the baryon and antibaryon in the ground state octet, , through the intermediate state of the correlated pair or the meson and calculate the effective coupling constants of . We will make use of dispersion relations following Ref. [39], and similar methods have been used in, e.g., Refs. [37, 38] to derive the baryon-baryon- couplings, and Ref. [40] to derive the coupling to heavy mesons. Here, we will match the dispersive amplitudes of at low energies to the chiral amplitudes up to the next-to-leading order (NLO).
This paper is structured as follows. The formalism is presented in Sec. II, including the calculation of the OBE amplitude in Sec. II.1 and the amplitude from the dispersion relation (DR) with a careful treatment of kinematical singularities in Sec. II.2. In Sec. III, we conduct an analysis of the two amplitudes and present the scalar coupling constants along with an error analysis. This includes a comparison of the -channel processes, as detailed in Sec. III.1, and a comparison of the corresponding -channel processes utilizing the crossing symmetry in Sec. III.2. A brief summary is given in Sec. IV. The adopted conventions, the result of coupling in the SU(2) framework, and crossing symmetry relations are relegated to the appendices.
II Formalism
To determine the scalar coupling, , between the baryon in the ground baryon octet and the meson, we first utilize the DR and chiral perturbation theory (ChPT) to calculate the amplitude of with a correlated intermediate state, as depicted in Fig. 1(a). Here, and represent the four momentum and the helicity of particle or , respectively. Additionally, we restrict the quantum numbers of the two-body intermediate state, , to be . This -wave amplitude can be denoted as , where the subscript 0 indicates the -wave, represents the square of the total energy of the system in the center-of-mass (c.m.) frame,111For an -channel process of , as illustrated in Fig. 1, while . and the superscript DR indicates the result obtained from the DR. Next, we proceed to calculate the same amplitude, with the intermediate meson, as depicted in Fig. 1(b). In this case, we utilize the OBE model, and the corresponding amplitude can be expressed as . Finally, we compare the aforementioned amplitudes to extract the coupling constant in the phenomenological baryon-baryon- coupling Lagrangian.
II.1 The OBE amplitude
According to the following effective Lagrangian coupling the meson to the baryons in the SU(3) flavor octet [18, 41],
(1) |
the OBE amplitude for the Feynman diagram depicted in Fig. 1(b) reads
(2) |
where is a flavor factor, , , and . For simplicity, we choose throughout the paper.222Other choices, e.g., physical amplitudes using the orbital-spin basis are also accessible. The final results do not depend on the choice. With this choice we have
(3) |
where is the isospin averaged mass of the baryon .333Since we are not interested in the isospin symmetry breaking effects, the isospin averaged mass is used for all particles within the same isospin multiplet.
II.2 The dispersive representation
II.2.1 The DR and the kinematical singularity
One can write down a dispersive representation of the scattering amplitude corresponding to Fig. 1(a) as
(4) |
Here a once-subtracted dispersive integral is employed to facilitate the convergence of the dispersive integral. The threshold is chosen as the subtraction point, and we set the subtraction constant as Eq. (3) since the two amplitudes will be matched later.
In order to capture the rescattering in the region and avoid the interference from other resonances, e.g., the , the upper limit of the dispersive integral in Eq. (4) is set to , as in Ref. [39] (see also Ref. [42]). We will investigate the impact of varying the upper limit of the integration on the final result and regard it as a part of the uncertainty of the coupling constants.
Next, let us discuss the discontinuity. Taking into account the unitary relation that is fulfilled by the partial-wave -matrix elements, we can express the discontinuity of the -wave amplitude (here the partial wave refers to that between the pions) along the cut in terms of and . However, it is crucial to notice that when dealing with systems that involve spins, particularly those containing fermions, kinematical singularities arise [43]. These singularities stem from the definition of the wave functions for the initial and final states. Following Ref. [43], we introduce the kinematical-singularity-free amplitudes,
(5) | |||
(6) |
Then the unitary relation for the -wave -matrix elements is given by
(7) |
where is the two-body phase space factor. Moreover, as we will discuss in detail in Sec. II.2.4, the treatment of kinematical singularity plays a vital role in guaranteeing the self-consistency of the theory.
Furthermore, with the phase conventions outlined in Appendix A and considering the isospin scalar system, we obtain the following relations
(8) |
Then, we obtain the following discontinuities,
(9) | |||
(10) | |||
(11) | |||
(12) |
II.2.2 The SU(3) ChPT framework
To obtain the low-energy -wave amplitudes , we need the corresponding chiral baryon-meson Lagrangian. The leading-order (LO) chiral Lagrangian is given by [44],
(13) |
which contains three low-energy constants (LECs), , and . Here, means trace in the flavor space, the baryon octet are collected in the matrix,
(17) |
and the chiral vielbein and covariant derivative are given by
(18) |
with , , where
(22) |
Notice that the chiral connection containing two pions is a vector, the two pions from that term cannot be in the -wave. As a result, only the - and -channel exchanges depicted in Fig. 2 (a) and (b), which contribute to the LHC part of , will be present in the LO calculation. In addition, the LO Lagrangian contains the coupling terms of the form and . Therefore, it is necessary to consider the exchange of in the process and the exchange of in the process.
The chiral Lagrangian contains the contact contribution with the -wave pion pair, as illustrated in Fig. 2 (c). At the NLO, the number of LECs increases, and the Lagrangian reads [45, 46],444Here we use the Lagrangian in [46] while the one in Ref. [45] has redundant terms. Note also the ordering of the operators in Ref. [46] is different from that in Refs. [45, 47].
(23) | ||||
where with a constant related to the quark condensate in the chiral limit and the light-quark mass matrix. We will use the values of involved LECs from Fit II in Ref. [48], which are , , , , , , , , , , , , .
II.2.3 The partial-wave amplitudes
Using the LO and NLO Lagrangians given in Eqs. (13, 23), we can calculate the tree-level amplitude for the process of as depicted in Fig. 2. However, in order to determine the final state with , we need to perform a partial-wave (PW) expansion. The generalized PW expansion of the helicity amplitude for arbitrary spin can be found in Ref. [49]. The final PW amplitude for reads
(24) |
where is the relative orbital angular momentum of the pions, and are the third components of the helicities of and . The basis is such that is expanded in terms of , and the relative momentum is chosen to be along the axis so that ; are the polar and azimuthal angles of the relative momentum.
For the tree-level -wave amplitude for , the LHC part from the - and -channel baryon exchange is555Here we consider only the baryon exchanges such that the two mesons emitted are two pions since we focus on the correlated -wave two-pion exchange. That is, although we use an SU(3) chiral Lagrangian, the exchanged baryon has the same strangeness as the external ones. The framework may be understood as an SU(2) one for each of the baryons, but with the LECs matched to those in the SU(3) Lagrangian.
(25) | ||||
(26) | ||||
(27) | ||||
(28) |
where
The contact terms, which are from the NLO Lagrangian and contribute to the RHC part of after taking into account the rescattering, read
(29) | ||||
(30) | ||||
(31) | ||||
(32) |
where the parameter is the decay constant of the in the chiral limit. Since we use the LECs determined in Ref. [48], we adopt the same value MeV [50] for consistency.
Moreover, employing Eq. (5), the tree-level -wave amplitudes for after eliminating the kinematical singularities read, for the LHC part,
(33) | ||||
(34) | ||||
(35) | ||||
(36) |
and for the contact term part,
(37) | ||||
(38) | ||||
(39) | ||||
(40) |
II.2.4 The Muskhelishvili-Omnès representation
We now incorporate the rescattering based on the tree-level amplitude, into the Muskhelishvili-Omnès representation. For the process, we partition the total -wave kinematical-singularity-free amplitude into the LHC and the RHC parts,
(41) |
Utilizing the amplitude in the scalar-isoscalar channel , where is the -wave isoscalar phase shift, and since there is no overlap between the LHC and RHC for kinematic-singularity-free amplitudes,666The RHC is chosen to be along the positive axis in the interval . The LHC is in the interval for the - or -channel process of , where represents the mass of the exchanged particle. It can be easily proven that . the unitary relation implies,
(42) |
To solve this equation, we first define the Omnès function [51],
(43) |
By using , we further derive
(44) |
Therefore, we can derive a DR with subtractions,
(45) |
where is an arbitrary polynomial of order . Finally, we obtain a DR for as
(46) |
For the phase shift , we take the parametrization in Ref. [52]. For the Omnès function, we take the matrix element of the coupled-channel Omnès matrix for the - -wave interaction obtained in Ref. [53].
The above equation provides a reasonable form that incorporates the rescattering. The LHC part and the polynomial may be determined by matching at low energies to the chiral amplitudes as done in Refs. [54, 55, 56, 57, 58]. We perform the matching when the rescattering is switched off, i.e., , which leads to . Consequently, for the process, we can approximate and .
Moreover, there is a polynomial ambiguity as discussed in Refs. [59, 60]. If the asymptotic value of the phase shift is not 0 but as , the corresponding Omnès function will approach asymptotically. In our case, the phase shift implies , thus the general solution of the unitarity condition (42) contains free parameters [59, 60] (assuming that is asymptotically bounded by ). However, although the standard twice subtracted DR via Eq. (46) indeed grows like (notice that ), it contains only free parameters in the polynomial, i.e., one parameter less than the general solution. Hence we propose an oversubtracted DR (twice subtracted DR with an order-2 polynomial matching to the ChPT amplitudes) that can be solved uniquely. In summary, the final DR is given as
(47) |
From the above derivation, it is important to note that Eq. (47) can only be applied when the singularity of the LHC is exclusively included in , and there is no overlap between the LHC and RHC. The original - and -channel exchange amplitudes Eqs. (25-28) do not satisfy this condition due to the factor . Let us take as an example. From Fig. 3, it becomes apparent that the amplitude in Eq. (26) includes the LHC derived from the particle exchanging in the crossed channel, as well as a kinematical cut in the physical region. Therefore, directly substituting Eq. (26) into Eq. (47) is invalid and disrupts the self-consistency of the theory. By employing the method described in Sec. II.2.1 to eliminate the kinematical singularities, the kinematical-singularity-free -wave amplitude in Eq. (34) has only the LHC and satisfies the condition for Eq. (47), as demonstrated in Fig. 4.
At this stage, we can substitute the amplitude given by Eqs. (33-40) into Eq. (47) to obtain the amplitude denoted as . It includes the -wave rescattering and does not exhibit any kinematical singularities. Then, utilizing Eqs. (9-12), we get the discontinuity in Eq. (4), and finally, the DR amplitude for from exchanging correlated -wave is obtained by performing the dispersive integral.
III Determination of coupling constants
III.1 Matching -channel amplitudes
Let us first compare the two amplitudes in Eqs. (3, 4) in the -channel physical region, specifically . Since the amplitudes from exchanging and from exchanging the correlated -wave have the same Lorentz structure, we can compare the two amplitudes at large values so that the pion masses and the mass in the OBE amplitude play little role. A comparison of and in the physical region of is shown in Fig. 5, where has been adjusted so that the two amplitudes coincide in the physical region and GeV is taken. In fact, matching Eqs. (3, 4) at , one gets
(48) |
Since is much larger than both or , one obtains the following sum rule:
(49) |
The numerical results of the scalar coupling constants are presented in Table 1, where the uncertainties in the second to fourth columns arise from the error propagated from those of the NLO LECs and the choice of the upper limit for the dispersive integral (see below), corresponding to Eqs. (4, 47). In addition to the results obtained in the SU(3) framework, we also investigate in the SU(2) framework. The details are presented in Appendix B, and the results are listed in the last row in Table 1, labeled as . Moreover, remarks are made on the difference in under the SU(2) and SU(3) frameworks in Appendix B.
Let us comment on the calculation of the two dispersive integrals. The first one, given by Eq. (47), is computed over the integration range of . The second one, given by Eq. (4), is integrated over .777Here, represents a small positive quantity that is relatively insignificant when compared to and . As long as it is much smaller than , the specific value has negligible impact on the results. Note that the range of the second integral is completely covered by that of the first one to avoid unphysical singularities.
The central values in Table 1 are obtained by setting to 0.8 GeV as in Ref. [39] and utilizing the central values of the NLO LECs provided in Ref. [48]. The uncertainties of the NLO LECs as determined in Ref. [48] are propagated to the coupling constants by using the bootstrap method. The resulting average values and corresponding standard deviations introduce the first source of errors in the third and fourth columns in Table 1 (the LECs appear only in the RHC contributions, and we have fixed the pion decay constant; thus the second column does not have errors from LECs). Furthermore, we vary from 0.7 GeV to 0.9 GeV, which constitute the errors in the second column and the second source of errors in the third and fourth columns.
LHC | RHC | Total | [33] | [18] | [34] | [37] | [36] | [24] | [23] | ||
---|---|---|---|---|---|---|---|---|---|---|---|
- | - | 10.85(8.92) | 4.65 | - | - | - | |||||
- | - | - | - | - | 3.4 | - | |||||
- | - | 8.18(6.54) | 4.37 | - | - | 6.59 | |||||
12.78 | 8.46 | 8.46 | 8.58 | 13.85 | 10.2 | 9.86 | |||||
Results from other studies on these scalar couplings are also listed in Table 1. For that has been estimated in many works, we find a good agreement with existing results, which supports the validity of our framework. Here we briefly discuss the methods used in the literature. In Ref. [33], the authors investigated the -wave amplitudes with the rescattering and the results revealed that the intertwined contribution from the -wave can be elegantly described as a broad -meson with a mass of approximately and a coupling strength of . In Ref. [18], displaying the outcomes derived from the Bonn meson-exchange model, they found that the correlated -wave exchange can be further approximated by a zero width scalar exchange, with the corresponding mass and coupling constant readjusted to 550 MeV and 8.46, respectively. In Ref. [34], the authors also considered the exchange as an effective parameterization for the correlated -wave exchange contribution. They utilized the result from the full Bonn meson-exchange model [18] for the nucleon, i.e., the value in the sixth column of Table 1, and and are determined by a fit to the empirical hyperon-nucleon data using two different models, with the distinction lying in whether higher-order processes involving a spin- baryon in the intermediate state were considered in the hyperon-nucleon interaction. In Ref. [37], the authors calculated the and amplitudes in the light of hadron-exchange picture. Based on an ansatz for Lagrangian, various symmetries and assumptions, they reduced the number of free parameters as many as possible, and then the parameters were fixed by adjusting the amplitudes to the quasi-empirical data. With these parameters and the existing scattering phase shifts they got the and amplitudes in the pseudo-physical region after solving the Blankenbecler-Sugar equation. Then employing the DR they got the spectral function which denotes the strength of a hadron-exchange process, namely the coupling constants. The eighth column in Table 1 represents their results, which are also similar to those reported in Ref. [38]. In later development of the Jülich meson-exchange model in Ref. [36], the authors conducted an analysis of the coupled-channel dynamics and performed a simultaneous fit to the experimental data of various reactions, including , , and , with the intermediate state parameterized as the , and channels. In their fitting, the coupling constant is determined to be . In Ref. [23], the authors used from SU(3) consideration and took from Ref. [18]. In Ref. [24], was determined using the linear model [61]. Then under the assumption that the meson only couples to the and quarks, the authors got based on the quark model consideration. Additionally, in Ref. [62], the authors calculated the potential arising from the exchange of a correlated -wave isoscalar pion pair, i.e., the channel, utilizing a unitary approach based on the lowest order chiral Lagrangian and the Bethe-Salpeter equation for the analysis of scattering. A qualitative estimate for was obtained, at the right order of the values quoted in Table 1.
III.2 Matching -channel amplitudes
In the preceding subsection, it becomes evident that for the -channel process of , the selection of an apt coupling constant allows for the exchange to mimic the correlated intermediate state with in physical region, . However, when employing the OBE model to estimate the interaction between hadrons, the is exchanged in the - or -channel, as illustrated in Fig. 6 (b) rather than in the -channel, as demonstrated in Fig. 1 (b). Therefore, to derive the parameters for the exchange that can be used in the OBE model, one needs to conduct an analysis of the - and -channel meson-exchange processes. As elaborated in Appendix C, the crossing symmetry relations provide a means to relate the -channel process to the -channel one. It becomes evident that, should we manage to align the two amplitudes within the non-physical region of the -channel process, specifically , we can subsequently match the corresponding pair of amplitudes within the physical region of the -channel process, i.e., , relevant for the low-energy scattering.
In order for the exchange to approximate the -wave correlated two pions in the few hundred MeV region, we also need to adjust the mass in addition to the couplings derived above.101010Since in the scattering physical region, the exchanged two pions cannot go on shell, a real mass, instead of the complex pole, for the meson in the OBE model should be used. As an example, in Fig. 7, we show the comparison of the OBE amplitude and the DR amplitude for the case at the -channel threshold. One finds from Fig. 7 (b) that by adjusting the mass to about MeV, the DR amplitude using the central values of the LECs can be very well reproduced. The matching point has been chosen to be GeV2, corresponding to the -channel threshold. To see the dependence on the mass, we also show the comparison for MeV in Fig. 7 (a).
The aforementioned analysis can be readily extended to the other ground state octet baryons, yielding the results shown in Fig. 8. From Figs. 5, 7 (b) and 8, it is apparent that if our aim is to use a simple exchange in the OBE model to concurrently match a complex correlated exchange with in the -, - and -channel physical region, the values required by different processes differ. Specifically, we find MeV, MeV, MeV and MeV, where the uncertainties correspond to those of the couplings added in quadrature.111111The superscript of is utilized to represent the mass of this which is derived from the process of . These values are listed in the last column of Table 1. This echoes previous attempts to modify the mass of , a broad resonance with a mass approximately equal to [33], to a mass of 550 MeV with a zero width [18], which is within all the above ranges. The goal of such modification was to allow a single exchange to more accurately replicate the results of a correlated exchange with .
IV summary
In this work, we evaluate the couplings of the meson to the ground state light baryons, which are essential inputs of the OBE models, by matching the baryon-baryon scattering amplitudes through correlated -wave isoscalar intermediate state to the OBE ones. Using the LO and NLO SU(3) chiral baryon-meson Lagrangians, we carefully handle the kinematical singularities and utilize DR and incorporate the rescattering by Muskhelishvili-Omnès representation to obtain the DR amplitude. Considering the phenomenological exchange as an effective parameterization for the correlated exchange contribution in the channel, we determine the scalar coupling constants from the -channel matching, as listed in Table 1. Specifically, , , , and , where the errors are obtained by adding the corresponding ones in Table 1 in quadrature. This is achieved by comparing the DR amplitude and OBE amplitude in the physical region of the -channel process, specifically, . Concurrently, we estimate the uncertainties of the scalar coupling constants arising from the NLO LECs [48] and variation of the upper limit for the dispersive integral. Moreover, by extending the analysis to the physical region of the corresponding -channel process via the crossing relation, we obtain the mass to be used together with the determined coupling constant. The value depends on the process but is always around 550 MeV. We also compute the coupling by matching to the SU(2) CHPT amplitude with the LECs determined in Refs. [63, 64], and the result is .
The effective coupling constants obtained here can be used to describe the interaction between light hadrons and other hadrons through the exchange. The same method can be applied to the determination of the coupling constants of and other hadrons, such as heavy mesons and baryons, the interactions between which are crucial to understand the abundance of exotic hadron candidates observed at various experiments in last two decades.
Acknowledgements.
We would like to thank Ulf-G. Meißner for a careful reading of the manuscript. This work is supported in part by the Chinese Academy of Sciences under Grants No. XDB34030000 and No. YSBR-101; by the National Natural Science Foundation of China (NSFC) and the Deutsche Forschungsgemeinschaft (DFG) through the funds provided to the Sino-German Collaborative Research Center TRR110 “Symmetries and the Emergence of Structure in QCD” (NSFC Grant No. 12070131001, DFG Project-ID 196253076); by the NSFC under Grants No. 12125507, No. 11835015, and No. 12047503; and by the Postdoctoral Fellowship Program of China Postdoctoral Science Foundation (CPSF) under No. GZC20232773 and the CPSF No. 2023M743601.Appendix A Isospin conventions
In this work, we use the following isospin conventions [65]:
Therefore, we can readily obtain the isoscalar state composed of , and in the particle basis.
Appendix B from SU(2) ChPT
It is worth mentioning that in the context of interaction, it is more common to utilize the Lagrangian within the SU(2) framework, the LO Lagrangian is given by
(50) |
where represents the nucleon axial-vector coupling constant in the chiral limit and is related to the SU(3) LECs via . At the NLO,
(51) |
which contains seven LECs [66, 67, 68, 69], the first four of which are determined in Refs. [64, 63] as (in units of ),
(52) |
By utilizing the Eqs. (50, 51) and the above LECs, we obtain the following results through the -channel matching as detailed in Sect. III.1:
(53) |
Notice that here for consistency with the values, we take MeV and used in Refs. [64, 63], larger than the value used in the main text. Meanwhile, from matching the -channel amplitudes, we find MeV. The value geiven above is close to the real part of the coupling defined as the residue of the amplitude at the pole obtained in Ref. [70], which is .
As per Table 1, the central value calculated using the ChPT NLO Lagrangian within the SU(2) framework deviates from its value within the SU(3) framework. In Fig. 9, we show a comparison of the -wave tree-level amplitudes of the contact terms for the process from the SU(3) chiral Lagrangian with that from the SU(2) chiral Lagrangian, the LECs of which are taken from Ref. [48] and Refs. [64, 63], respectively. One sees a clear deviation. We have checked that the deviation from the SU(2) result would be larger if we use the central values of the SU(3) LECs determined by other groups [71, 72, 73, 47]. Nevertheless, the values of and from RHC contributions agree within uncertainties. One notices that Refs. [48] and Refs. [64, 63] considered different experimental and lattice data sets.

Appendix C The crossing relation
Based on the crossing symmetry, we can establish a relation between the -channel helicity amplitude of and the -channel helicity amplitude of or the -channel helicity amplitude of . Using crossing symmetry relations for systems with spin [74, 43, 75],121212In the context of crossing relation, for a -channel process of , as illustrated in Fig. 6, refers to while refers to . the amplitude for the -channel process of via the correlated intermediate state with can be expressed as
(54) |
where represents the Wigner rotation angles corresponding to the Lorentz transformation from the -channel c.m. frame to the -channel c.m. frame, and the subscript signifies that the of either the -channel process or the -channel process forms an isoscalar -wave. Considering that the crossing relation, Eq. (54), is solely dependent on the particles of the external lines, the same relation is applicable regardless of whether there is a exchange or a correlated exchange, namely,
(55) | ||||
(56) |
We then obtain
(57) |
Since our goal is to ensure that the amplitude of the correlated exchange with and that of the exchange are approximately the same for the -channel process of within the -channel physical region, i.e., , we require the corresponding -channel amplitudes of to approximate each other as well as possible, i.e.,
(58) |
when . The -channel process mirrors this exactly.
References
- Hosaka et al. [2016] A. Hosaka, T. Iijima, K. Miyabayashi, Y. Sakai, and S. Yasui, Exotic hadrons with heavy flavors: , , , and related states, PTEP 2016, 062C01 (2016), arXiv:1603.09229 [hep-ph] .
- Richard [2016] J.-M. Richard, Exotic hadrons: review and perspectives, Few Body Syst. 57, 1185 (2016), arXiv:1606.08593 [hep-ph] .
- Lebed et al. [2017] R. F. Lebed, R. E. Mitchell, and E. S. Swanson, Heavy-Quark QCD Exotica, Prog. Part. Nucl. Phys. 93, 143 (2017), arXiv:1610.04528 [hep-ph] .
- Esposito et al. [2017] A. Esposito, A. Pilloni, and A. D. Polosa, Multiquark Resonances, Phys. Rept. 668, 1 (2017), arXiv:1611.07920 [hep-ph] .
- Ali et al. [2017] A. Ali, J. S. Lange, and S. Stone, Exotics: Heavy Pentaquarks and Tetraquarks, Prog. Part. Nucl. Phys. 97, 123 (2017), arXiv:1706.00610 [hep-ph] .
- Olsen et al. [2018] S. L. Olsen, T. Skwarnicki, and D. Zieminska, Nonstandard heavy mesons and baryons: Experimental evidence, Rev. Mod. Phys. 90, 015003 (2018), arXiv:1708.04012 [hep-ph] .
- Guo et al. [2018] F.-K. Guo, C. Hanhart, U.-G. Meißner, Q. Wang, Q. Zhao, and B.-S. Zou, Hadronic molecules, Rev. Mod. Phys. 90, 015004 (2018), [Erratum: Rev.Mod.Phys. 94, 029901 (2022)], arXiv:1705.00141 [hep-ph] .
- Brambilla et al. [2020] N. Brambilla, S. Eidelman, C. Hanhart, A. Nefediev, C.-P. Shen, C. E. Thomas, A. Vairo, and C.-Z. Yuan, The states: experimental and theoretical status and perspectives, Phys. Rept. 873, 1 (2020), arXiv:1907.07583 [hep-ex] .
- Yamaguchi et al. [2020] Y. Yamaguchi, A. Hosaka, S. Takeuchi, and M. Takizawa, Heavy hadronic molecules with pion exchange and quark core couplings: a guide for practitioners, J. Phys. G 47, 053001 (2020), arXiv:1908.08790 [hep-ph] .
- Dong et al. [2021a] X.-K. Dong, F.-K. Guo, and B.-S. Zou, A survey of heavy-antiheavy hadronic molecules, Progr. Phys. 41, 65 (2021a), arXiv:2101.01021 [hep-ph] .
- Dong et al. [2021b] X.-K. Dong, F.-K. Guo, and B.-S. Zou, A survey of heavy–heavy hadronic molecules, Commun. Theor. Phys. 73, 125201 (2021b), arXiv:2108.02673 [hep-ph] .
- Chen et al. [2023] H.-X. Chen, W. Chen, X. Liu, Y.-R. Liu, and S.-L. Zhu, An updated review of the new hadron states, Rept. Prog. Phys. 86, 026201 (2023), arXiv:2204.02649 [hep-ph] .
- Meng et al. [2023] L. Meng, B. Wang, G.-J. Wang, and S.-L. Zhu, Chiral perturbation theory for heavy hadrons and chiral effective field theory for heavy hadronic molecules, Phys. Rept. 1019, 1 (2023), arXiv:2204.08716 [hep-ph] .
- Mai et al. [2023] M. Mai, U.-G. Meißner, and C. Urbach, Towards a theory of hadron resonances, Phys. Rept. 1001, 1 (2023), arXiv:2206.01477 [hep-ph] .
- Dong et al. [2021c] X.-K. Dong, F.-K. Guo, and B.-S. Zou, Explaining the Many Threshold Structures in the Heavy-Quark Hadron Spectrum, Phys. Rev. Lett. 126, 152001 (2021c), arXiv:2011.14517 [hep-ph] .
- Yukawa [1935] H. Yukawa, On the Interaction of Elementary Particles I, Proc. Phys. Math. Soc. Jap. 17, 48 (1935).
- Durso et al. [1977] J. W. Durso, M. Saarela, G. E. Brown, and A. D. Jackson, Isobars, Transition Potentials and Short Range Repulsion in the Nucleon-Nucleon Interaction, Nucl. Phys. A 278, 445 (1977).
- Machleidt et al. [1987] R. Machleidt, K. Holinde, and C. Elster, The Bonn Meson Exchange Model for the Nucleon Nucleon Interaction, Phys. Rept. 149, 1 (1987).
- Törnqvist [1994] N. A. Törnqvist, From the deuteron to deusons, an analysis of deuteron-like meson meson bound states, Z. Phys. C 61, 525 (1994), arXiv:hep-ph/9310247 .
- Ding et al. [2009] G.-J. Ding, J.-F. Liu, and M.-L. Yan, Dynamics of Hadronic Molecule in One-Boson Exchange Approach and Possible Heavy Flavor Molecules, Phys. Rev. D 79, 054005 (2009), arXiv:0901.0426 [hep-ph] .
- Calle Cordon and Ruiz Arriola [2010] A. Calle Cordon and E. Ruiz Arriola, Renormalization vs Strong Form Factors for One Boson Exchange Potentials, Phys. Rev. C 81, 044002 (2010), arXiv:0905.4933 [nucl-th] .
- Sun et al. [2011] Z.-F. Sun, J. He, X. Liu, Z.-G. Luo, and S.-L. Zhu, and as the and molecular states, Phys. Rev. D 84, 054002 (2011), arXiv:1106.2968 [hep-ph] .
- Zhao et al. [2013] L. Zhao, N. Li, S.-L. Zhu, and B.-S. Zou, Meson-exchange model for the interaction, Phys. Rev. D 87, 054034 (2013), arXiv:1302.1770 [hep-ph] .
- Liu et al. [2018] M.-Z. Liu, T.-W. Wu, J.-J. Xie, M. Pavon Valderrama, and L.-S. Geng, and molecular states from one boson exchange, Phys. Rev. D 98, 014014 (2018), arXiv:1805.08384 [hep-ph] .
- Liu et al. [2019] M.-Z. Liu, T.-W. Wu, M. Pavon Valderrama, J.-J. Xie, and L.-S. Geng, Heavy-quark spin and flavor symmetry partners of the X(3872) revisited: What can we learn from the one boson exchange model?, Phys. Rev. D 99, 094018 (2019), arXiv:1902.03044 [hep-ph] .
- Zhou et al. [2005] Z. Y. Zhou, G. Y. Qin, P. Zhang, Z. Xiao, H. Q. Zheng, and N. Wu, The Pole structure of the unitary, crossing symmetric low energy scattering amplitudes, JHEP 02, 043, arXiv:hep-ph/0406271 .
- Caprini et al. [2006] I. Caprini, G. Colangelo, and H. Leutwyler, Mass and width of the lowest resonance in QCD, Phys. Rev. Lett. 96, 132001 (2006), arXiv:hep-ph/0512364 .
- García-Martín et al. [2011a] R. García-Martín, R. Kamiński, J. R. Peláez, and J. Ruiz de Elvira, Precise determination of the and pole parameters from a dispersive data analysis, Phys. Rev. Lett. 107, 072001 (2011a), arXiv:1107.1635 [hep-ph] .
- Peláez [2016] J. R. Peláez, From controversy to precision on the sigma meson: a review on the status of the non-ordinary resonance, Phys. Rept. 658, 1 (2016), arXiv:1510.00653 [hep-ph] .
- Yao et al. [2021] D.-L. Yao, L.-Y. Dai, H.-Q. Zheng, and Z.-Y. Zhou, A review on partial-wave dynamics with chiral effective field theory and dispersion relation, Rept. Prog. Phys. 84, 076201 (2021), arXiv:2009.13495 [hep-ph] .
- Cao et al. [2023] X.-H. Cao, Q.-Z. Li, Z.-H. Guo, and H.-Q. Zheng, Roy equation analyses of scatterings at unphysical pion masses, Phys. Rev. D 108, 034009 (2023), arXiv:2303.02596 [hep-ph] .
- Rodas et al. [2023] A. Rodas, J. J. Dudek, and R. G. Edwards, Constraining the quark mass dependence of the lightest resonance in QCD (2023), arXiv:2304.03762 [hep-lat] .
- Durso et al. [1980] J. W. Durso, A. D. Jackson, and B. J. Verwest, Models of pseudophysical amplitudes, Nucl. Phys. A 345, 471 (1980).
- Holzenkamp et al. [1989] B. Holzenkamp, K. Holinde, and J. Speth, A Meson Exchange Model for the Hyperon Nucleon Interaction, Nucl. Phys. A 500, 485 (1989).
- Meißner [1991] U.-G. Meißner, Chiral dynamics: Where are the scalars?, Comments Nucl. Part. Phys. 20, 119 (1991).
- Rönchen et al. [2013] D. Rönchen, M. Döring, F. Huang, H. Haberzettl, J. Haidenbauer, C. Hanhart, S. Krewald, U.-G. Meißner, and K. Nakayama, Coupled-channel dynamics in the reactions , Eur. Phys. J. A 49, 44 (2013), arXiv:1211.6998 [nucl-th] .
- Reuber et al. [1996] A. Reuber, K. Holinde, H.-C. Kim, and J. Speth, Correlated and exchange in the baryon baryon interaction, Nucl. Phys. A 608, 243 (1996), arXiv:nucl-th/9511011 .
- Haidenbauer and Meißner [2005] J. Haidenbauer and U.-G. Meißner, The Julich hyperon-nucleon model revisited, Phys. Rev. C 72, 044005 (2005), arXiv:nucl-th/0506019 .
- Donoghue [2006] J. F. Donoghue, Sigma exchange in the nuclear force and effective field theory, Phys. Lett. B 643, 165 (2006), arXiv:nucl-th/0602074 .
- Kim and Kim [2020] H.-J. Kim and H.-C. Kim, and coupling constants for the charmed and beauty mesons, Phys. Rev. D 102, 014026 (2020), arXiv:1912.11622 [hep-ph] .
- Yalikun and Zou [2022] N. Yalikun and B.-S. Zou, Anticharmed strange pentaquarks from the one-boson-exchange model, Phys. Rev. D 105, 094026 (2022), arXiv:2112.06426 [hep-ph] .
- Gasser and Meißner [1991] J. Gasser and U.-G. Meißner, Chiral expansion of pion form factors beyond one loop, Nucl. Phys. B 357, 90 (1991).
- Martin and Spearman [1970] A. D. Martin and T. D. Spearman, Elementary Particle Theory (North-Holland Publishing Co., Amsterdam, 1970).
- Krause [1990] A. Krause, Baryon Matrix Elements of the Vector Current in Chiral Perturbation Theory, Helv. Phys. Acta 63, 3 (1990).
- Frink and Meißner [2004] M. Frink and U.-G. Meißner, Chiral extrapolations of baryon masses for unquenched three flavor lattice simulations, JHEP 07, 028, arXiv:hep-lat/0404018 .
- Oller et al. [2006] J. A. Oller, M. Verbeni, and J. Prades, Meson-baryon effective chiral lagrangians to , JHEP 09, 079, arXiv:hep-ph/0608204 .
- Mai and Meißner [2013] M. Mai and U.-G. Meißner, New insights into antikaon-nucleon scattering and the structure of the , Nucl. Phys. A 900, 51 (2013), arXiv:1202.2030 [nucl-th] .
- Ren et al. [2012] X.-L. Ren, L. S. Geng, J. Martin Camalich, J. Meng, and H. Toki, Octet baryon masses in next-to-next-to-next-to-leading order covariant baryon chiral perturbation theory, JHEP 12, 073, arXiv:1209.3641 [nucl-th] .
- Jacob and Wick [1959] M. Jacob and G. C. Wick, On the General Theory of Collisions for Particles with Spin, Annals Phys. 7, 404 (1959).
- Amoros et al. [2001] G. Amoros, J. Bijnens, and P. Talavera, QCD isospin breaking in meson masses, decay constants and quark mass ratios, Nucl. Phys. B 602, 87 (2001), arXiv:hep-ph/0101127 .
- Omnès [1958] R. Omnès, On the Solution of certain singular integral equations of quantum field theory, Nuovo Cim. 8, 316 (1958).
- García-Martín et al. [2011b] R. García-Martín, R. Kamiński, J. R. Peláez, J. Ruiz de Elvira, and F. J. Ynduráin, The Pion-pion scattering amplitude. IV: Improved analysis with once subtracted Roy-like equations up to 1100 MeV, Phys. Rev. D 83, 074004 (2011b), arXiv:1102.2183 [hep-ph] .
- Ropertz et al. [2018] S. Ropertz, C. Hanhart, and B. Kubis, A new parametrization for the scalar pion form factors, Eur. Phys. J. C 78, 1000 (2018), arXiv:1809.06867 [hep-ph] .
- Donoghue [1996] J. F. Donoghue, Dispersion relations and effective field theory, in Advanced School on Effective Theories (1996) arXiv:hep-ph/9607351 .
- Kang et al. [2014] X.-W. Kang, B. Kubis, C. Hanhart, and U.-G. Meißner, decays and the extraction of , Phys. Rev. D 89, 053015 (2014), arXiv:1312.1193 [hep-ph] .
- Chen et al. [2016] Y.-H. Chen, J. T. Daub, F.-K. Guo, B. Kubis, U.-G. Meißner, and B.-S. Zou, Effect of states on decays, Phys. Rev. D 93, 034030 (2016), arXiv:1512.03583 [hep-ph] .
- Dong et al. [2021d] X.-K. Dong, V. Baru, F.-K. Guo, C. Hanhart, A. Nefediev, and B.-S. Zou, Is the existence of a bound state plausible?, Sci. Bull. 66, 2462 (2021d), arXiv:2107.03946 [hep-ph] .
- Chen [2019] Y.-H. Chen, Chromopolarizability of Charmonium and Final State Interaction Revisited, Adv. High Energy Phys. 2019, 7650678 (2019), arXiv:1901.04126 [hep-ph] .
- Anisovich and Leutwyler [1996] A. V. Anisovich and H. Leutwyler, Dispersive analysis of the decay , Phys. Lett. B 375, 335 (1996), arXiv:hep-ph/9601237 .
- Colangelo et al. [2018] G. Colangelo, S. Lanz, H. Leutwyler, and E. Passemar, Dispersive analysis of , Eur. Phys. J. C 78, 947 (2018), arXiv:1807.11937 [hep-ph] .
- Gell-Mann and Lévy [1960] M. Gell-Mann and M. Lévy, The Axial Vector Current in Beta Decay, Nuovo Cim. 16, 705 (1960).
- Oset et al. [2000] E. Oset, H. Toki, M. Mizobe, and T. T. Takahashi, sigma exchange in the NN interaction within the chiral unitary approach, Prog. Theor. Phys. 103, 351 (2000), arXiv:nucl-th/0011008 .
- Hoferichter et al. [2016] M. Hoferichter, J. Ruiz de Elvira, B. Kubis, and U.-G. Meißner, Roy–Steiner-equation analysis of pion–nucleon scattering, Phys. Rept. 625, 1 (2016), arXiv:1510.06039 [hep-ph] .
- Hoferichter et al. [2015] M. Hoferichter, J. Ruiz de Elvira, B. Kubis, and U.-G. Meißner, Matching pion-nucleon Roy-Steiner equations to chiral perturbation theory, Phys. Rev. Lett. 115, 192301 (2015), arXiv:1507.07552 [nucl-th] .
- de Swart [1963] J. J. de Swart, The Octet model and its Clebsch-Gordan coefficients, Rev. Mod. Phys. 35, 916 (1963), [Erratum: Rev.Mod.Phys. 37, 326–326 (1965)].
- Fettes et al. [1998] N. Fettes, U.-G. Meißner, and S. Steininger, Pion-nucleon scattering in chiral perturbation theory. 1. Isospin symmetric case, Nucl. Phys. A 640, 199 (1998), arXiv:hep-ph/9803266 .
- Fettes et al. [2000] N. Fettes, U.-G. Meißner, M. Mojžiš, and S. Steininger, The Chiral Effective Pion-Nucleon Lagrangian of Order , Annals Phys. 283, 273 (2000), [Erratum: Annals Phys. 288, 249–250 (2001)], arXiv:hep-ph/0001308 .
- Scherer and Schindler [2012] S. Scherer and M. R. Schindler, A Primer for Chiral Perturbation Theory, Vol. 830 (2012).
- Meißner and Rusetsky [2022] U.-G. Meißner and A. Rusetsky, Effective Field Theories (Cambridge University Press, 2022).
- Hoferichter et al. [2023] M. Hoferichter, J. R. de Elvira, B. Kubis, and U.-G. Meißner, Nucleon resonance parameters from Roy-Steiner equations, (2023), arXiv:2312.15015 [hep-ph] .
- Borasoy and Meißner [1997] B. Borasoy and U.-G. Meißner, Chiral Expansion of Baryon Masses and -Terms, Annals Phys. 254, 192 (1997), arXiv:hep-ph/9607432 .
- Ikeda et al. [2012] Y. Ikeda, T. Hyodo, and W. Weise, Chiral SU(3) theory of antikaon-nucleon interactions with improved threshold constraints, Nucl. Phys. A 881, 98 (2012), arXiv:1201.6549 [nucl-th] .
- Guo and Oller [2013] Z.-H. Guo and J. A. Oller, Meson-baryon reactions with strangeness within a chiral framework, Phys. Rev. C 87, 035202 (2013), arXiv:1210.3485 [hep-ph] .
- Hara [1970] Y. Hara, On crossing relations for helicity amplitudes, J. Math. Phys. 11, 253 (1970).
- Hebbar et al. [2022] A. Hebbar, D. Karateev, and J. Penedones, Spinning S-matrix bootstrap in 4d, JHEP 01, 060, arXiv:2011.11708 [hep-th] .