Doubly heavy tetraquarks in a chiral-diquark picture
Abstract
Energy spectrum of doubly heavy tetraquarks, ( with and ), is studied in the potential chiral-diquark model. Using the chiral effective theory of diquarks and the quark-diquark-based potential model, the , , and tetraquarks are described as a three-body system composed of two heavy quarks and an antidiquark. We find several bound states, while no and (deep) bound state is seen. We also study the change of the tetraquark masses under restoration of chiral symmetry.
I Introduction
Hadron spectroscopy is one of the basis to show the properties of the quantum field theory. Hadrons are mostly observed as two types of structures, the quark-antiquark configuration for the conventional mesons and the three-quark configuration for the conventional baryons. However, the Quantum Chromodinamics (QCD) allows some exotic color-singlet hadrons such as compact multiquark states, hadronic molecules, and glueballs. Observation and interpretation of these unconventional states with the special properties have provided us a various attempt for the discussion of hadron physics for a long time.
A recent discovery of exotic hadrons is the tetraquark reported by the LHCb Collaboration Aaij et al. (2021a, b). This is composed of two charm quarks and two light antiquarks, and the total spin and parity is assumed to be . The mass relative to threshold and the decay width were firstly determined as keV and keV with the Breit-Wigner profile Aaij et al. (2021a), and these values have been updated as keV and keV with the unitarized Breit-Wigner profile Aaij et al. (2021b), respectively. On the theoretical points of view, the early studies of exotic hadrons including two heavy quarks, ( and ), have been made in the 1980s Ader et al. (1982); Ballot and Richard (1983); Lipkin (1986); Zouzou et al. (1986); Heller and Tjon (1987); Carlson et al. (1988). Until these days, many researchers attempted to search for the features of doubly heavy tetraquarks with various approaches such as quark-level models Silvestre-Brac and Semay (1993); Semay and Silvestre-Brac (1994); Chow (1995a, b); Pepin et al. (1997); Brink and Stancu (1998); Gelman and Nussinov (2003); Vijande et al. (2004); Janc and Rosina (2004); Vijande et al. (2006); Ebert et al. (2007); Zhang et al. (2008); Vijande et al. (2009); Yang et al. (2009); Feng et al. (2013); Karliner and Rosner (2017); Eichten and Quigg (2017); Yan et al. (2018); Ali et al. (2018); Park et al. (2019); Deng et al. (2020); Caramés et al. (2019); Hernández et al. (2020); Yang et al. (2020a); Bedolla et al. (2020); Yu et al. (2020); Wallbott et al. (2020); Tan et al. (2020); Lü et al. (2020); Braaten et al. (2021); Yang et al. (2020b); Qin et al. (2021); Meng et al. (2021, 2022); Chen (2021); Jin et al. (2021); Chen et al. (2021); Andreev (2021); Deng and Zhu (2021), chromomagnetic interaction models Lee et al. (2008); Lee and Yasui (2009); Hyodo et al. (2013); Luo et al. (2017); Hyodo et al. (2017); Cheng et al. (2020, 2021); Weng et al. (2022); Guo et al. (2022), QCD sum rules Navarra et al. (2007); Dias et al. (2011); Du et al. (2013); Chen et al. (2014); Wang (2018); Wang and Yan (2018); Agaev et al. (2019); Sundu et al. (2019); Agaev et al. (2020a); Tang et al. (2020); Agaev et al. (2020b, c); Wang and Chen (2020); Agaev et al. (2021a, b, 2022); Azizi and Özdem (2021); Aliev et al. (2021); Özdem (2022); Bilmiş (2021), lattice QCD simulations Detmold et al. (2007); Wagner (2010); Bali and Hetzenegger (2010); Bicudo and Wagner (2013); Brown and Orginos (2012); Ikeda et al. (2014); Bicudo et al. (2015, 2016); Francis et al. (2017); Bicudo et al. (2017a, b); Francis et al. (2019); Junnarkar et al. (2019); Leskovec et al. (2019); Hudspith et al. (2020); Mohanta and Basak (2020); Bicudo et al. (2021); Padmanath and Mathur (2021) and so on. Most of them predict a stable doubly-bottom tetraquark below the strong-decay threshold, but no stable doubly charmed tetraquark. The predictions for the bottom-charmed tetraquark are divided, whether it is stable or not.
Multiquark states are often considered from the perspective of their substructures (clusters). The tetraquarks are often discussed as diquark-antidiquark or meson-meson structure. In this paper, we apply the chiral effective theory of diquarks and the quark-diquark potential model Harada et al. (2020); Kim et al. (2020, 2021). In Ref. Harada et al. (2020), we proposed a chiral effective theory of the spin 0, color , scalar/pseudoscalar diquarks based on the chiral symmetry. Here, the diquarks belong to the same chiral multiplet, forming chiral partners and their masses are to be degenerate when the chiral symmetry is restored. Similarly, in Ref. Kim et al. (2021), the spin 1, color , vector/axial-vector diquarks are introduced as another set of chiral partners. Then we applied the chiral effective theories to singly heavy baryons and studied the spectrum using the nonrelativistic potential heavy-quark–diquark model.
In the present work, tetraquarks are described as a three-body system composed of two heavy quarks and an antidiquark. We calculate the mass spectrum and the wave functions to study how chiral symmetry is realized in the tetraquark systems. We also investigate the dependence of tetraquark masses on the chiral symmetry breaking parameter.
II Theoretical Framework
In this section, firstly, we introduce the mass formulas of diquarks given by the chiral effective theory of scalar/pseudoscalar diquarks Harada et al. (2020); Kim et al. (2020) and of vector/axial-vector diquarks Kim et al. (2021). Then we construct the nonrelativistic potential model for the tetraquark states including the antidiquark cluster.
II.1 Chiral effective theory of diquarks
Following Ref. Kim et al. (2021), we consider the chiral effective theory of diquarks given by the chiral effective Lagrangian,
(1) |
Here the first term describes the effective Lagrangian for the scalar and pseudoscalar diquarks, and for the vector and axial-vector diquarks. Their explicit forms are given below. The last term, , describes the coupling between the scalar and vector diquarks. As it does not contribute to the diquark masses, here we omit this term. The third and forth terms of Eq. (1) are the kinetic and potential terms for the meson field , respectively. This meson operator contains nonet scalar and pseudoscalar mesons, whose chiral transform is given by
(2) |
The potential term for the meson field, , causes spontaneous chiral symmetry breaking, represented by the vacuum expectation value of scalar meson as
(3) |
where MeV is the pion decay constant. In this vacuum, meson is regarded as the massless Nambu-Goldstone bosons.
Using the meson field and the chiral diquark operators summarized in Table. 1, the effective Lagrangians, and , are expressed as
(4) |
(5) |
Chiral operator | Spin | Color | Flavor |
---|---|---|---|
Since the diquark is not a color-singlet state, we introduce the color-gauge-covariant derivative, , with being the gluon field and being the color generator for the color representation. It is used for the kinetic term of these Lagrangians. in the first term of Eq. (5) shows the strength of chiral vector diquark fields. The self-interaction terms of gluons are omitted, and all the color indices are contracted and not explicitly written.
The masses of scalar and pseudoscalar diquarks are given by the three parameters , , and in Eq. (4). is the chiral invariant mass which is the degenerate mass for the two diquarks in the chiral symmetry restored phase. Mass splitting between diquarks induced by the chiral symmetry breaking is given by and . In particular, the term also causes the symmetry breaking ’t Hooft (1976, 1986). On the other hand, the masses of vector and axial-vector diquarks are given by three parameters , , and in Eq. (5). Similar to the scalar and pseudoscalar diquarks, is the chiral invariant mass of vector and axial-vector diquarks, and their mass splitting is given by and .
The non-zero strange quark mass brings about the explicit chiral symmetry breaking and the flavor symmetry breaking. In this term, the masses of scalar and pseudoscalar diquarks, and , are expressed in Ref. Harada et al. (2020) as
(6) | |||||
(7) | |||||
(8) | |||||
(9) |
where the index stands for or quark, and for quark, meaning two constituent quarks of a diquark. The constant is given by
(10) |
with , where , and stands for the mass of strange quark, the quark-meson coupling constant, and the decay constants of pion and kaon, respectively. From these mass formulas, Eqs. (6)–(9), we can give the mass relation of scalar and pseudoscalar diquarks Harada et al. (2020); Kim et al. (2020, 2021) as
(11) | |||||
From Eq. (11), we see the non-strange pseudoscalar diquark is heavier than the singly strange pseudoscalar diquark, , which is called the inverse mass hierarchy.
The masses of axial-vector and vector diquarks, and , are expressed in Ref. Kim et al. (2021) as
(12) | |||||
(13) | |||||
(14) | |||||
(15) | |||||
(16) | |||||
For the masses of axial-vector diquarks, Eqs. (12)–(14), the generalized Gell-Mann–Okubo mass formula Kim et al. (2021) which is analogous to the conventional one Gell-Mann (1962); Okubo (1962) is obtained as
(17) | |||||
Here the square mass differences are characterized by the number of quark. On the other hand, the square mass difference between nonstrange and singly strange vector diquarks is given by
(18) |
As shown in Table. 2, the parameter generally takes the negative value Kim et al. (2021), so that the mass difference between nonstrange and singly strange vector diquarks becomes much larger than that of axial-vector diquarks. Here we call this the enhanced mass hierarchy of vector diquarks in this paper.
II.2 Potential models
In order to calculate the spectrum of tetraquarks, we use the chiral effective theory of diquark Harada et al. (2020); Kim et al. (2020, 2021) and consider these tetraquarks as the three-body system with two heavy quarks () and one antidiquark (). We need two kinds of two-body potentials, one for the interaction between two heavy quarks, , and the other for between a heavy quark and a color antidiquark, .
The Hamiltonian for this tetraquark system is written as
(19) | |||||
where and are the mass and momentum of constituent particles, respectively. is the kinetic energy of the center of mass, which is subtracted and the remained kinetic energy is for the and parts of Jacobi coordinates in Fig. 1.

For the potential between two heavy-quarks, , we choose the form of the “ potential” in Ref. Silvestre-Brac (1996), which reproduces the spectra of heavy mesons, color-singlet . The potential contains the power-law confinement term and one-gluon-exchange term expressed as
(20) | |||||
with the mass-dependent parameter
(21) |
The parameters in Eqs. (20) and (21) are summarized in Table. 3 with the values of effective quark masses. In addition, the calculated masses of singly and doubly heavy mesons are also summarized for the verification of this potential model.
Masses () | ||||
---|---|---|---|---|
Parameters () | Mesons | Silvestre-Brac (1996) | Exp. Zyla et al. (2020) | |
1881 | 1868.04 | |||
(GeV5/3) | 2033 | 2009.12 | ||
(GeV) | 1955 | 1968.34 | ||
2107 | 2112.20 | |||
(GeVB-1) | 5311 | 5279.44 | ||
5367 | 5324.70 | |||
(GeV) | 5356 | 5366.88 | ||
(GeV) | 5418 | 5415.40 | ||
(GeV) | 2982 | 2983.90 | ||
(GeV) | 3103 | 3096.90 | ||
6269 | 6274.90 | |||
9401 | 9398.70 | |||
9461 | 9460.30 |
Here we consider the effect of color structure expressed as by the Gell-Mann matrices. As in Table. 4, the value of this factor for the quark-antiquark picture inside a meson () is two times larger than that for the two quarks with color representation (). For this reason, we assume that the potential in Eq. (20) can be generalized to the form proportional to this factor, , and we write the potential between two heavy-quarks of color as
(22) |
Two particles | Color structure | |
---|---|---|
, | ||
, |
Similarly, to construct the potential for between a heavy quark and a color antidiquark, , we apply the potential diquark–heavy-quark model which gives the spectra of singly heavy baryons. For this type of potential model, we employ the “Y-potential” in the chiral effective theory of diquarks Kim et al. (2021), which is firstly made from the quark model as in Refs. Yoshida et al. (2015); Kim et al. (2020). Similar to the Eq. (20), this is expressed with the confinement term and one-gluon-exchange term as
(23) | |||||
where the parameters are summarized in Table. 5 with the calculated masses of singly heavy baryons. Note that we have adjusted the constant shift parameters from the original values in Ref. Kim et al. (2021), so as to use the same heavy-quark mass values with the potential model summarized in Table. 3. For the other parameters, we use the values of diquark masses summarized in Table. 2. Although the Y-potential includes the spin-orbit term and the tensor term Kim et al. (2021), we here neglect them for simplicity.
Now we consider the color structure and the factor again. From the value of summarized in Table. 4 and the assumption that the Y-potential is proportional to the factor , , we can write the potential between the heavy quark and the color antidiquark inside the color-singlet tetraquark as
(24) |
Masses () | ||||
---|---|---|---|---|
Parameters (Y-pot.) | Baryons | Y-pot. Kim et al. (2021) | Exp. Zyla et al. (2020) | |
() | 2284 | 2286.46 | ||
(GeV2) | () | 2451 | 2453.54 | |
(GeV) | () | 2513 | 2518.13 | |
(GeV) | () | 2467 | 2469.42 | |
() | 2581 | 2578.80 | ||
() | 2639 | 2645.88 | ||
(GeV) | () | 2699 | 2695.20 | |
() | 2752 | 2765.90 | ||
() | 5619 | 5619.60 | ||
() | 5810 | 5813.10 | ||
() | 5829 | 5832.53 | ||
() | 5796 | 5794.45 | ||
() | 5934 | 5935.02 | ||
() | 5952 | 5953.82 | ||
() | 6046 | 6046.10 | ||
() | 6063 |
To solve the Schrdinger equation for the Hamiltonian in Eq. (19), we use the Gaussian expansion method Kamimura (1988); Hiyama et al. (2003). Here the variational wave function of tetraquarks with the total spin is expressed as
(25) | |||||
where and stand for the spin wave function of the heavy quark and the antidiquark, respectively. and are the Gaussian basis functions, which show the spatial wave function for and parts of Jacobi coordinates with the coefficients . The index denotes the quantum numbers related to this calculation method, , where and are the number of basis functions, and are the orbital angular momentum, and are the spin of two heavy quarks and all constituent particles, respectively. Using the Gaussian expansion method, the coefficients are determined by the Rayleigh-Ritz variational principle.
In calculating the spectrum of tetraquarks with the antidiquark cluster, we only consider -wave and -wave for the orbital angular momentum. For the -wave states, we calculate the masses of states which the principal quantum number or . For the -wave excited states, we classify these states into three types based on the channel of the Jacobi coordinate in Fig. 1: the -mode which is the excitation between two heavy quarks, the -mode which is that between the pair of two heavy quarks and an antidiquark, and the -mode which is that between two antiquarks inside the pseudoscalar or vector antidiquark. In particular, we calculate the states summarized in Tables. 6 and 7, which are represented as the spectra introduced in the next Sec. III. Note that, for the and tetraquarks, the quamtum numbers are restricted by the Pauli principle to for the -mode states and for the -wave and -mode states.
State | Antidiquark () | |||||||
---|---|---|---|---|---|---|---|---|
1 | 0 | 0 | Scalar | 0 | 0 | * | ||
1 | 0 | 0 | Scalar | 1 | 1 | |||
2 | 0 | 0 | Scalar | 0 | 0 | * | ||
2 | 0 | 0 | Scalar | 1 | 1 | |||
1 | 1 | 0 | Scalar | 0 | 0 | |||
1 | 1 | 0 | Scalar | 1 | 1 | * | ||
1 | 0 | 1 | Scalar | 0 | 0 | * | ||
1 | 0 | 1 | Scalar | 1 | 1 | |||
1 | 0 | 0 | Axial-vector | 0 | 1 | * | ||
1 | 0 | 0 | Axial-vector | 1 | 0 | |||
1 | 0 | 0 | Axial-vector | 1 | 1 | |||
1 | 0 | 0 | Axial-vector | 1 | 2 | |||
2 | 0 | 0 | Axial-vector | 0 | 1 | * | ||
2 | 0 | 0 | Axial-vector | 1 | 0 | |||
2 | 0 | 0 | Axial-vector | 1 | 1 | |||
2 | 0 | 0 | Axial-vector | 1 | 2 | |||
1 | 1 | 0 | Axial-vector | 0 | 1 | |||
1 | 1 | 0 | Axial-vector | 1 | 0 | * | ||
1 | 1 | 0 | Axial-vector | 1 | 1 | * | ||
1 | 1 | 0 | Axial-vector | 1 | 2 | * | ||
1 | 0 | 1 | Axial-vector | 0 | 1 | * | ||
1 | 0 | 1 | Axial-vector | 1 | 0 | |||
1 | 0 | 1 | Axial-vector | 1 | 1 | |||
1 | 0 | 1 | Axial-vector | 1 | 2 |
State | Antidiquark () | |||||||
---|---|---|---|---|---|---|---|---|
1 | 0 | 0 | Pseudoscalar | 0 | 0 | * | ||
1 | 0 | 0 | Pseudoscalar | 1 | 1 | |||
1 | 1 | 0 | Pseudoscalar | 0 | 0 | |||
1 | 1 | 0 | Pseudoscalar | 1 | 1 | * | ||
1 | 0 | 1 | Pseudoscalar | 0 | 0 | * | ||
1 | 0 | 1 | Pseudoscalar | 1 | 1 | |||
1 | 0 | 0 | Vector | 0 | 1 | * | ||
1 | 0 | 0 | Vector | 1 | 0 | |||
1 | 0 | 0 | Vector | 1 | 1 | |||
1 | 0 | 0 | Vector | 1 | 2 | |||
1 | 1 | 0 | Vector | 0 | 1 | |||
1 | 1 | 0 | Vector | 1 | 0 | * | ||
1 | 1 | 0 | Vector | 1 | 1 | * | ||
1 | 1 | 0 | Vector | 1 | 2 | * | ||
1 | 0 | 1 | Vector | 0 | 1 | * | ||
1 | 0 | 1 | Vector | 1 | 0 | |||
1 | 0 | 1 | Vector | 1 | 1 | |||
1 | 0 | 1 | Vector | 1 | 2 |
III Results
III.1 Spectrum of and tetraquarks




Here we discuss the spectra of the and tetraquarks, in which two constituent heavy quarks are the same. The energy spectra of these tetraquarks are shown in Figs. 3, 3, 5, and 5. Each figure is classified by the flavor of constituent antidiquarks and the heavy quarks.
Figs. 3 and 3 are the spectra of the strangeness and tetraquarks, respectively. Here the flavor of all the constituent antidiquarks is , meaning the scalar, pseudoscalar, and vector antidiquarks. In particular, the red and green lines stand for the masses of tetraquarks including the scalar antidiquark, which correspond to the -wave states and the -wave states, respectively. For the other colors, the blue and magenta lines, indicate the tetraquarks including the pseudoscalar antidiquark and the vector antidiquark, respectively.
About the spectra in Fig. 3, all the tetraquarks including the scalar antidiquarks are lighter than those containing the pseudoscalar or vector antidiquarks. On the other hand, in Fig. 3, not all the tetraquarks including the scalar antidiquarks are lighter than the others. In detail, when a constituent antidiquark has an strange antiquark,, the masses of -mode states become lighter than those of states and approach to the lightest -wave states (- and -mode states). This is caused by the inverse mass hierarchy of pseudoscalar diquarks in Eq. (11). Moreover, all the tetraquarks containing the -mode state become heavier than the other states by the inclusion of a strange antiquark, and this is caused by the enhanced mass hierarchy of vector diquarks in Eq. (18).
The thresholds for the strong decay of the ground states is given by the sum of and singly heavy mesons ( or ). Note that the ground state is not allowed to decay into two mesons. In Figs. 3 and 3, we show two sets of threshold values, one (exp.) given by the experimental meson masses in the Particle Data Group Zyla et al. (2020) (black-dashed lines), and another (AP1) from the calculation with the potential in Eqs. (20) and (21) (blue-dotted lines). One sees that the ground states of tetraquarks are below the thresholds, so that they are bound states and do not decay by the strong interaction. From the experimental thresholds, their binding energies are MeV for the and MeV for the . On the other hand, no bound state appears in the spectra of tetraquarks.
Next we discuss the spectra of the tetraquarks containing the axial-vector antidiquarks. These are illustrated in Figs. 5 and 5, which show the spectrum of and tetraquarks, respectively. The red and green lines stand for the -wave states and the -wave states, respectively. In these figures, the number of strange antiquark is increasing from left to right.
Comparing three spectra in each figure, the masses of these tetraquarks in each state become heavier as the number of the strange antiquark increases, and the mass difference between and is almost equal to that between and . This behavior comes from the generalized Gell-Mann–Okubo mass formula for the axial-vector diquarks in Eq. (17).
The thresholds shown in Figs. 5 and 5 are for the tetraquarks, and are the total mass of two singly heavy mesons with . Although the ground state with are the lightest, they are above the thresholds and therefore unstable. Discrepancy between the tetraquark mass and the threshold is even larger for the strange tetraquarks.
About the spin-spin potential terms in the two potential models, Eqs. (20) and (23), both of them are in inverse proportion to the masses of two interacting particles. Therefore, in the heavy-quark limit (), these terms disappear, and the splitted states by the spin-spin interaction become degenerate. This is called the heavy-quark spin (HQS) multiplet which appears frequently in the tetraquark spectra. For example, in Figs. 5 and 5, the and states of tetraquarks with , and have the HQS triplet structure. Comparing these structures in Figs. 5 and 5, we can see that the mass differences for tetraquarks are smaller than those for tetraquarks, which is caused by the bottom quark being heavier than the charm quark.
III.2 Spectrum of tetraquarks




We here discuss the spectra of tetraquarks. Since two constituent heavy quarks are different from each other, the spin of two heavy quarks, , can take two kinds of values as and . The energy spectra of these tetraquarks are shown in Figs. 7, 7, 9, and 9, classified by the flavor of constituent antidiquarks and the quantum number . Figs. 7 and 7 are the spectra of the strangeness and tetraquarks with a flavor antidiquark, respectively. The four colors of lines stand for the types of states and constituent antidiquarks, which is the same as in Figs. 3 and 3. Also, Figs. 9 and 9 are the spectra of the tetraquarks with and , including a flavor axial-vector antidiquark, respectively. The two colors of lines stand for the -wave states and the -wave states, which is the same as Figs. 5 and 5.
The spectra of tetraquarks are similar to those of and tetraquarks when the constituent antidiquarks are the same. Although there are less states for the tetraquarks with , we can see the effect of chiral effective theory of diquarks: the inverse mass hierarchy of pseudoscalar diquarks and the enhanced mass hierarchy of vector diquarks in Figs. 7 and 7, and the generalized Gell-Mann–Okubo mass formula for the axial-vector diquarks in Figs. 9 and 9.
In the states given in Tables. 6 and 7, one sees that some -wave states have the same spin quantum numbers, such as and modes with or 1. These states may be mixed by the spin-spin interaction in . (Note that the mixing is not allowed in and due to the Pauli principle.) The subscripts of and characters mean the probabilities of each state, calculated by the two-body density distribution between two heavy quarks. For example, about two -wave states with green lines in Fig. 7, the - and -mode states mix each other. As a result, in the lower state, the probability of the -mode state becomes , and the remained is that of the -mode state.
As with the and spectra, the black-dashed (exp.) and the blue-dotted (AP1) lines in Figs. 7, 7, 9, and 9 are the relevant two-meson thresholds. We find no bound tetraquarks except the tetraquark with at 3 MeV below the AP1 threshold in Fig. 7. Therefore, we conclude that most of the tetraquarks with the antidiquark cluster are unstable in strong decays.
III.3 Binding energies of lowest tetraquark states
Tetraquark | State | Binding energy [MeV] | |
---|---|---|---|
type exp | type AP1 | ||
We summarize the binding energies of the lowest-lying states in Table 8. Two lowest states of tetraquarks are well below the threshold and expected to be stable.
On the other hand, we have no bound state. It may seem contradictory to the recent observation of , at about 300 keV below the threshold by the LHCb collaboration Aaij et al. (2021a, b). As the observed state is very close to the threshold, it may couple strongly to the state and thus have a molecule-like structure. In contrast, the tetraquark in our diquark picture does not dissociate into , and therefore it does not couple to loosely bound states. In other words, our model can reproduce only tightly bound tetraquark states. Thus our results may not be inconsistent with the LHCb observation.
In Table. 9, we also summarize the stability of lowest tetraquark states given by some previous works. Most of the studies in this table predicted that the non-strange and singly antistrange tetraquarks, and , are stable, which is consistent with our results. On the other hand, the other tetraquark systems seem to be unstable or uncertain.
Reference | |||||||||
Quark-level models | |||||||||
Lipkin (1986) Lipkin (1986) | ND | S | |||||||
ZouZou et al. (1986) Zouzou et al. (1986) | S | S | S | ||||||
Carlson-Heller-Tjon (1988) Carlson et al. (1988) | S | S | ND | ||||||
Semay–Silvestre-Brac (1994) Semay and Silvestre-Brac (1994) | ND | S | S | ND | |||||
Pepin et al. (1997) Pepin et al. (1997) | S | S | |||||||
Janc-Rosina (2004) Janc and Rosina (2004) | S | S | |||||||
Ebert et al. (2007) Ebert et al. (2007) | US | US | US | S | US | US | US | US | US |
Zhang-Zhang-Zhang (2008) Zhang et al. (2008) | US | US | S | US | |||||
Vijande-Valcarce-Barnea (2009) Vijande et al. (2009) | S | S | |||||||
Yang et al. (2009) Yang et al. (2009) | ND | S | |||||||
Feng-Guo-Zou (2013) Feng et al. (2013) | S | S | S | ||||||
Karliner-Rosner (2017) Karliner and Rosner (2017) | US | S | ND | ||||||
Eichten-Quigg (2017) Eichten and Quigg (2017) | US | US | US | S | S | US | US | US | US |
Park-Noh-Lee (2019) Park et al. (2019) | US | S | S | US | |||||
Deng-Chen-Ping (2020) Deng et al. (2020) | S | US | US | S | S | US | S | US | US |
Yang-Ping-Segovia (2020a) Yang et al. (2020a) | S | S | S | ||||||
Tan-Lu-Ping (2020) Tan et al. (2020) | S | S | S | ||||||
Lü-Chen-Dong (2020) Lü et al. (2020) | US | US | US | S | US | US | US | US | US |
Braaten-He-Mohapatra (2021) Braaten et al. (2021) | US | US | US | S | S | US | US | US | US |
Yang-Ping-Segovia (2020b) Yang et al. (2020b) | US | US | US | ||||||
Meng et al. (2021) Meng et al. (2021) | S | S | S | S | |||||
(This work) | US | US | US | S | S | US | ND | US | US |
Chromomagnetic interaction models | |||||||||
Lee-Yasui (2009) Lee and Yasui (2009) | S | S | S | S | S | US | |||
Luo et al. (2017) Luo et al. (2017) | S | S | US | S | S | US | S | S | US |
Cheng et al. (2020) Cheng et al. (2020) | ND | US | |||||||
Cheng et al. (2021) Cheng et al. (2021) | US | US | US | S | S | US | ND | US | US |
Weng-Deng-Zhu (2022) Weng et al. (2022) | S | ND | US | S | S | ND | S | ND | ND |
Guo et al. (2022) Guo et al. (2022) | S | S | US | S | S | ND | S | S | US |
QCD sum rules | |||||||||
Navarra-Nielsen-Lee (2007) Navarra et al. (2007) | ND | S | |||||||
Du et al. (2013) Du et al. (2013) | US | US | S | S | S | ||||
Chen-Steele-Zhu (2014) Chen et al. (2014) | ND | S | |||||||
Wang (2018) Wang (2018) | ND | ND | S | S | |||||
Wang-Yan (2018) Wang and Yan (2018) | ND | ND | ND | ||||||
Agaev et al. (2019) Agaev et al. (2019) | S | ||||||||
Agaev-Azizi-Sundu (2020) Agaev et al. (2020a) | ND | ||||||||
Agaev et al. (2020, 2021) Agaev et al. (2020b, 2021a) | S | ||||||||
Wang-Chen (2020) Wang and Chen (2020) | S | ||||||||
Lattice QCD | |||||||||
Francis et al. (2017) Francis et al. (2017) | S | S | |||||||
Francis et al. (2019) Francis et al. (2019) | S | S | S | ||||||
Junnarkar-Mathur-Padmanath (2019) Junnarkar et al. (2019) | S | S | US | S | S | ND | |||
Leskovec et al. (2019) Leskovec et al. (2019) | S | ||||||||
Hudspith et al. (2020) Hudspith et al. (2020) | S | S | US | S | |||||
Mohanta-Basak (2020) Mohanta and Basak (2020) | S |
III.4 Density distribution of bound states
Now we discuss the density distributions between two constituent particles in each of the tetraquark bound states. The two-body density distribution is expressed by the variational wave function in Eq. (25) as
(26) |
where and are the distance and the angular parts of the relative coordinate for two particles in channel , respectively. From this equation (26), we can give the density distribution between two bottom quarks with channel , and that between a bottom quark and a scalar antidiquark with channel . The functions of each pair are shown in Fig. 10 for both tetraquarks with the strangeness and .

We see that the distance between the pair of two bottom quarks () is nearer than the pair of a bottom quark and an antidiquark () for both strangeness and . This is due to the suppression of kinetic energy which is inversely proportional to the reduced mass of two interacting particles. Comparing with the different strangeness , the red solid line and the blue dashed line are almost same so that there is no difference between the pair . On the other hand, for between the pair , the nonstrange antidiquark seems to be more extended from a bottom quark than that including one strange antiquark, because the scalar diquark becomes heavier by including a strange quark.
Tetraquark | State | rms distance [fm] | |
---|---|---|---|
0.30 | 0.56 | ||
0.30 | 0.53 |
Also we summarize the root-mean-square (rms) distances of between two particles in Table. 10, which is given by the density distribution function as
(27) |
For the rms distance between a bottom quark and an antidiquark (), these values are similar to that between a bottom quark and a diquark inside the ground state of ( fm) and ( fm) Kim et al. (2020). Although the interaction between a heavy quark and an antidiquark is half as large as that between a heavy quark and a diquark from Eq. (24), its effect is mostly suppressed by the interaction between two bottom quarks.
III.5 tetraquarks toward chiral restoration
In the chiral effective theory based on the linear sigma model, the spontaneous chiral symmetry breaking is controlled by the vacuum expectation value of . It is interesting to see how the tetraquark masses are modified when the value of changes. Such a situation may be realized in nature when the tetraquarks are placed or produced in the hot/dense matter, where chiral symmetry tends to be restored. According to Ref. Kim et al. (2021), when we multiply the chiral symmetry breaking factor to , where its range is , we can express the masses of diquarks from the chiral symmetry restored phase () to the ordinary vacuum state ().
For example, the masses of nonstrange diquarks are given by
(28) | |||
(29) | |||
(30) | |||
(31) |
in which only the mass of scalar diquark becomes heavier and the others become lighter with approaching to the chiral symmetry restored phase .




Figs. 11, 12, 13, and 14 show the change of masses of tetraquarks by the factor , which are all ground states including a scalar or an axial-vector antidiquark. For all tetraquarks in these figures, these masses change, which is similar to the mass of an antidiquark they include. As the masses of scalar and axial-vector diquarks cross at around and are reversed, the crossing of masses of tetraquarks also occurs near . The tetraquarks including an axial-vector antidiquark become lighter than those including a scalar antidiquark. In other words, the isospin of stable tetraquarks changes from to as the chiral symmetry is restored. In particular, if we assume that meson-meson thresholds do not change by the degree of chiral symmetry breaking, tetraquarks with a constituent axial-vector antidiquark are below the thresholds when the parameter is less than .
In the chiral symmetry restored phase, the masses of nonstrange vector diquark and axial-vector diquark are degenerate with the mass value according to Eqs. (30) and (31). Therefore, in the chiral symmetry restored phase, not only the positive-parity states but also the negative-parity states with the -mode are below the thresholds for the nonstrange tetraquark. On the other hand, the masses of nonstrange pseudoscalar diquark and scalar diquark are not degenerate due to the anomaly, and the pseudoscalar diquark is about 300 MeV heavier than the scalar diquark even at Kim et al. (2021). Thus there are no -mode state below the thresholds for the nonstrange tetraquark.
IV Summary
In this paper, we have studied the spectrum of doubly heavy tetraquarks in the potential chiral-diquark model. Applying the chiral effective theory of diquarks and the potential quark-diquark model Harada et al. (2020); Kim et al. (2020, 2021), we have calculated the lower energy spectrum and the wave functions of -wave and -wave tetraquarks. Here we have employed the Gaussian expansion method Kamimura (1988); Hiyama et al. (2003) for solving the three-body system consisting of two heavy quarks and one antidiquark. We have also investigated the behavior of the ground-state energies of the nonstrange tetraquarks under the chiral restoration.
For the tetraquarks including an antidiquark, we have obtained the following:
-
(i)
For the ground states of the tetraquarks, we have found that the tetraquarks including a scalar antidiquark are stable in strong decays with the binding energies, 115 MeV for and 28 MeV for . No other states, such as and tetraquarks, are significantly below the two-meson thresholds and thus not stable.
-
(ii)
For the excited states of the tetraquarks, we have found the inverse mass hierarchy in states with the -mode and the enhanced mass hierarchy in states with the -mode. As a result, for the singly antistrange tetraquarks, there are one more -wave states between and states compared with the nonstrange tetraquarks.
-
(iii)
For two stable bound states of tetraquarks, we have investigated the structure by calculating the density distributions and rms distances from the wave functions. We found that the size of nonstrange and singly antistrange tetraquarks are similar to each other and have less difference by the existence of an strange antiquark.
-
(iv)
For the ground states of the nonstrange tetraquarks, we have investigated the change of masses by the degree of chiral symmetry breaking. Here we used the mass formulas of non-strange diquarks expressed with the chiral symmetry breaking factor , Eqs. (28) – (31). We found that the axial-vector antidiquark makes the tetraquarks more stable than the scalar antidiquark near the chiral symmetry restored phase.
Finally, we comment on possible improvements of our model. In this work, we have treated the antidiquark as a point-like particle, where the distance between two light antiquarks inside the antidiquark is neglected. It will be important to take into account the finite size of diquarks and improve the heavy-quark–diquark potential model, as done by Refs. Kumakawa and Jido (2017, 2021). In such a model, the parameters of the improved heavy-quark–diquark potential model will be fixed to reproduce the low-lying masses of the singly heavy baryons, and the results of the excited tetraquark states would be improved.
The chiral effective field theory constructed in Refs. Harada et al. (2020); Kim et al. (2020, 2021), includes only the scalar, axial-vector, pseudoscalar, and vector diquarks. As summarized in Table. 1, all the diquarks considered in this work have color , so that the energy spectra of tetraquarks including such a diquark are related only to the color interaction between the light antidiquark and the heavy diquark. On the other hand, another type is the color diquark. If we consider the color diquarks in the chiral effective theory of diquarks, the energy spectra of tetraquarks is related also to the color interaction. According to Refs. Deng et al. (2020); Lü et al. (2020); Chen (2021); Hyodo et al. (2013); Luo et al. (2017); Hyodo et al. (2017); Cheng et al. (2020, 2021); Weng et al. (2022); Guo et al. (2022), the energy spectra of tetraquarks including the color antidiquark would appear above the ground states given by the color scalar antidiquarks. Also, tensor diquarks with color and flavor , or with color and flavor , would be the new source for the excited states of tetraquarks. Such studies are left for future works.
Acknowledgments
We would like to thank Yan-Rui Liu for helpful discussions. This work was supported by Grants-in-Aid for Scientific Research No. JP20K03959 (M.O.), No. JP21H00132 (M. O.), No. JP17K14277 (K. S.), and No. JP20K14476 (K. S.), and for JSPS Fellows No. JP21J20048 (Y. K.) from Japan Society for the Promotion of Science.
References
- Aaij et al. (2021a) Roel Aaij et al. (LHCb), “Observation of an exotic narrow doubly charmed tetraquark,” (2021a), arXiv:2109.01038 [hep-ex] .
- Aaij et al. (2021b) Roel Aaij et al. (LHCb), “Study of the doubly charmed tetraquark ,” (2021b), arXiv:2109.01056 [hep-ex] .
- Ader et al. (1982) J. P. Ader, J. M. Richard, and P. Taxil, “Do narrow heavy multiquark states exist?” Phys. Rev. D 25, 2370 (1982).
- Ballot and Richard (1983) J. l. Ballot and J. M. Richard, “Four quark states in additive potentials,” Phys. Lett. B 123, 449–451 (1983).
- Lipkin (1986) Harry J. Lipkin, “A model-independent approach to multiquark bound states,” Phys. Lett. B 172, 242–247 (1986).
- Zouzou et al. (1986) S. Zouzou, B. Silvestre-Brac, C. Gignoux, and J. M. Richard, “Four-quark bound states,” Z. Phys. C 30, 457 (1986).
- Heller and Tjon (1987) L. Heller and J. A. Tjon, “On the Existence of Stable Dimesons,” Phys. Rev. D 35, 969 (1987).
- Carlson et al. (1988) J. Carlson, L. Heller, and J. A. Tjon, “Stability of Dimesons,” Phys. Rev. D 37, 744 (1988).
- Silvestre-Brac and Semay (1993) B. Silvestre-Brac and C. Semay, “Systematics of systems,” Z. Phys. C 57, 273–282 (1993).
- Semay and Silvestre-Brac (1994) C. Semay and B. Silvestre-Brac, “Diquonia and potential models,” Z. Phys. C 61, 271–275 (1994).
- Chow (1995a) Chi-Keung Chow, “Semileptonic decays of heavy tetraquarks,” Phys. Rev. D 51, 3541–3543 (1995a), arXiv:hep-ph/9411221 .
- Chow (1995b) Chi-Keung Chow, “From tetraquarks to hexaquarks: A Systematic study of heavy exotics in the large limit,” Phys. Rev. D 51, 6327–6331 (1995b), arXiv:hep-ph/9412242 .
- Pepin et al. (1997) S. Pepin, F. Stancu, M. Genovese, and J. M. Richard, “Tetraquarks with colour-blind forces in chiral quark models,” Phys. Lett. B 393, 119–123 (1997), arXiv:hep-ph/9609348 .
- Brink and Stancu (1998) D. M. Brink and Fl. Stancu, “Tetraquarks with heavy flavors,” Phys. Rev. D 57, 6778–6787 (1998).
- Gelman and Nussinov (2003) Boris A. Gelman and Shmuel Nussinov, “Does a narrow tetraquark state exist?” Phys. Lett. B 551, 296–304 (2003), arXiv:hep-ph/0209095 .
- Vijande et al. (2004) J. Vijande, F. Fernandez, A. Valcarce, and B. Silvestre-Brac, “Tetraquarks in a chiral constituent quark model,” Eur. Phys. J. A 19, 383 (2004), arXiv:hep-ph/0310007 .
- Janc and Rosina (2004) D. Janc and M. Rosina, “The molecular state,” Few Body Syst. 35, 175–196 (2004), arXiv:hep-ph/0405208 .
- Vijande et al. (2006) J. Vijande, A. Valcarce, and K. Tsushima, “Dynamical study of bf mesons,” Phys. Rev. D 74, 054018 (2006), arXiv:hep-ph/0608316 .
- Ebert et al. (2007) D. Ebert, R. N. Faustov, V. O. Galkin, and W. Lucha, “Masses of tetraquarks with two heavy quarks in the relativistic quark model,” Phys. Rev. D 76, 114015 (2007), arXiv:0706.3853 [hep-ph] .
- Zhang et al. (2008) M. Zhang, H. X. Zhang, and Z. Y. Zhang, “ four-quark bound states in chiral SU(3) quark model,” Commun. Theor. Phys. 50, 437–440 (2008), arXiv:0711.1029 [nucl-th] .
- Vijande et al. (2009) J. Vijande, A. Valcarce, and N. Barnea, “Exotic meson-meson molecules and compact four–quark states,” Phys. Rev. D 79, 074010 (2009), arXiv:0903.2949 [hep-ph] .
- Yang et al. (2009) Youchang Yang, Chengrong Deng, Jialun Ping, and T. Goldman, “-wave state in the constituent quark model,” Phys. Rev. D 80, 114023 (2009).
- Feng et al. (2013) G. Q. Feng, X. H. Guo, and B. S. Zou, “ bound state in the Bethe-Salpeter equation approach,” (2013), arXiv:1309.7813 [hep-ph] .
- Karliner and Rosner (2017) Marek Karliner and Jonathan L. Rosner, “Discovery of doubly-charmed baryon implies a stable () tetraquark,” Phys. Rev. Lett. 119, 202001 (2017), arXiv:1707.07666 [hep-ph] .
- Eichten and Quigg (2017) Estia J. Eichten and Chris Quigg, “Heavy-quark symmetry implies stable heavy tetraquark mesons ,” Phys. Rev. Lett. 119, 202002 (2017), arXiv:1707.09575 [hep-ph] .
- Yan et al. (2018) Xiaojun Yan, Bin Zhong, and Ruilin Zhu, “Doubly charmed tetraquarks in a diquark–antidiquark model,” Int. J. Mod. Phys. A 33, 1850096 (2018), arXiv:1804.06761 [hep-ph] .
- Ali et al. (2018) Ahmed Ali, Qin Qin, and Wei Wang, “Discovery potential of stable and near-threshold doubly heavy tetraquarks at the LHC,” Phys. Lett. B 785, 605–609 (2018), arXiv:1806.09288 [hep-ph] .
- Park et al. (2019) Woosung Park, Sungsik Noh, and Su Houng Lee, “Masses of the doubly heavy tetraquarks in a constituent quark model,” Nucl. Phys. A 983, 1–19 (2019), arXiv:1809.05257 [nucl-th] .
- Deng et al. (2020) Chengrong Deng, Hong Chen, and Jialun Ping, “Systematical investigation on the stability of doubly heavy tetraquark states,” Eur. Phys. J. A 56, 9 (2020), arXiv:1811.06462 [hep-ph] .
- Caramés et al. (2019) Teresa F. Caramés, Javier Vijande, and Alfredo Valcarce, “Exotic four-quark states,” Phys. Rev. D 99, 014006 (2019), arXiv:1812.08991 [hep-ph] .
- Hernández et al. (2020) E. Hernández, J. Vijande, A. Valcarce, and Jean-Marc Richard, “Spectroscopy, lifetime and decay modes of the tetraquark,” Phys. Lett. B 800, 135073 (2020), arXiv:1910.13394 [hep-ph] .
- Yang et al. (2020a) Gang Yang, Jialun Ping, and Jorge Segovia, “Doubly-heavy tetraquarks,” Phys. Rev. D 101, 014001 (2020a), arXiv:1911.00215 [hep-ph] .
- Bedolla et al. (2020) M. A. Bedolla, J. Ferretti, C. D. Roberts, and E. Santopinto, “Spectrum of fully-heavy tetraquarks from a diquark+antidiquark perspective,” Eur. Phys. J. C 80, 1004 (2020), arXiv:1911.00960 [hep-ph] .
- Yu et al. (2020) Meng-Ting Yu, Zhi-Yong Zhou, Dian-Yong Chen, and Zhiguang Xiao, “Possible molecular states in scatterings,” Phys. Rev. D 101, 074027 (2020), arXiv:1912.07348 [hep-ph] .
- Wallbott et al. (2020) Paul C. Wallbott, Gernot Eichmann, and Christian S. Fischer, “Disentangling different structures in heavy-light four-quark states,” Phys. Rev. D 102, 051501 (2020), arXiv:2003.12407 [hep-ph] .
- Tan et al. (2020) Yue Tan, Weichang Lu, and Jialun Ping, “Systematics of in a chiral constituent quark model,” Eur. Phys. J. Plus 135, 716 (2020), arXiv:2004.02106 [hep-ph] .
- Lü et al. (2020) Qi-Fang Lü, Dian-Yong Chen, and Yu-Bing Dong, “Masses of doubly heavy tetraquarks in a relativized quark model,” Phys. Rev. D 102, 034012 (2020), arXiv:2006.08087 [hep-ph] .
- Braaten et al. (2021) Eric Braaten, Li-Ping He, and Abhishek Mohapatra, “Masses of doubly heavy tetraquarks with error bars,” Phys. Rev. D 103, 016001 (2021), arXiv:2006.08650 [hep-ph] .
- Yang et al. (2020b) Gang Yang, Jialun Ping, and Jorge Segovia, “ tetraquarks in the chiral quark model,” Phys. Rev. D 102, 054023 (2020b), arXiv:2007.05190 [hep-ph] .
- Qin et al. (2021) Qin Qin, Yin-Fa Shen, and Fu-Sheng Yu, “Discovery potentials of double-charm tetraquarks,” Chin. Phys. C 45, 103106 (2021), arXiv:2008.08026 [hep-ph] .
- Meng et al. (2021) Q. Meng, E. Hiyama, A. Hosaka, M. Oka, P. Gubler, K. U. Can, T. T. Takahashi, and H. S. Zong, “Stable double-heavy tetraquarks: spectrum and structure,” Phys. Lett. B 814, 136095 (2021), arXiv:2009.14493 [nucl-th] .
- Meng et al. (2022) Qi Meng, Masayasu Harada, Emiko Hiyama, Atsushi Hosaka, and Makoto Oka, “Doubly heavy tetraquark resonant states,” Phys. Lett. B 824, 136800 (2022), arXiv:2106.11868 [hep-ph] .
- Chen (2021) Xiaoyun Chen, “Doubly heavy tetraquark states and ,” (2021), arXiv:2109.02828 [hep-ph] .
- Jin et al. (2021) Yi Jin, Shi-Yuan Li, Yan-Rui Liu, Qin Qin, Zong-Guo Si, and Fu-Sheng Yu, “Color and baryon number fluctuation of preconfinement system in production process and structure,” Phys. Rev. D 104, 114009 (2021), arXiv:2109.05678 [hep-ph] .
- Chen et al. (2021) Kan Chen, Rui Chen, Lu Meng, Bo Wang, and Shi-Lin Zhu, “Systematics of the heavy flavor hadronic molecules,” (2021), arXiv:2109.13057 [hep-ph] .
- Andreev (2021) Oleg Andreev, “On the -Quark Potential in String Models,” (2021), arXiv:2111.14418 [hep-ph] .
- Deng and Zhu (2021) Chengrong Deng and Shi-Lin Zhu, “ and its partners,” (2021), arXiv:2112.12472 [hep-ph] .
- Lee et al. (2008) Su Houng Lee, Shigehiro Yasui, Wei Liu, and Che Ming Ko, “Charmed exotics in Heavy Ion Collisions,” Eur. Phys. J. C 54, 259–265 (2008), arXiv:0707.1747 [hep-ph] .
- Lee and Yasui (2009) Su Houng Lee and Shigehiro Yasui, “Stable multiquark states with heavy quarks in a diquark model,” Eur. Phys. J. C 64, 283–295 (2009), arXiv:0901.2977 [hep-ph] .
- Hyodo et al. (2013) Tetsuo Hyodo, Yan-Rui Liu, Makoto Oka, Kazutaka Sudoh, and Shigehiro Yasui, “Production of doubly charmed tetraquarks with exotic color configurations in electron-positron collisions,” Phys. Lett. B 721, 56–60 (2013), arXiv:1209.6207 [hep-ph] .
- Luo et al. (2017) Si-Qiang Luo, Kan Chen, Xiang Liu, Yan-Rui Liu, and Shi-Lin Zhu, “Exotic tetraquark states with the configuration,” Eur. Phys. J. C 77, 709 (2017), arXiv:1707.01180 [hep-ph] .
- Hyodo et al. (2017) Tetsuo Hyodo, Yan-Rui Liu, Makoto Oka, and Shigehiro Yasui, “Spectroscopy and production of doubly charmed tetraquarks,” (2017), arXiv:1708.05169 [hep-ph] .
- Cheng et al. (2020) Jian-Bo Cheng, Shi-Yuan Li, Yan-Rui Liu, Yu-Nan Liu, Zong-Guo Si, and Tao Yao, “Spectrum and rearrangement decays of tetraquark states with four different flavors,” Phys. Rev. D 101, 114017 (2020), arXiv:2001.05287 [hep-ph] .
- Cheng et al. (2021) Jian-Bo Cheng, Shi-Yuan Li, Yan-Rui Liu, Zong-Guo Si, and Tao Yao, “Double-heavy tetraquark states with heavy diquark-antiquark symmetry,” Chin. Phys. C 45, 043102 (2021), arXiv:2008.00737 [hep-ph] .
- Weng et al. (2022) Xin-Zhen Weng, Wei-Zhen Deng, and Shi-Lin Zhu, “Doubly heavy tetraquarks in an extended chromomagnetic model *,” Chin. Phys. C 46, 013102 (2022), arXiv:2108.07242 [hep-ph] .
- Guo et al. (2022) Tao Guo, Jianing Li, Jiaxing Zhao, and Lianyi He, “Mass spectra of doubly heavy tetraquarks in an improved chromomagnetic interaction model,” Phys. Rev. D 105, 014021 (2022), arXiv:2108.10462 [hep-ph] .
- Navarra et al. (2007) Fernando S. Navarra, Marina Nielsen, and Su Houng Lee, “QCD sum rules study of mesons,” Phys. Lett. B 649, 166–172 (2007), arXiv:hep-ph/0703071 .
- Dias et al. (2011) J. M. Dias, S. Narison, F. S. Navarra, M. Nielsen, and J. M. Richard, “Relation between and from QCD,” Phys. Lett. B 703, 274–280 (2011), arXiv:1105.5630 [hep-ph] .
- Du et al. (2013) Meng-Lin Du, Wei Chen, Xiao-Lin Chen, and Shi-Lin Zhu, “Exotic , and states,” Phys. Rev. D 87, 014003 (2013), arXiv:1209.5134 [hep-ph] .
- Chen et al. (2014) Wei Chen, T. G. Steele, and Shi-Lin Zhu, “Exotic open-flavor , and , tetraquark states,” Phys. Rev. D 89, 054037 (2014), arXiv:1310.8337 [hep-ph] .
- Wang (2018) Zhi-Gang Wang, “Analysis of the axialvector doubly heavy tetraquark states with QCD sum rules,” Acta Phys. Polon. B 49, 1781 (2018), arXiv:1708.04545 [hep-ph] .
- Wang and Yan (2018) Zhi-Gang Wang and Ze-Hui Yan, “Analysis of the scalar, axialvector, vector, tensor doubly charmed tetraquark states with QCD sum rules,” Eur. Phys. J. C 78, 19 (2018), arXiv:1710.02810 [hep-ph] .
- Agaev et al. (2019) S. S. Agaev, K. Azizi, B. Barsbay, and H. Sundu, “Weak decays of the axial-vector tetraquark ,” Phys. Rev. D 99, 033002 (2019), arXiv:1809.07791 [hep-ph] .
- Sundu et al. (2019) H. Sundu, S. S. Agaev, and K. Azizi, “Semileptonic decays of the scalar tetraquark ,” Eur. Phys. J. C 79, 753 (2019), arXiv:1903.05931 [hep-ph] .
- Agaev et al. (2020a) S. S. Agaev, K. Azizi, and H. Sundu, “Double-heavy axial-vector tetraquark ,” Nucl. Phys. B 951, 114890 (2020a), arXiv:1905.07591 [hep-ph] .
- Tang et al. (2020) Liang Tang, Bin-Dong Wan, Kim Maltman, and Cong-Feng Qiao, “Doubly Heavy Tetraquarks in QCD Sum Rules,” Phys. Rev. D 101, 094032 (2020), arXiv:1911.10951 [hep-ph] .
- Agaev et al. (2020b) S. S. Agaev, K. Azizi, B. Barsbay, and H. Sundu, “Heavy exotic scalar meson ,” Phys. Rev. D 101, 094026 (2020b), arXiv:1912.07656 [hep-ph] .
- Agaev et al. (2020c) S. S. Agaev, K. Azizi, B. Barsbay, and H. Sundu, “Stable scalar tetraquark ,” Eur. Phys. J. A 56, 177 (2020c), arXiv:2001.01446 [hep-ph] .
- Wang and Chen (2020) Qi-Nan Wang and Wei Chen, “Fully open-flavor tetraquark states and with ,” Eur. Phys. J. C 80, 389 (2020), arXiv:2002.04243 [hep-ph] .
- Agaev et al. (2021a) S. S. Agaev, K. Azizi, B. Barsbay, and H. Sundu, “A family of double-beauty tetraquarks: Axial-vector state ,” Chin. Phys. C 45, 013105 (2021a), arXiv:2002.04553 [hep-ph] .
- Agaev et al. (2021b) S. S. Agaev, K. Azizi, B. Barsbay, and H. Sundu, “Semileptonic and nonleptonic decays of the axial-vector tetraquark ,” Eur. Phys. J. A 57, 106 (2021b), arXiv:2008.02049 [hep-ph] .
- Agaev et al. (2022) S. S. Agaev, K. Azizi, and H. Sundu, “Newly observed exotic doubly charmed meson ,” Nucl. Phys. B 975, 115650 (2022), arXiv:2108.00188 [hep-ph] .
- Azizi and Özdem (2021) K. Azizi and U. Özdem, “Magnetic dipole moments of the and tetraquark states,” Phys. Rev. D 104, 114002 (2021), arXiv:2109.02390 [hep-ph] .
- Aliev et al. (2021) T. M. Aliev, S. Bilmis, and M. Savci, “Determination of the spectroscopic parameters of beauty-partners of from QCD,” (2021), arXiv:2111.01081 [hep-ph] .
- Özdem (2022) Ulaş Özdem, “Magnetic moments of the doubly charged axial-vector states,” Phys. Rev. D 105, 054019 (2022), arXiv:2112.10402 [hep-ph] .
- Bilmiş (2021) Selçuk Bilmiş, “Mixing angles between the tetraquark states with two heavy quarks within QCD sum rules,” Turk. J. Phys. 45, 390–399 (2021), arXiv:2201.06109 [hep-ph] .
- Detmold et al. (2007) William Detmold, Kostas Orginos, and Martin J. Savage, “ Potentials in Quenched Lattice QCD,” Phys. Rev. D 76, 114503 (2007), arXiv:hep-lat/0703009 .
- Wagner (2010) Marc Wagner (ETM, Y), “Forces between static-light mesons,” PoS LATTICE2010, 162 (2010), arXiv:1008.1538 [hep-lat] .
- Bali and Hetzenegger (2010) Gunnar Bali and Martin Hetzenegger (QCDSF), “Static-light meson-meson potentials,” PoS LATTICE2010, 142 (2010), arXiv:1011.0571 [hep-lat] .
- Bicudo and Wagner (2013) Pedro Bicudo and Marc Wagner (European Twisted Mass), “Lattice QCD signal for a bottom-bottom tetraquark,” Phys. Rev. D 87, 114511 (2013), arXiv:1209.6274 [hep-ph] .
- Brown and Orginos (2012) Zachary S. Brown and Kostas Orginos, “Tetraquark bound states in the heavy-light heavy-light system,” Phys. Rev. D 86, 114506 (2012), arXiv:1210.1953 [hep-lat] .
- Ikeda et al. (2014) Yoichi Ikeda, Bruno Charron, Sinya Aoki, Takumi Doi, Tetsuo Hatsuda, Takashi Inoue, Noriyoshi Ishii, Keiko Murano, Hidekatsu Nemura, and Kenji Sasaki, “Charmed tetraquarks and from dynamical lattice QCD simulations,” Phys. Lett. B 729, 85–90 (2014), arXiv:1311.6214 [hep-lat] .
- Bicudo et al. (2015) Pedro Bicudo, Krzysztof Cichy, Antje Peters, Björn Wagenbach, and Marc Wagner, “Evidence for the existence of and the non-existence of and tetraquarks from lattice QCD,” Phys. Rev. D 92, 014507 (2015), arXiv:1505.00613 [hep-lat] .
- Bicudo et al. (2016) Pedro Bicudo, Krzysztof Cichy, Antje Peters, and Marc Wagner, “ interactions with static bottom quarks from Lattice QCD,” Phys. Rev. D 93, 034501 (2016), arXiv:1510.03441 [hep-lat] .
- Francis et al. (2017) Anthony Francis, Renwick J. Hudspith, Randy Lewis, and Kim Maltman, “Lattice Prediction for Deeply Bound Doubly Heavy Tetraquarks,” Phys. Rev. Lett. 118, 142001 (2017), arXiv:1607.05214 [hep-lat] .
- Bicudo et al. (2017a) Pedro Bicudo, Jonas Scheunert, and Marc Wagner, “Including heavy spin effects in the prediction of a tetraquark with lattice QCD potentials,” Phys. Rev. D 95, 034502 (2017a), arXiv:1612.02758 [hep-lat] .
- Bicudo et al. (2017b) Pedro Bicudo, Marco Cardoso, Antje Peters, Martin Pflaumer, and Marc Wagner, “ tetraquark resonances with lattice QCD potentials and the Born-Oppenheimer approximation,” Phys. Rev. D 96, 054510 (2017b), arXiv:1704.02383 [hep-lat] .
- Francis et al. (2019) Anthony Francis, Renwick J. Hudspith, Randy Lewis, and Kim Maltman, “Evidence for charm-bottom tetraquarks and the mass dependence of heavy-light tetraquark states from lattice QCD,” Phys. Rev. D 99, 054505 (2019), arXiv:1810.10550 [hep-lat] .
- Junnarkar et al. (2019) Parikshit Junnarkar, Nilmani Mathur, and M. Padmanath, “Study of doubly heavy tetraquarks in Lattice QCD,” Phys. Rev. D 99, 034507 (2019), arXiv:1810.12285 [hep-lat] .
- Leskovec et al. (2019) Luka Leskovec, Stefan Meinel, Martin Pflaumer, and Marc Wagner, “Lattice QCD investigation of a doubly-bottom tetraquark with quantum numbers ,” Phys. Rev. D 100, 014503 (2019), arXiv:1904.04197 [hep-lat] .
- Hudspith et al. (2020) R. J. Hudspith, B. Colquhoun, A. Francis, R. Lewis, and K. Maltman, “A lattice investigation of exotic tetraquark channels,” Phys. Rev. D 102, 114506 (2020), arXiv:2006.14294 [hep-lat] .
- Mohanta and Basak (2020) Protick Mohanta and Subhasish Basak, “Construction of tetraquark states on lattice with NRQCD bottom and HISQ up and down quarks,” Phys. Rev. D 102, 094516 (2020), arXiv:2008.11146 [hep-lat] .
- Bicudo et al. (2021) Pedro Bicudo, Antje Peters, Sebastian Velten, and Marc Wagner, “Importance of meson-meson and of diquark-antidiquark creation operators for a tetraquark,” Phys. Rev. D 103, 114506 (2021), arXiv:2101.00723 [hep-lat] .
- Padmanath and Mathur (2021) M. Padmanath and Nilmani Mathur, “four-quark states from Lattice QCD,” in 38th International Symposium on Lattice Field Theory (2021) arXiv:2111.01147 [hep-lat] .
- Harada et al. (2020) Masayasu Harada, Yan-Rui Liu, Makoto Oka, and Kei Suzuki, “Chiral effective theory of diquarks and the anomaly,” Phys. Rev. D 101, 054038 (2020), arXiv:1912.09659 [hep-ph] .
- Kim et al. (2020) Yonghee Kim, Emiko Hiyama, Makoto Oka, and Kei Suzuki, “Spectrum of singly heavy baryons from a chiral effective theory of diquarks,” Phys. Rev. D 102, 014004 (2020), arXiv:2003.03525 [hep-ph] .
- Kim et al. (2021) Yonghee Kim, Yan-Rui Liu, Makoto Oka, and Kei Suzuki, “Heavy baryon spectrum with chiral multiplets of scalar and vector diquarks,” Phys. Rev. D 104, 054012 (2021), arXiv:2105.09087 [hep-ph] .
- ’t Hooft (1976) Gerard ’t Hooft, “Computation of the Quantum Effects Due to a Four-Dimensional Pseudoparticle,” Phys. Rev. D 14, 3432–3450 (1976), [Erratum: Phys.Rev.D 18, 2199 (1978)].
- ’t Hooft (1986) Gerard ’t Hooft, “How Instantons Solve the Problem,” Phys. Rept. 142, 357–387 (1986).
- Gell-Mann (1962) M. Gell-Mann, “Symmetries of baryons and mesons,” Phys. Rev. 125, 1067–1084 (1962).
- Okubo (1962) Susumu Okubo, “Note on unitary symmetry in strong interactions,” Prog. Theor. Phys. 27, 949–966 (1962).
- Silvestre-Brac (1996) B. Silvestre-Brac, “Spectrum and static properties of heavy baryons,” Few Body Syst. 20, 1 (1996).
- Zyla et al. (2020) P.A. Zyla et al. (Particle Data Group), “Review of Particle Physics,” PTEP 2020, 083C01 (2020).
- Yoshida et al. (2015) Tetsuya Yoshida, Emiko Hiyama, Atsushi Hosaka, Makoto Oka, and Katsunori Sadato, “Spectrum of heavy baryons in the quark model,” Phys. Rev. D92, 114029 (2015), arXiv:1510.01067 [hep-ph] .
- Kamimura (1988) M. Kamimura, “Nonadiabatic coupled-rearrangement-channel approach to muonic molecules,” Phys. Rev. A38, 621 (1988).
- Hiyama et al. (2003) E. Hiyama, Y. Kino, and M. Kamimura, “Gaussian expansion method for few-body systems,” Prog. Part. Nucl. Phys. 51, 223 (2003).
- Kumakawa and Jido (2017) Kento Kumakawa and Daisuke Jido, “Excitation energy spectra of the and baryons in a finite-size diquark model,” PTEP 2017, 123D01 (2017), arXiv:1708.02012 [nucl-th] .
- Kumakawa and Jido (2021) Kento Kumakawa and Daisuke Jido, “Excitation spectra of heavy baryons in diquark models,” (2021), arXiv:2107.11950 [hep-ph] .