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

Doubly heavy tetraquarks in a chiral-diquark picture

Yonghee Kim kimu.ryonhi@phys.kyushu-u.ac.jp Department of Physics, Kyushu University, Fukuoka 819-0395, Japan    Makoto Oka oka@post.j-parc.jp Advanced Science Research Center, Japan Atomic Energy Agency (JAEA), Tokai 319-1195, Japan Nishina Center for Accelerator-Based Science, RIKEN, Wako 351-0198, Japan    Kei Suzuki k.suzuki.2010@th.phys.titech.ac.jp Advanced Science Research Center, Japan Atomic Energy Agency (JAEA), Tokai 319-1195, Japan
Abstract

Energy spectrum of doubly heavy tetraquarks, TQQT_{QQ} (QQq¯q¯QQ\bar{q}\bar{q} with Q=c,bQ=c,b and q=u,d,sq=u,d,s), is studied in the potential chiral-diquark model. Using the chiral effective theory of diquarks and the quark-diquark-based potential model, the TbbT_{bb}, TccT_{cc}, and TcbT_{cb} tetraquarks are described as a three-body system composed of two heavy quarks and an antidiquark. We find several TbbT_{bb} bound states, while no TccT_{cc} and TcbT_{cb} (deep) bound state is seen. We also study the change of the TQQT_{QQ} 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 Tcc+T_{cc}^{+} 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 JP=1+J^{P}=1^{+}. The mass relative to D+D0D^{*+}-D^{0} threshold and the decay width were firstly determined as 273±61-273\pm 61 keV and 410±165410\pm 165 keV with the Breit-Wigner profile Aaij et al. (2021a), and these values have been updated as 361±40-361\pm 40 keV and 47.8±1.947.8\pm 1.9 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, QQq¯q¯QQ\bar{q}\bar{q} (Q=c,bQ=c,b and q=u,d,sq=u,d,s), 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 TQQT_{QQ} 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 TbbT_{bb} tetraquark below the strong-decay threshold, but no stable doubly charmed TccT_{cc} tetraquark. The predictions for the bottom-charmed TcbT_{cb} tetraquark are divided, whether it is stable or not.

Multiquark states are often considered from the perspective of their substructures (clusters). The TQQT_{QQ} 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 𝟑¯\bf{\bar{3}}, scalar/pseudoscalar diquarks based on the SU(3)R×SU(3)LSU(3)_{R}\times SU(3)_{L} chiral symmetry. Here, the 0±0^{\pm} 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 𝟑¯\bf{\bar{3}}, 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, TQQT_{QQ} 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 TQQT_{QQ} tetraquark masses on the chiral symmetry breaking parameter.

This paper is organized as follows. In Sec. II, we introduce the chiral effective theory of diquarks and construct the potential models for the diquark-pictured TQQT_{QQ} tetraquarks. In Sec. III, we show the results for various TQQT_{QQ} tetraquark states. The last section IV is for our conclusion and outlook.

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 TQQT_{QQ} 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,

=S+V+14Tr[μΣμΣ]V(Σ)+VS.\displaystyle\mathcal{L}=\mathcal{L}_{S}+\mathcal{L}_{V}+\frac{1}{4}{\rm Tr}[\partial^{\mu}\Sigma^{\dagger}\partial_{\mu}\Sigma]-V(\Sigma)+\mathcal{L}_{V-S}. (1)

Here the first term S\mathcal{L}_{S} describes the effective Lagrangian for the scalar and pseudoscalar diquarks, and V\mathcal{L}_{V} for the vector and axial-vector diquarks. Their explicit forms are given below. The last term, VS\mathcal{L}_{V-S}, 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 Σ\Sigma, respectively. This meson operator Σ\Sigma contains nonet scalar σ\sigma and pseudoscalar π\pi mesons, whose chiral transform is given by

Σij=σij+iπijUL,ikΣkmUR,mj.\Sigma_{ij}=\sigma_{ij}+i\pi_{ij}\rightarrow U_{L,ik}\Sigma_{km}U^{{\dagger}}_{R,mj}. (2)

The potential term for the Σ\Sigma meson field, V(Σ)V(\Sigma), causes spontaneous chiral symmetry breaking, represented by the vacuum expectation value of scalar meson σ\sigma as

Σij=σij=fπδij,πij=0,\langle\Sigma_{ij}\rangle=\langle\sigma_{ij}\rangle=f_{\pi}\delta_{ij}~{},~{}~{}~{}\langle\pi_{ij}\rangle=0, (3)

where fπ92f_{\pi}\simeq 92 MeV is the pion decay constant. In this vacuum, π\pi meson is regarded as the massless Nambu-Goldstone bosons.

Using the Σ\Sigma meson field and the chiral diquark operators summarized in Table. 1, the effective Lagrangians, S\mathcal{L}_{S} and V\mathcal{L}_{V}, are expressed as

S=𝒟μdR,i(𝒟μdR,i)+𝒟μdL,i(𝒟μdL,i)mS02(dR,idR,i+dL,idL,i)mS12fπ(dR,iΣijdL,j+dL,iΣijdR,j)mS222fπ2ϵijkϵlmn(dR,kΣliΣmjdL,n+dL,kΣliΣm,jdR,n),\displaystyle\begin{split}\mathcal{L}_{S}&=\mathcal{D}_{\mu}d_{R,i}(\mathcal{D}^{\mu}d_{R,i})^{\dagger}+\mathcal{D}_{\mu}d_{L,i}(\mathcal{D}^{\mu}d_{L,i})^{\dagger}\\ &-m_{S0}^{2}(d_{R,i}d_{R,i}^{\dagger}+d_{L,i}d_{L,i}^{\dagger})\\ &-\frac{m_{S1}^{2}}{f_{\pi}}(d_{R,i}\Sigma_{ij}^{\dagger}d_{L,j}^{\dagger}+d_{L,i}\Sigma_{ij}d_{R,j}^{\dagger})\\ &-\frac{m_{S2}^{2}}{2f_{\pi}^{2}}\epsilon_{ijk}\epsilon_{lmn}(d_{R,k}\Sigma_{li}\Sigma_{mj}d_{L,n}^{\dagger}+d_{L,k}\Sigma^{\dagger}_{li}\Sigma^{\dagger}_{m,j}d_{R,n}^{\dagger}),\\ \end{split} (4)
V=12Tr[FμνFμν]+mV02Tr[dμdμ]+mV12fπ2Tr[ΣdμΣTdμT]+mV22fπ2[Tr{ΣTΣTdμdμ}+Tr{ΣΣdμdμ}].\displaystyle\begin{split}\mathcal{L}_{V}&=\frac{1}{2}{\rm Tr}[F^{\mu\nu}F^{{\dagger}}_{\mu\nu}]+m_{V0}^{2}{\rm Tr}[d^{\mu}d^{{\dagger}}_{\mu}]\\ &+\frac{m_{V1}^{2}}{f_{\pi}^{2}}{\rm Tr}[\Sigma^{{\dagger}}d^{\mu}\Sigma^{T}d^{{\dagger}T}_{\mu}]\\ &+\frac{m_{V2}^{2}}{f_{\pi}^{2}}[{\rm Tr}\{\Sigma^{T}\Sigma^{{\dagger}T}d_{\mu}^{{\dagger}}d^{\mu}\}+{\rm Tr}\{\Sigma\Sigma^{{\dagger}}d^{\mu}d^{{\dagger}}_{\mu}\}].\\ \end{split} (5)
Table 1: Diquark operators in the chiral basis.
Chiral operator Spin Color Flavor
dR,ia=ϵabcϵijk(qR,jbTCqR,kc)d^{a}_{R,i}=\epsilon_{abc}\epsilon_{ijk}(q^{bT}_{R,j}Cq^{c}_{R,k}) 0 𝟑¯\overline{\bm{3}} 𝟑¯\bar{\bm{3}}
dL,ia=ϵabcϵijk(qL,jbTCqL,kc)d^{a}_{L,i}=\epsilon_{abc}\epsilon_{ijk}(q^{bT}_{L,j}Cq^{c}_{L,k}) 0 𝟑¯\overline{\bm{3}} 𝟑¯\bar{\bm{3}}
dija,μ=ϵabc(qL,ibTCγμqR,jc)d^{a,\mu}_{ij}=\epsilon_{abc}(q^{bT}_{L,i}C\gamma^{\mu}q^{c}_{R,j}) 11 𝟑¯\overline{\bm{3}} 𝟑¯,𝟔\bar{\bm{3}},\bm{6}

Since the diquark is not a color-singlet state, we introduce the color-gauge-covariant derivative, 𝒟μ=μ+igTαGα,μ\mathcal{D}^{\mu}=\partial^{\mu}+igT^{\alpha}G^{\alpha,\mu}, with Gα,μG^{\alpha,\mu} being the gluon field and TαT^{\alpha} being the color SU(3)SU(3) generator for the color 𝟑¯\bf\bar{3} representation. It is used for the kinetic term of these Lagrangians. Fμν=𝒟μdν𝒟νdμF^{\mu\nu}=\mathcal{D}^{\mu}d^{\nu}-\mathcal{D}^{\nu}d^{\mu} 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 mS02m_{S0}^{2}, mS12m_{S1}^{2}, and mS22m_{S2}^{2} in Eq. (4). mS0m_{S0} 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 mS1m_{S1} and mS2m_{S2}. In particular, the mS12m_{S1}^{2} term also causes the UA(1)U_{A}(1) symmetry breaking ’t Hooft (1976, 1986). On the other hand, the masses of vector and axial-vector diquarks are given by three parameters mV02m_{V0}^{2}, mV12m_{V1}^{2}, and mV22m_{V2}^{2} in Eq. (5). Similar to the scalar and pseudoscalar diquarks, mV0m_{V0} is the chiral invariant mass of vector and axial-vector diquarks, and their mass splitting is given by mV1m_{V1} and mV2m_{V2}.

The non-zero strange quark mass brings about the explicit chiral symmetry breaking and the flavor SU(3)SU(3) symmetry breaking. In this term, the masses of scalar and pseudoscalar diquarks, M(0+)M(0^{+}) and M(0)M(0^{-}), are expressed in Ref. Harada et al. (2020) as

[Mud(0+)]2\displaystyle\left[M_{ud}(0^{+})\right]^{2} =mS02(1+ϵ)mS12mS22,\displaystyle=m_{S0}^{2}-(1+\epsilon)m_{S1}^{2}-m_{S2}^{2}, (6)
[Mns(0+)]2\displaystyle\left[M_{ns}(0^{+})\right]^{2} =mS02mS12(1+ϵ)mS22,\displaystyle=m_{S0}^{2}-m_{S1}^{2}-(1+\epsilon)m_{S2}^{2}, (7)
[Mud(0)]2\displaystyle\left[M_{ud}(0^{-})\right]^{2} =mS02+(1+ϵ)mS12+mS22,\displaystyle=m_{S0}^{2}+(1+\epsilon)m_{S1}^{2}+m_{S2}^{2}, (8)
[Mns(0)]2\displaystyle\left[M_{ns}(0^{-})\right]^{2} =mS02+mS12+(1+ϵ)mS22,\displaystyle=m_{S0}^{2}+m_{S1}^{2}+(1+\epsilon)m_{S2}^{2}, (9)

where the index nn stands for uu or dd quark, and ss for ss quark, meaning two constituent quarks of a diquark. The constant ϵ\epsilon is given by

ϵ=fsfπ(1+msgsfπgsfs)2/3,\displaystyle\epsilon=\frac{f_{s}}{f_{\pi}}\left(1+\frac{m_{s}-g_{s}f_{\pi}}{g_{s}f_{s}}\right)\simeq 2/3, (10)

with fs=2fKfπf_{s}=2f_{K}-f_{\pi}, where ms,gs,fπm_{s},g_{s},f_{\pi}, and fKf_{K} 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

[Mns(0+)]2[Mud(0+)]2=[Mud(0)]2\displaystyle[M_{ns}(0^{+})]^{2}-[M_{ud}(0^{+})]^{2}=[M_{ud}(0^{-})]^{2}- [Mns(0)]2\displaystyle[M_{ns}(0^{-})]^{2} (11)
>0.\displaystyle>0.

From Eq. (11), we see the non-strange pseudoscalar diquark is heavier than the singly strange pseudoscalar diquark, Mud(0)>Mns(0)M_{ud}(0^{-})>M_{ns}(0^{-}), which is called the inverse mass hierarchy.

The masses of axial-vector and vector diquarks, M(1+)M(1^{+}) and M(1)M(1^{-}), are expressed in Ref. Kim et al. (2021) as

[Mnn(1+)]2\displaystyle\left[M_{nn}(1^{+})\right]^{2} =mV02+mV12+2mV22,\displaystyle=m_{V0}^{2}+m_{V1}^{2}+2m_{V2}^{2}, (12)
[Mns(1+)]2\displaystyle\left[M_{ns}(1^{+})\right]^{2} =mV02+mV12+2mV22\displaystyle=m_{V0}^{2}+m_{V1}^{2}+2m_{V2}^{2} (13)
+ϵ(mV12+2mV22),\displaystyle+\epsilon(m_{V1}^{2}+2m_{V2}^{2}),
[Mss(1+)]2\displaystyle\left[M_{ss}(1^{+})\right]^{2} =mV02+mV12+2mV22\displaystyle=m_{V0}^{2}+m_{V1}^{2}+2m_{V2}^{2} (14)
+2ϵ(mV12+2mV22),\displaystyle+2\epsilon(m_{V1}^{2}+2m_{V2}^{2}),
[Mud(1)]2\displaystyle\left[M_{ud}(1^{-})\right]^{2} =mV02mV12+2mV22,\displaystyle=m_{V0}^{2}-m_{V1}^{2}+2m_{V2}^{2}, (15)
[Mns(1)]2\displaystyle\left[M_{ns}(1^{-})\right]^{2} =mV02mV12+2mV22\displaystyle=m_{V0}^{2}-m_{V1}^{2}+2m_{V2}^{2} (16)
+ϵ(mV12+2mV22).\displaystyle+\epsilon(-m_{V1}^{2}+2m_{V2}^{2}).

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

[Mss(1+)]2[Mns(1+)]2\displaystyle[M_{ss}(1^{+})]^{2}-[M_{ns}(1^{+})]^{2} =[Mns(1+)]2[Mnn(1+)]2\displaystyle=[M_{ns}(1^{+})]^{2}-[M_{nn}(1^{+})]^{2} (17)
=ϵ(mV12+2mV22).\displaystyle=\epsilon(m_{V1}^{2}+2m_{V2}^{2}).

Here the square mass differences are characterized by the number of ss quark. On the other hand, the square mass difference between nonstrange and singly strange vector diquarks is given by

[Mns(1)]2[Mud(1)]2=ϵ(mV12+2mV22).\displaystyle[M_{ns}(1^{-})]^{2}-[M_{ud}(1^{-})]^{2}=\epsilon(-m_{V1}^{2}+2m_{V2}^{2}). (18)

As shown in Table. 2, the parameter mV12m_{V1}^{2} 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.

Table 2: Diquark masses and parameters of the chiral effective Lagrangian taken from Refs. Kim et al. (2020, 2021), whose values are determined by the Y-potential model.
Diquark masses (MeV\rm MeV) Parameters (MeV2\rm MeV^{2})
Mud(0+)M_{ud}(0^{+})~{} 725 mS02~{}~{}m_{S0}^{2}~{} (1119)2(1119)^{2}
Mns(0+)M_{ns}(0^{+})~{} 942 mS12~{}~{}m_{S1}^{2}~{} (690)2(690)^{2}
Mud(0)M_{ud}(0^{-})~{} 1406 mS22~{}~{}m_{S2}^{2}~{} (258)2-(258)^{2}
Mns(0)M_{ns}(0^{-})~{} 1271
Mnn(1+)M_{nn}(1^{+})~{} 973 mV02~{}~{}m_{V0}^{2}~{} (708)2(708)^{2}
Mns(1+)M_{ns}(1^{+})~{} 1116 mV12~{}~{}m_{V1}^{2}~{} (757)2-(757)^{2}
Mss(1+)M_{ss}(1^{+})~{} 1242 mV22~{}~{}m_{V2}^{2}~{} (714)2(714)^{2}
Mud(1)M_{ud}(1^{-})~{} 1447
Mns(1)M_{ns}(1^{-})~{} 1776

II.2 Potential models

In order to calculate the spectrum of TQQT_{QQ} 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 (Q1,Q2Q_{1},Q_{2}) and one antidiquark (d¯\bar{d}). We need two kinds of two-body potentials, one for the interaction between two heavy quarks, VQ1Q2V_{Q_{1}Q_{2}}, and the other for between a heavy quark and a color 𝟑\bm{3} antidiquark, VQid¯V_{Q_{i}\bar{d}}.

The Hamiltonian for this tetraquark system is written as

H=k={Q1,Q2,d¯}\displaystyle H=\sum_{k=\{Q_{1},Q_{2},\bar{d}\}} (Mk+𝒑k22Mk)Kc.m.\displaystyle\left(M_{k}+\frac{\bm{p}_{k}^{2}}{2M_{k}}\right)-K_{c.m.} (19)
+VQ1Q2+VQ1d¯+VQ2d¯,\displaystyle+V_{Q_{1}Q_{2}}+V_{Q_{1}\bar{d}}+V_{Q_{2}\bar{d}},

where MQi/d¯M_{Q_{i}/\bar{d}} and 𝒑Qi/d¯\bm{p}_{Q_{i}/\bar{d}} are the mass and momentum of constituent particles, respectively. Kc.m.K_{c.m.} is the kinetic energy of the center of mass, which is subtracted and the remained kinetic energy is for the rcr_{c} and RcR_{c} parts of Jacobi coordinates in Fig. 1.

Refer to caption
Figure 1: Jacobi coordinates for the three-body system. In each channel (c=1,2,3c=1,2,3), QQ and q¯\bar{q} denote the heavy quark and the light antiquark, respectively.

For the potential between two heavy-quarks, VQ1Q2V_{Q_{1}Q_{2}}, we choose the form of the “AP1AP1 potential” in Ref. Silvestre-Brac (1996), which reproduces the spectra of heavy mesons, color-singlet QQ¯Q\bar{Q}. The AP1AP1 potential contains the power-law confinement term and one-gluon-exchange term expressed as

VQ1Q¯2(r)\displaystyle V_{Q_{1}\bar{Q}_{2}}(r) =αr+λr2/3+C\displaystyle=-\frac{\alpha}{r}+\lambda r^{2/3}+C (20)
+(𝒔Q1𝒔Q2)8κ3MQ1MQ2πer2/r02r03,\displaystyle+({\bm{s}}_{Q_{1}}\cdot{\bm{s}}_{Q_{2}})\frac{8\kappa}{3M_{Q_{1}}M_{Q_{2}}\sqrt{\pi}}\frac{e^{-r^{2}/r_{0}^{2}}}{r_{0}^{3}},

with the mass-dependent parameter

r0=A(2MQ1MQ2MQ1+MQ2)B.\displaystyle r_{0}=A\left(\frac{2M_{Q_{1}}M_{Q_{2}}}{M_{Q_{1}}+M_{Q_{2}}}\right)^{-B}. (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.

Table 3: Parameters and quark masses of the AP1AP1 potential VQ1Q¯2(r)V_{Q_{1}\bar{Q}_{2}}(r) and the calculated masses of ground-state heavy mesons compared with the experimental values in Ref. Zyla et al. (2020).
Masses (MeV\rm MeV)
Parameters (AP1AP1) Mesons AP1AP1 Silvestre-Brac (1996) Exp. Zyla et al. (2020)
α\alpha 0.42420.4242 D(0)D(0^{-}) 1881 1868.04
λ\lambda (GeV5/3) 0.38980.3898 D(1)D^{*}(1^{-}) 2033 2009.12
CC (GeV) 1.1313-1.1313 Ds(0)D_{s}(0^{-}) 1955 1968.34
κ\kappa 1.80251.8025 Ds(1)D_{s}^{*}(1^{-}) 2107 2112.20
AA (GeVB-1) 1.52961.5296 B(0)B(0^{-}) 5311 5279.44
BB 0.32630.3263 B(1)B^{*}(1^{-}) 5367 5324.70
MnM_{n} (GeV) 0.2770.277 Bs(0)B_{s}(0^{-}) 5356 5366.88
MsM_{s} (GeV) 0.5530.553 Bs(1)B_{s}^{*}(1^{-}) 5418 5415.40
McM_{c} (GeV) 1.8191.819 ηc(0)\eta_{c}(0^{-}) 2982 2983.90
MbM_{b} (GeV) 5.2065.206 J/ψ(1)J/\psi(1^{-}) 3103 3096.90
Bc(0)B_{c}(0^{-}) 6269 6274.90
ηb(0)\eta_{b}(0^{-}) 9401 9398.70
Υ(1)\Upsilon(1^{-}) 9461 9460.30

Here we consider the effect of color structure expressed as (𝝀i𝝀j)(\bm{\lambda}_{i}\cdot\bm{\lambda}_{j}) by the Gell-Mann matrices. As in Table. 4, the value of this factor for the quark-antiquark picture inside a meson (𝟑𝟑¯=𝟏\bm{3}\otimes\bar{\bm{3}}=\bm{1}) is two times larger than that for the two quarks with color 𝟑¯\bar{\bm{3}} representation (𝟑𝟑=𝟑¯\bm{3}\otimes\bm{3}=\bar{\bm{3}}). For this reason, we assume that the AP1AP1 potential in Eq. (20) can be generalized to the form proportional to this factor, Vqq¯(r)(𝝀i𝝀j)V_{q\bar{q}}(r)\propto(\bm{\lambda}_{i}\cdot\bm{\lambda}_{j}), and we write the potential between two heavy-quarks of color 𝟑¯\bar{\bm{3}} as

VQ1Q2(r)=12VQ1Q¯2(r).\displaystyle V_{Q_{1}Q_{2}}(r)=\frac{1}{2}V_{Q_{1}\bar{Q}_{2}}(r). (22)
Table 4: Color structures and the values of (𝝀i𝝀j)(\bm{\lambda}_{i}\cdot\bm{\lambda}_{j}), where QQ and dd denote the quark and the diquark, respectively.
Two particles Color structure (𝝀i𝝀j)(\bm{\lambda}_{i}\cdot\bm{\lambda}_{j})
QQ¯Q-\bar{Q}, QdQ-d 𝟑𝟑¯=𝟏\bm{3}\otimes\bar{\bm{3}}=\bm{1} 16/3-16/3
QQQ-Q, Qd¯Q-\bar{d} 𝟑𝟑=𝟑¯\bm{3}\otimes\bm{3}=\bar{\bm{3}} 8/3-8/3

Similarly, to construct the potential for between a heavy quark and a color 𝟑\bm{3} antidiquark, VQid¯V_{Q_{i}\bar{d}}, 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

VQd(r)\displaystyle V_{Qd}(r) =αr+λr+CQ\displaystyle=-\frac{\alpha^{\prime}}{r}+\lambda^{\prime}r+C^{\prime}_{Q} (23)
+(𝒔Q𝒔d)κQMQMdΛ2rexp(Λr),\displaystyle+({\bm{s}}_{Q}\cdot{\bm{s}}_{d})\frac{\kappa^{\prime}_{Q}}{M_{Q}M_{d}}\frac{\Lambda^{\prime 2}}{r}\exp{(-\Lambda^{\prime}r)},

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 Cc,bC_{c,b} from the original values in Ref. Kim et al. (2021), so as to use the same heavy-quark mass values with the AP1AP1 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 (𝝀i𝝀j)(\bm{\lambda}_{i}\cdot\bm{\lambda}_{j}) again. From the value of (𝝀i𝝀j)(\bm{\lambda}_{i}\cdot\bm{\lambda}_{j}) summarized in Table. 4 and the assumption that the Y-potential is proportional to the factor (𝝀i𝝀j)(\bm{\lambda}_{i}\cdot\bm{\lambda}_{j}), VQd(r)(𝝀i𝝀j)V_{Qd}(r)\propto(\bm{\lambda}_{i}\cdot\bm{\lambda}_{j}), we can write the potential between the heavy quark and the color 𝟑\bm{3} antidiquark inside the color-singlet TQQT_{QQ} tetraquark as

VQid¯(r)=12VQd(r).\displaystyle V_{Q_{i}\bar{d}}(r)=\frac{1}{2}V_{Qd}(r). (24)
Table 5: Parameters of the Y-potential and the calculated masses of ground-state singly heavy baryons compared with the experimental values in Ref. Zyla et al. (2020). For the parameters, μ\mu in the parameter α\alpha is the reduced mass of two interacting particles, and Cc,bC_{c,b} are tuned from Ref. Kim et al. (2021). For the quark and diquark masses, we use the value summarized in Tables. 2 and 3.
Masses (MeV\rm MeV)
Parameters (Y-pot.) Baryons Y-pot. Kim et al. (2021) Exp. Zyla et al. (2020)
α\alpha^{\prime} 60/μ60/\mu Λc\Lambda_{c} (1/2+1/2^{+}) 2284 2286.46
λ\lambda^{\prime} (GeV2) 0.1650.165 Σc\Sigma_{c} (1/2+1/2^{+}) 2451 2453.54
CcC^{\prime}_{c} (GeV) 0.900-0.900 Σc\Sigma_{c} (3/2+3/2^{+}) 2513 2518.13
CbC^{\prime}_{b} (GeV) 0.913-0.913 Ξc\Xi_{c} (1/2+1/2^{+}) 2467 2469.42
κc\kappa^{\prime}_{c} 0.85860.8586 Ξc\Xi^{\prime}_{c} (1/2+1/2^{+}) 2581 2578.80
κb\kappa^{\prime}_{b} 0.66350.6635 Ξc\Xi^{\prime}_{c} (3/2+3/2^{+}) 2639 2645.88
Λ\Lambda^{\prime} (GeV) 0.6910.691 Ωc\Omega_{c} (1/2+1/2^{+}) 2699 2695.20
Ωc\Omega_{c} (3/2+3/2^{+}) 2752 2765.90
Λb\Lambda_{b} (1/2+1/2^{+}) 5619 5619.60
Σb\Sigma_{b} (1/2+1/2^{+}) 5810 5813.10
Σb\Sigma_{b} (3/2+3/2^{+}) 5829 5832.53
Ξb\Xi_{b} (1/2+1/2^{+}) 5796 5794.45
Ξb\Xi^{\prime}_{b} (1/2+1/2^{+}) 5934 5935.02
Ξb\Xi^{\prime}_{b} (3/2+3/2^{+}) 5952 5953.82
Ωb\Omega_{b} (1/2+1/2^{+}) 6046 6046.10
Ωb\Omega_{b} (3/2+3/2^{+}) 6063 \ldots

To solve the Schro¨\ddot{\rm o}dinger 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 TQQT_{QQ} tetraquarks with the total spin (J,M)(J,M) is expressed as

ΨJM\displaystyle\Psi_{JM} =cβCc,β[[ϕνl(c)(𝒓c)ψNL(c)(𝑹c)]Λ\displaystyle=\sum_{c}\sum_{\beta}C_{c,\beta}\left[[\phi_{\nu l}^{(c)}({\bm{r}}_{c})\psi_{NL}^{(c)}({\bm{R}}_{c})]_{\Lambda}\right. (25)
[[χ1/2(Q1)χ1/2(Q2)]sχsd¯(=0,1)(d¯)]σ]JM,\displaystyle\left.\otimes\left[[\chi_{1/2}(Q_{1})\chi_{1/2}(Q_{2})]_{s}\chi_{s_{\bar{d}}(=0,1)}(\bar{d})\right]_{\sigma}\right]_{JM},

where χ1/2\chi_{1/2} and χsd¯\chi_{s_{\bar{d}}} stand for the spin wave function of the heavy quark and the antidiquark, respectively. ϕ\phi and ψ\psi are the Gaussian basis functions, which show the spatial wave function for 𝒓c{\bm{r}}_{c} and 𝑹c{\bm{R}}_{c} parts of Jacobi coordinates with the coefficients Cc,βC_{c,\beta}. The index β\beta denotes the quantum numbers related to this calculation method, β={ν,N,l,L,Λ,s,σ}\beta=\{\nu,N,l,L,\Lambda,s,\sigma\}, where ν\nu and NN are the number of basis functions, l,Ll,L and Λ\Lambda are the orbital angular momentum, ss and σ\sigma are the spin of two heavy quarks and all constituent particles, respectively. Using the Gaussian expansion method, the coefficients Cc,βC_{c,\beta} are determined by the Rayleigh-Ritz variational principle.

In calculating the spectrum of TQQT_{QQ} tetraquarks with the antidiquark cluster, we only consider SS-wave and PP-wave for the orbital angular momentum. For the SS-wave states, we calculate the masses of states which the principal quantum number 𝒩=1\mathcal{N}=1 or 22. For the PP-wave excited states, we classify these states into three types based on the channel c=3c=3 of the Jacobi coordinate in Fig. 1: the ρ\rho-mode which is the excitation between two heavy quarks, the λ\lambda-mode which is that between the pair of two heavy quarks and an antidiquark, and the ξd\xi_{d}-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 TbbT_{bb} and TccT_{cc} tetraquarks, the quamtum numbers are restricted by the Pauli principle to s=0s=0 for the ρ\rho-mode states and s=1s=1 for the SS-wave and λ\lambda-mode states.

Table 6: Quantum numbers of the ground and excited states of the TQQT_{QQ} tetraquarks including a scalar or an axial-vector antidiquark. The asterisk of the rightmost column indicates the forbidden state of TbbT_{bb} and TccT_{cc} tetraquarks.
State 𝒩\mathcal{N} ll LL Antidiquark (JPJ^{P}) ss σ\sigma JPJ^{P}
1S1S 1 0 0 Scalar (0+)(0^{+}) 0 0 0+0^{+} *
1S1S 1 0 0 Scalar (0+)(0^{+}) 1 1 1+1^{+}
2S2S 2 0 0 Scalar (0+)(0^{+}) 0 0 0+0^{+} *
2S2S 2 0 0 Scalar (0+)(0^{+}) 1 1 1+1^{+}
ρ\rho 1 1 0 Scalar (0+)(0^{+}) 0 0 11^{-}
ρ\rho 1 1 0 Scalar (0+)(0^{+}) 1 1 0,1,20,1,2^{-} *
λ\lambda 1 0 1 Scalar (0+)(0^{+}) 0 0 11^{-} *
λ\lambda 1 0 1 Scalar (0+)(0^{+}) 1 1 0,1,20,1,2^{-}
1S1S 1 0 0 Axial-vector (1+)(1^{+}) 0 1 1+1^{+} *
1S1S 1 0 0 Axial-vector (1+)(1^{+}) 1 0 0+0^{+}
1S1S 1 0 0 Axial-vector (1+)(1^{+}) 1 1 1+1^{+}
1S1S 1 0 0 Axial-vector (1+)(1^{+}) 1 2 2+2^{+}
2S2S 2 0 0 Axial-vector (1+)(1^{+}) 0 1 1+1^{+} *
2S2S 2 0 0 Axial-vector (1+)(1^{+}) 1 0 0+0^{+}
2S2S 2 0 0 Axial-vector (1+)(1^{+}) 1 1 1+1^{+}
2S2S 2 0 0 Axial-vector (1+)(1^{+}) 1 2 2+2^{+}
ρ\rho 1 1 0 Axial-vector (1+)(1^{+}) 0 1 0,1,20,1,2^{-}
ρ\rho 1 1 0 Axial-vector (1+)(1^{+}) 1 0 11^{-} *
ρ\rho 1 1 0 Axial-vector (1+)(1^{+}) 1 1 0,1,20,1,2^{-} *
ρ\rho 1 1 0 Axial-vector (1+)(1^{+}) 1 2 1,2,31,2,3^{-} *
λ\lambda 1 0 1 Axial-vector (1+)(1^{+}) 0 1 0,1,20,1,2^{-} *
λ\lambda 1 0 1 Axial-vector (1+)(1^{+}) 1 0 11^{-}
λ\lambda 1 0 1 Axial-vector (1+)(1^{+}) 1 1 0,1,20,1,2^{-}
λ\lambda 1 0 1 Axial-vector (1+)(1^{+}) 1 2 1,2,31,2,3^{-}
Table 7: Quantum numbers of the excited states of the TQQT_{QQ} tetraquarks including a pseudoscalar or a vector antidiquark.The asterisk of the rightmost column indicates the forbidden state of TbbT_{bb} and TccT_{cc} tetraquarks.
State 𝒩\mathcal{N} ll LL Antidiquark (JPJ^{P}) ss σ\sigma JPJ^{P}
ξP\xi_{P} 1 0 0 Pseudoscalar (0)(0^{-}) 0 0 00^{-} *
ξP\xi_{P} 1 0 0 Pseudoscalar (0)(0^{-}) 1 1 11^{-}
ξPρ\xi_{P}\rho 1 1 0 Pseudoscalar (0)(0^{-}) 0 0 1+1^{+}
ξPρ\xi_{P}\rho 1 1 0 Pseudoscalar (0)(0^{-}) 1 1 0,1,2+0,1,2^{+} *
ξPλ\xi_{P}\lambda 1 0 1 Pseudoscalar (0)(0^{-}) 0 0 1+1^{+} *
ξPλ\xi_{P}\lambda 1 0 1 Pseudoscalar (0)(0^{-}) 1 1 0,1,2+0,1,2^{+}
ξV\xi_{V} 1 0 0 Vector (1)(1^{-}) 0 1 11^{-} *
ξV\xi_{V} 1 0 0 Vector (1)(1^{-}) 1 0 00^{-}
ξV\xi_{V} 1 0 0 Vector (1)(1^{-}) 1 1 11^{-}
ξV\xi_{V} 1 0 0 Vector (1)(1^{-}) 1 2 22^{-}
ξVρ\xi_{V}\rho 1 1 0 Vector (1)(1^{-}) 0 1 0,1,2+0,1,2^{+}
ξVρ\xi_{V}\rho 1 1 0 Vector (1)(1^{-}) 1 0 1+1^{+} *
ξVρ\xi_{V}\rho 1 1 0 Vector (1)(1^{-}) 1 1 0,1,2+0,1,2^{+} *
ξVρ\xi_{V}\rho 1 1 0 Vector (1)(1^{-}) 1 2 1,2,3+1,2,3^{+} *
ξVλ\xi_{V}\lambda 1 0 1 Vector (1)(1^{-}) 0 1 0,1,2+0,1,2^{+} *
ξVλ\xi_{V}\lambda 1 0 1 Vector (1)(1^{-}) 1 0 1+1^{+}
ξVλ\xi_{V}\lambda 1 0 1 Vector (1)(1^{-}) 1 1 0,1,2+0,1,2^{+}
ξVλ\xi_{V}\lambda 1 0 1 Vector (1)(1^{-}) 1 2 1,2,3+1,2,3^{+}

III Results

III.1 Spectrum of TbbT_{bb} and TccT_{cc} tetraquarks

Refer to caption
Figure 2: The energy spectra of nonstrange TbbT_{bb} and TccT_{cc} tetraquarks with the flavor 𝟑\bm{3} antidiquarks. The colors of lines show the types of states and constituent antidiquarks, such as red for the SS-wave state with the scalar antidiquark, green for the PP-wave state with the scalar antidiquark, blue for the ones with the pseudoscalar antidiquark, and magenta for the ones with the vector antidiquark. The dashed and dotted lines are the thresholds relevant to the strong decay of the ground state, where the black-dashed ones are calculated from the experimental values of the meson masses in Ref. Zyla et al. (2020) and the blue-dotted ones are from the meson masses calculated in the AP1AP1 potential model, Eqs. (20) and (21).
Refer to caption
Figure 3: The energy spectra of strange TbbT_{bb} and TccT_{cc} tetraquarks with the flavor 𝟑\bm{3} antidiquarks. The notations are the same as Fig. 3.
Refer to caption
Figure 4: The energy spectra of TbbT_{bb} tetraquarks with the flavor 𝟔¯\bar{\bm{6}} antidiquark. The red and green states are the SS-wave and PP-wave states with the axial-vector antidiquark, respectively. The other notations are the same as Fig. 3.
Refer to caption
Figure 5: The energy spectra of TccT_{cc} tetraquarks with the flavor 𝟔¯\bar{\bm{6}} antidiquark. The other notations are the same as Fig. 5.

Here we discuss the spectra of the TbbT_{bb} and TccT_{cc} 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 S=0S=0 and +1+1 tetraquarks, respectively. Here the flavor of all the constituent antidiquarks is 𝟑\bm{3}, 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 SS-wave states and the PP-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 ξP\xi_{P}-mode states become lighter than those of 2S2S states and approach to the lightest PP-wave states (ρ\rho- and λ\lambda-mode states). This is caused by the inverse mass hierarchy of pseudoscalar diquarks in Eq. (11). Moreover, all the tetraquarks containing the ξV\xi_{V}-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 1+1^{+} ground states is given by the sum of JP=0J^{P}=0^{-} and 11^{-} singly heavy mesons (BB or DD). Note that the 1+1^{+} ground state is not allowed to decay into two 00^{-} 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 AP1AP1 in Eqs. (20) and (21) (blue-dotted lines). One sees that the 1S1S ground states of TbbT_{bb} tetraquarks are below the B(s)BB_{(s)}-B^{*} thresholds, so that they are bound states and do not decay by the strong interaction. From the experimental thresholds, their binding energies are 115115 MeV for the Tbb;u¯d¯T_{bb;\bar{u}\bar{d}} and 2828 MeV for the Tbb;n¯s¯T_{bb;\bar{n}\bar{s}}. On the other hand, no bound state appears in the spectra of TccT_{cc} 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 TbbT_{bb} and TccT_{cc} tetraquarks, respectively. The red and green lines stand for the SS-wave states and the PP-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 TQQ;n¯n¯T_{QQ;\bar{n}\bar{n}} and TQQ;n¯s¯T_{QQ;\bar{n}\bar{s}} is almost equal to that between TQQ;n¯s¯T_{QQ;\bar{n}\bar{s}} and TQQ;s¯s¯T_{QQ;\bar{s}\bar{s}}. 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 0+0^{+} tetraquarks, and are the total mass of two singly heavy mesons with JP=0J^{P}=0^{-}. Although the 1S1S ground state with JP=0+J^{P}=0^{+} 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 (MQiM_{Q_{i}}\rightarrow\infty), 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 1S1S and 2S2S states of tetraquarks with JP=0+,1+J^{P}=0^{+},1^{+}, and 2+2^{+} have the HQS triplet structure. Comparing these structures in Figs. 5 and 5, we can see that the mass differences for TbbT_{bb} tetraquarks are smaller than those for TccT_{cc} tetraquarks, which is caused by the bottom quark being heavier than the charm quark.

III.2 Spectrum of TcbT_{cb} tetraquarks

Refer to caption
Figure 6: The energy spectra of nonstrange TcbT_{cb} tetraquarks with the flavor 𝟑\bm{3} antidiquarks, classified by the total spin of heavy quarks, s=0s=0, or 11. The other notations are the same as Fig. 3.
Refer to caption
Figure 7: The energy spectra of strange TcbT_{cb} tetraquarks with the flavor 𝟑\bm{3} antidiquarks, classified by the total spin of heavy quarks, s=0s=0, or 11. The other notations are the same as Fig. 3.
Refer to caption
Figure 8: The energy spectra of TcbT_{cb} tetraquarks with the flavor 𝟔¯\bar{\bm{6}} antidiquarks, for the total spin of heavy quarks, s=0s=0. The other notations are the same as Fig. 5.
Refer to caption
Figure 9: The energy spectra of TcbT_{cb} tetraquarks with the flavor 𝟔¯\bar{\bm{6}} antidiquarks, for the total spin of heavy quarks, s=1s=1. The other notations are the same as Fig. 5.

We here discuss the spectra of TcbT_{cb} tetraquarks. Since two constituent heavy quarks are different from each other, the spin of two heavy quarks, ss, can take two kinds of values as s=0s=0 and 11. 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 ss. Figs. 7 and 7 are the spectra of the strangeness S=0S=0 and +1+1 tetraquarks with a flavor 𝟑\bm{3} 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 TcbT_{cb} tetraquarks with s=0s=0 and 11, including a flavor 𝟔¯\bar{\bm{6}} axial-vector antidiquark, respectively. The two colors of lines stand for the SS-wave states and the PP-wave states, which is the same as Figs. 5 and 5.

The spectra of TcbT_{cb} tetraquarks are similar to those of TbbT_{bb} and TccT_{cc} tetraquarks when the constituent antidiquarks are the same. Although there are less states for the TcbT_{cb} tetraquarks with s=0s=0, 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 PP-wave states have the same spin quantum numbers, such as ρ\rho and λ\lambda modes with s=σ=0s=\sigma=0 or 1. These states may be mixed by the spin-spin interaction in TcbT_{cb}. (Note that the mixing is not allowed in TbbT_{bb} and TccT_{cc} due to the Pauli principle.) The subscripts of ρ\rho and λ\lambda characters mean the probabilities of each state, calculated by the two-body density distribution between two heavy quarks. For example, about two PP-wave states with green lines in Fig. 7, the ρ\rho- and λ\lambda-mode states mix each other. As a result, in the lower state, the probability of the ρ\rho-mode state becomes 68.5%68.5\%, and the remained 31.5%31.5\% is that of the λ\lambda-mode state.

As with the TbbT_{bb} and TccT_{cc} 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 TcbT_{cb} tetraquarks except the Tcb;u¯d¯T_{cb;\bar{u}\bar{d}} tetraquark with s=1s=1 at 3 MeV below the AP1 threshold in Fig. 7. Therefore, we conclude that most of the TcbT_{cb} tetraquarks with the antidiquark cluster are unstable in strong decays.

III.3 Binding energies of lowest TQQT_{QQ} tetraquark states

Table 8: The binding energies of the TQQT_{QQ} tetraquark ground states including a scalar antidiquark, which are based on two types of thresholds: type exp from the experimental masses of mesons in Ref. Zyla et al. (2020), and type AP1 from the theoretical masses of mesons given by the AP1AP1 potential model. The character II shows for the isospin.
Tetraquark State Binding energy [MeV]
TQQ;q¯q¯T_{QQ;\bar{q}\bar{q}} I(JP)I(J^{P}) type exp type AP1
Tbb;u¯d¯T_{bb;\bar{u}\bar{d}} 0(1+)0(1^{+}) 115115 189189
Tbb;n¯s¯T_{bb;\bar{n}\bar{s}} 1/2(1+)1/2(1^{+}) 2828 5959
Tcc;u¯d¯T_{cc;\bar{u}\bar{d}} 0(1+)0(1^{+}) (86)(-86) (47)(-47)
Tcc;n¯s¯T_{cc;\bar{n}\bar{s}} 1/2(1+)1/2(1^{+}) (166)(-166) (153)(-153)
Tcb;u¯d¯T_{cb;\bar{u}\bar{d}} 0(0+)0(0^{+}) (81)(-81) (32)(-32)
Tcb;n¯s¯T_{cb;\bar{n}\bar{s}} 1/2(0+)1/2(0^{+}) (170)(-170) (165)(-165)
Tcb;u¯d¯T_{cb;\bar{u}\bar{d}} 0(1+)0(1^{+}) (56)(-56) 33
Tcb;n¯s¯T_{cb;\bar{n}\bar{s}} 1/2(1+)1/2(1^{+}) (143)(-143) (124)(-124)

We summarize the binding energies of the lowest-lying states in Table 8. Two lowest states of TbbT_{bb} tetraquarks are well below the threshold and expected to be stable.

On the other hand, we have no bound TccT_{cc} state. It may seem contradictory to the recent observation of Tcc+T_{cc}^{+}, at about 300 keV below the D+D0D^{*+}-D^{0} 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 D+D0D^{*+}-D^{0} state and thus have a molecule-like structure. In contrast, the tetraquark in our diquark picture does not dissociate into DDD-D, and therefore it does not couple to loosely bound D()DD^{(*)}-D 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 TQQT_{QQ} tetraquark states given by some previous works. Most of the studies in this table predicted that the non-strange and singly antistrange TbbT_{bb} tetraquarks, Tbb;u¯d¯T_{bb;\bar{u}\bar{d}} and Tbb;n¯s¯T_{bb;\bar{n}\bar{s}}, are stable, which is consistent with our results. On the other hand, the other tetraquark systems seem to be unstable or uncertain.

Table 9: Comparison between theoretical studies for the stability of doubly heavy tetraquarks. “S”, “US”, and “ND” denote “stable” (below the threshold), “unstable” (above the threshold), and “not determined” (in other words, not conclusive within theoretical uncertainties), respectively. For a similar summary, see Ref. Cheng et al. (2021).
Reference Tcc;u¯d¯T_{cc;\bar{u}\bar{d}} Tcc;n¯s¯T_{cc;\bar{n}\bar{s}} Tcc;s¯s¯T_{cc;\bar{s}\bar{s}} Tbb;u¯d¯T_{bb;\bar{u}\bar{d}} Tbb;n¯s¯T_{bb;\bar{n}\bar{s}} Tbb;s¯s¯T_{bb;\bar{s}\bar{s}} Tcb;u¯d¯T_{cb;\bar{u}\bar{d}} Tcb;n¯s¯T_{cb;\bar{n}\bar{s}} Tcb;s¯s¯T_{cb;\bar{s}\bar{s}}
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 TbbT_{bb} bound states

Now we discuss the density distributions between two constituent particles in each of the TbbT_{bb} tetraquark bound states. The two-body density distribution is expressed by the variational wave function in Eq. (25) as

ρ(rc)=𝑑𝒓^c𝑑𝑹c|ΨJM(𝒓c,𝑹c)|2,\displaystyle\rho(r_{c})=\int d\hat{\bm{r}}_{c}d{\bm{R}}_{c}|\Psi_{JM}({\bm{r}}_{c},{\bm{R}}_{c})|^{2}, (26)

where rc=|𝒓c|r_{c}=|{\bm{r}}_{c}| and 𝒓^c\hat{\bm{r}}_{c} are the distance and the angular parts of the relative coordinate for two particles in channel cc, respectively. From this equation (26), we can give the density distribution between two bottom quarks with channel c=3c=3, and that between a bottom quark and a scalar antidiquark with channel c=1,2c=1,2. The functions r2ρ(r)r^{2}\rho(r) of each pair are shown in Fig. 10 for both TbbT_{bb} tetraquarks with the strangeness S=0S=0 and +1+1.

Refer to caption
Figure 10: Density distributions between two constituent bottom quarks (bbb-b) and between a bottom quark and an antidiquark (bd¯b-\bar{d}) inside the bound states of TbbT_{bb} tetraquarks. The solid lines are for the strangeness S=0S=0 tetraquark, and the dashed lines are for the strangeness S=+1S=+1 tetraquark, respectively.

We see that the distance between the pair of two bottom quarks (bbb-b) is nearer than the pair of a bottom quark and an antidiquark (bd¯b-\bar{d}) for both strangeness S=0S=0 and +1+1. 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 SS, the red solid line and the blue dashed line are almost same so that there is no difference between the pair bbb-b. On the other hand, for between the pair bd¯b-\bar{d}, 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.

Table 10: The rms distances of the TbbT_{bb} tetraquark bound states.
Tetraquark State rms distance r^2\sqrt{\hat{r}^{2}} [fm]
Tbb;q¯q¯T_{bb;{\bar{q}}{\bar{q}}} I(JP)I(J^{P}) bb~{}~{}{b-b}~{}~{} bd¯{b-\bar{d}}
Tbb;u¯d¯T_{bb;{\bar{u}}{\bar{d}}} 0(1+)0(1^{+}) 0.30 0.56
Tbb;n¯s¯T_{bb;{\bar{n}}{\bar{s}}} 1/2(1+)1/2(1^{+}) 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 ρ(r)\rho(r) as

r^c2=(rc2ρ(rc)rc2𝑑rc)/(ρ(rc)rc2𝑑rc).\displaystyle\sqrt{\hat{r}_{c}^{2}}=\sqrt{\left(\int r_{c}^{2}\rho(r_{c})r_{c}^{2}dr_{c}\right)/\left(\int\rho(r_{c})r_{c}^{2}dr_{c}\right)}. (27)

For the rms distance between a bottom quark and an antidiquark (bd¯b-\bar{d}), these values are similar to that between a bottom quark and a diquark inside the ground state of Λb\Lambda_{b} (0.55\simeq 0.55 fm) and Ξb\Xi_{b} (0.51\simeq 0.51 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 TQQT_{QQ} 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 Σ\langle\Sigma\rangle. It is interesting to see how the tetraquark masses are modified when the value of Σ\langle\Sigma\rangle 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 xx to Σ\langle\Sigma\rangle, where its range is 0x10\leq x\leq 1, we can express the masses of diquarks from the chiral symmetry restored phase (x=0x=0) to the ordinary vacuum state (x=1x=1).

For example, the masses of nonstrange diquarks are given by

Mud(0+)=mS02(x+ϵ)mS12x2mS22,\displaystyle M_{ud}(0^{+})=\sqrt{m^{2}_{S0}-(x+\epsilon)m^{2}_{S1}-x^{2}m^{2}_{S2}}, (28)
Mud(0)=mS02+(x+ϵ)mS12+x2mS22,\displaystyle M_{ud}(0^{-})=\sqrt{m^{2}_{S0}+(x+\epsilon)m^{2}_{S1}+x^{2}m^{2}_{S2}}, (29)
Mnn(1+)=mV02+x2(mV12+2mV22),\displaystyle M_{nn}(1^{+})=\sqrt{m^{2}_{V0}+x^{2}(m^{2}_{V1}+2m^{2}_{V2})}, (30)
Mud(1)=mV02+x2(mV12+2mV22),\displaystyle M_{ud}(1^{-})=\sqrt{m^{2}_{V0}+x^{2}(-m^{2}_{V1}+2m^{2}_{V2})}, (31)

in which only the mass of scalar diquark Mud(0+)M_{ud}(0^{+}) becomes heavier and the others become lighter with approaching to the chiral symmetry restored phase (x0)(x\rightarrow 0).

Refer to caption
Figure 11: Dependence of the masses of nonstrange TbbT_{bb} tetraquarks (left scale) on the chiral symmetry breaking strength xx. Dependence of the masses of the diquarks (right scale) are also shown for reference.
Refer to caption
Figure 12: Dependence of the masses of nonstrange TccT_{cc} tetraquarks on the chiral symmetry breaking strength xx.
Refer to caption
Figure 13: Dependence of the masses of nonstrange TcbT_{cb} tetraquarks with s=0s=0 on the chiral symmetry breaking strength xx.
Refer to caption
Figure 14: Dependence of the masses of nonstrange TcbT_{cb} tetraquarks with s=1s=1 on the chiral symmetry breaking strength xx.

Figs. 11, 12, 13, and 14 show the change of masses of TQQT_{QQ} tetraquarks by the factor xx, which are all 1S1S 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 x=0.6x=0.6 and are reversed, the crossing of masses of TQQT_{QQ} tetraquarks also occurs near x=0.6x=0.6. The tetraquarks including an axial-vector antidiquark become lighter than those including a scalar antidiquark. In other words, the isospin of stable TQQT_{QQ} tetraquarks changes from 0 to 11 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, TbbT_{bb} tetraquarks with a constituent axial-vector antidiquark are below the thresholds when the parameter xx is less than 0.60.6.

In the chiral symmetry restored phase, the masses of nonstrange vector diquark and axial-vector diquark are degenerate with the mass value mV0m_{V0} 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 ξV\xi_{V}-mode are below the thresholds for the nonstrange TbbT_{bb} tetraquark. On the other hand, the masses of nonstrange pseudoscalar diquark and scalar diquark are not degenerate due to the UA(1)U_{A}(1) anomaly, and the pseudoscalar diquark is about 300 MeV heavier than the scalar diquark even at x=0x=0 Kim et al. (2021). Thus there are no ξP\xi_{P}-mode state below the thresholds for the nonstrange TbbT_{bb} tetraquark.

IV Summary

In this paper, we have studied the spectrum of doubly heavy TQQT_{QQ} 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 SS-wave and PP-wave TQQT_{QQ} 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 TQQT_{QQ} tetraquarks including an antidiquark, we have obtained the following:

  1. (i)

    For the ground states of the TQQT_{QQ} tetraquarks, we have found that the TbbT_{bb} tetraquarks including a scalar antidiquark are stable in strong decays with the binding energies, 115 MeV for Tbb;u¯d¯T_{bb;\bar{u}\bar{d}} and 28 MeV for Tbb;n¯s¯T_{bb;\bar{n}\bar{s}}. No other states, such as TccT_{cc} and TcbT_{cb} tetraquarks, are significantly below the two-meson thresholds and thus not stable.

  2. (ii)

    For the excited states of the TQQT_{QQ} tetraquarks, we have found the inverse mass hierarchy in states with the ξP\xi_{P}-mode and the enhanced mass hierarchy in states with the ξV\xi_{V}-mode. As a result, for the singly antistrange tetraquarks, there are one more PP-wave states between 1S1S and 2S2S states compared with the nonstrange tetraquarks.

  3. (iii)

    For two stable bound states of TbbT_{bb} 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 TbbT_{bb} tetraquarks are similar to each other and have less difference by the existence of an strange antiquark.

  4. (iv)

    For the ground states of the nonstrange TQQT_{QQ} 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 xx, Eqs. (28) – (31). We found that the axial-vector antidiquark makes the TQQT_{QQ} 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 TQQT_{QQ} 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 𝟑¯\bar{\bm{3}}, so that the energy spectra of TQQT_{QQ} tetraquarks including such a diquark are related only to the color 𝟑𝟑¯{\bm{3}}-\bar{\bm{3}} interaction between the light antidiquark and the heavy diquark. On the other hand, another type is the color 𝟔\bm{6} diquark. If we consider the color 𝟔\bm{6} diquarks in the chiral effective theory of diquarks, the energy spectra of TQQT_{QQ} tetraquarks is related also to the color 𝟔¯𝟔\bar{\bm{6}}-{\bm{6}} 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 TQQT_{QQ} tetraquarks including the color 𝟔¯\bar{\bm{6}} antidiquark would appear above the ground states given by the color 𝟑\bm{3} scalar antidiquarks. Also, tensor diquarks with color 𝟑¯\bar{\bm{3}} and flavor 𝟔\bm{6}, or with color 𝟔\bm{6} and flavor 𝟑¯\bar{\bm{3}}, would be the new source for the excited states of TQQT_{QQ} 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