Review of the QCD sum rules for exotic states

Zhi-Gang Wang

Front. Phys. ›› 2026, Vol. 21 ›› Issue (1) : 016300

PDF (12138KB)
Front. Phys. ›› 2026, Vol. 21 ›› Issue (1) : 016300 DOI: 10.15302/frontphys.2026.016300
REVIEW ARTICLE

Review of the QCD sum rules for exotic states

Author information +
History +
PDF (12138KB)

Abstract

We review the exotic states, such as the X, Y, Z, T and P states, and present their possible assignments based on the QCD sum rules. We present many predictions which can be confronted to the experimental data in the future to diagnose the exotic states. Furthermore, we also mention other theoretical methods.

Graphical abstract

Keywords

QCD / sum rules / exotic states

Cite this article

Download citation ▾
Zhi-Gang Wang. Review of the QCD sum rules for exotic states. Front. Phys., 2026, 21(1): 016300 DOI:10.15302/frontphys.2026.016300

登录浏览全文

4963

注册一个新账户 忘记密码

1 Introduction

In 1964, Gell−Mann suggested that multiquark states beyond the minimal valence quark constituents qq¯ and qqq might exist [1], a quantitative model for the tetraquark states with the quark constituents qqq¯q¯ was developed by Jaffe using the MIT bag model in 1977 [2, 3]. Later, the five-quark baryons with the quark constituents qqqqq¯ were developed [4], while the name pentaquark was introduced by Lipkin [5]. Also in 1964, Dyson and Xuang studied the dibaryon or six-quark states based on the SU(6) symmetry [6], for more literatures on this subject, one can consult Refs. [7, 8]. The QCD allows the existence of multiquark states and hybrid states which contain not only quarks but also gluonic degrees of freedom [9-11].

Before observation of the X(3872) by the Belle Collaboration in 2003 [12], the most promising and most hot subject is the nature of the light mesons below 1GeV, are they traditional 3P0 states, tetraquark states or molecular states? [2, 3, 13-16]. In fact, the observation of the X(3872) stimulates more motivations and curiosities in exploring the nature of the light scalar mesons [17-29], for more references, see the reviews [30, 31, 32]. However, it is an un-resoled problem until now. There have been several excellent reviews of the exotic states with emphasis on different aspects [33-47].

Firstly, let us see the experimental data on the exotic states pedagogically and dogmatically in the order X, Y, Z, T and P sequentially, and sort out of the exotic states according to the masses from low to high roughly. In fact, there are relations among those X, Y and Z states in one way or the other, it is impossible to sort out them distinctly, the most related ones are grouped together into one sub-section. Furthermore, we would like to emphasize the assignment based on the QCD sum rules at the end of every sub-section if there exists a room, as the predicted spectroscopy based on the QCD sum rules cannot accommodate all the exotic states. Other possible assignments would be presented in Sections 3−5 and 6.1.

1.1 X(3872)

In 2003, the Belle Collaboration observed a narrow charmonium-like state X(3872) with the mass 3872.0±0.6±0.5MeV near the DD¯ threshold in the π+πJ/ψ mass spectrum in the exclusive processes B±K±π+πJ/ψ [12]. The evidences for the decay modes X(3872)γJ/ψ,γψ observed by the Belle and BaBar Collaborations imply the positive charge conjugation C=+ [48-50]. Angular correlations between final state particles π+πJ/ψ analyzed by the CDF, Belle and LHCb Collaborations favor the JPC=1++ assignment [51-54]. It is a possible candidate for the tetraquark state [55-62], (not) molecular state ([63]) [64-84], (not) traditional charmonium χc1(2P) ([85]) [86-88], threshold cusp [89], etc. However, none of those available assignment has won an overall consensus, its nature is still under heated debates, the very narrow width and exotic branching fractions make it a hot potato, the door remains open for another binding mechanism. The QCD sum rules allow both the color 3¯3-type and 11-type tetraquark assignments, see Tab.9 in Section 3.1.1 and Tab.44 in Section 4.1.

1.2 X(3915), X(3940), Y(3940), Z(3930), X(4160), X(3860), X(3960)

In 2005, the Belle Collaboration observed the Y(3940) in the B-decays BKY(3940)KωJ/ψ [90], later, the BaBar Collaboration confirmed the Y(3940) in the ωJ/ψ mass spectrum with a mass about 3.915GeV in the decays BKωJ/ψ [91, 92].

In 2007, the Belle Collaboration observed the X(3940) in the process e+eJ/ψX(3940) with the subprocess X(3940)DD¯ [93]. In 2008, the Belle Collaboration confirmed the X(3940) in the same process, and observed the X(4160) in the subprocess X(4160)DD¯ [94].

In 2010, the X(3915) was observed in the process γγJ/ψω by the Belle Collaboration [95], then the BaBar Collaboration determined its quantum numbers JP=0+ [96], it is a good candidate for the conventional charmonium χc0(2P). The Y(3940) and X(3915) could be the same particle, and is denoted as the χc0(3915) with the assignment JPC=0++ in The Review of Particle Physics [97].

In 2006, the Z(3930) was observed in the process γγJ/ψω by the Belle Collaboration [98], then confirmed by the BaBar Collaboration in the same process [99], the two Collaborations both determined its quantum numbers to be JPC=2++, and it is widely assigned as the χc2(3930) [97].

In 2017, the Belle Collaboration performed a full amplitude analysis of the process e+eJ/ψDD¯, and observed a new charmonium-like state X(3860) which decays to the DD¯ pair, the measured mass and width are 386232+2613+40MeV and 20167+15482+88MeV, respectively [100]. The JPC=0++ hypothesis is favored over the 2++ hypothesis at the level of 2.5σ, and the Belle Collaboration assigned the X(3860) as an alternative χc0(2P) state [100].

In 2020, the LHCb Collaboration performed an amplitude analysis of the B+D+DK+ decays and observed that it is necessary to include the χc0(3930) and χc2(3930) with the JPC=0++ and 2++ respectively in the D+D channel, and to include the X0(2900) and X1(2900) with the JP=0+ and 1 respectively in the DK+ channel [101, 102]. The measured Breit−Wigner masses and widths are

χc0(3930):M=3923.8±1.5±0.4MeV,Γ=17.4±5.1±0.8MeV,χc2(3930):M=3926.8±2.4±0.8MeV,Γ=34.2±6.6±1.1MeV,

and the χc0(3915) and χc0(3930) can be identified as the χc0(2P).

In 2023, the LHCb Collaboration announced the observation of the X(3960) in the Ds+Ds mass spectrum in the B+Ds+DsK+ decays, and the assignment JPC=0++ is favored [103], the measured Breit−Wigner mass and width are 3956±5±10 MeV and 43±13±8 MeV, respectively.

The thresholds of the Ds+Ds and D+D are 3938MeV and 3739MeV, respectively, which favors assigning the X(3960) and χc0(3930) as the same particle. However, the ratio of the branching fractions [103],

Γ(XD+D)Γ(XDs+Ds)=0.29±0.09±0.10±0.08,

implies the exotic nature of this state, as it is harder to excite an ss¯ pair from the vacuum compared with the uu¯ or dd¯ pair, and the traditional charmonium states predominantly decay into the DD¯ and DD¯ states rather than into the DsD¯s and DsD¯s states. In addition, there is no room for the X(3860), so, at least one of the X(3860), X(3915), χc0(3930) and X(3960) should be exotic state [104-107].

For example, the updated nonrelativistic potential model (NR) and Godfrey−Isgur relativized potential model (GI) indicate that the 2P charmonium states have the masses (NR; GI)[108],

χc2(2P):3972MeV,3979MeV,χc1(2P):3925MeV,3953MeV,χc0(2P):3852MeV,3916MeV,

even the assignments of the χc0(2P) and χc2(2P) are not comfortable enough.

We can identify the X(3915), Y(3940) and χc0(3930) as the same particle tentatively, and assign it as the color 3¯3-type scalar tetraquark state based on the QCD sum rules, see Tab.9 in Section 3.1.1, furthermore, there exists a room for the X(3860). On the other hand, we can assign the X(3960) as the color 3¯3-type or 11-type tetraquark state, see Tab.10 in Section 3.1.1 and Tab.44 in Section 4.1.

1.3 ηc(3945), hc(4000), χc1(4010), hc(4300)

In 2024, the LHCb Collaboration explored the decays B+D+DK+ and DD+K+, and observed four charmonium(-like) states ηc(3945), hc(4000), χc1(4010) and hc(4300) with the quantum numbers JPC=0+, 1+, 1++ and 1+ respectively in the D±D mass spectrum [109]. The measured Breit−Wigner masses and widths are

ηc(3945):M=394517+2828+37MeV,Γ=13049+9270+101MeV,hc(4000):M=400014+1722+29MeV,Γ=18445+7161+97MeV,χc1(4010):M=4012.53.9+3.63.7+4.1MeV,Γ=62.76.4+7.06.6+6.4MeV,hc(4300):M=4307.36.6+6.44.1+3.3MeV,Γ=5816+2825+28MeV.

The updated nonrelativistic potential model (NR) and Godfrey−Isgur relativized potential model (GI) indicate that the 3S/2P/3P charmonium states have the masses (NR; GI) [108],

ηc(3S):4043MeV,4064MeV,χc1(2P):3925MeV,3953MeV,χc1(3P):4271MeV,4317MeV,hc(2P):3934MeV,3956MeV,hc(3P):4279MeV,4318MeV,

the assignments ηc(3945), hc(4000) and χc1(4010) in Ref. [109] are rather marginal.

Based on the predictions of the QCD sum rules, we can assign the hc(4000) and χc1(4010) as the color 3¯3-type tetraquark states tentatively based on the QCD sum rules, see Tab.9 in Section 3.1.1.

1.4 X(4140), X(4274), X(4500/4475), X(4700/4710), X(4685/4650), X(4630/4800)

In 2009, the CDF Collaboration observed the X(4140) in the J/ψϕ mass spectrum in the B+J/ψϕK+ decays with a significance larger than 3.8σ [110]. In 2011, the CDF Collaboration confirmed the Y(4140) in the B±J/ψϕK± decays with a significance greater than 5σ, and observed an evidence for the X(4274) with an approximate significance of 3.1σ [111]. In 2013, the CMS Collaboration also confirmed the X(4140) in the B±J/ψϕK± decays [112].

In 2016, the LHCb Collaboration performed the first full amplitude analysis of the B+J/ψϕK+ decays, confirmed the X(4140) and X(4274) in the J/ψϕ mass spectrum, and determined the spin-parity to be JP=1+ [113, 114]. Moreover, the LHCb Collaboration observed two new particles X(4500) and X(4700) in the J/ψϕ mass spectrum, and determined the spin-parity to be JP=0+ [113, 114]. The measured masses and widths are

X(4140):M=4146.5±4.52.8+4.6MeV,Γ=83±2114+21MeV,X(4274):M=4273.3±8.33.6+17.2MeV,Γ=56±1111+8MeV,X(4500):M=4506±1115+12MeV,Γ=92±2120+21MeV,X(4700):M=4704±1024+14MeV,Γ=120±3133+42MeV.

If they are tetraquark states, their quark constituents must be csc¯s¯. The S-wave J/ψϕ systems have the JPC=0++, 1++, 2++, while the P-wave J/ψϕ systems have the JPC=0+, 1+, 2+, 3+. The LHCb’s data rule out the 0++ or 2++ Ds+Ds molecule assignments.

In 2021, the LHCb Collaboration performed an improved full amplitude analysis of the exclusive process B+J/ψϕK+, observed the X(4685) (X(4630)) in the J/ψϕ mass spectrum with the JP=1+ (1) and Breit-Wigner masses and widths,

X(4685):M=4684±716+13MeV,Γ=126±1541+37MeV,X(4630):M=4626±16110+18MeV,Γ=174±2773+134MeV.

Furthermore, they observed the Zcs(4000) and Zcs(4220) with the JP=1+ in the J/ψK+ mass spectrum and confirmed the four old particles [115].

In 2024, the LHCb Collaboration performed the first full amplitude analysis of the decays B+ψ(2S)K+π+π, and they developed an amplitude model with 53 components comprising 11 hidden-charm exotic states, for example, the X(4475), X(4650), X(4710) and X(4800) in the ψ(2S)ρ0(770) mass spectrum with the JP=0+, 1+, 0+ and 1, respectively, while the X(4800) is just an effective description of generic partial wave-function [116].

The X(4475), X(4650), X(4710) and X(4800) have the isospin (I,I3)=(1,0), while the X(4500), X(4685), X(4700) and X(4630) have the isospin (I,I3)=(0,0) according to the final states ψ(2S)ρ0(770) and J/ψϕ. A possible explanation is that those states are genuinely different states, if the X(4475) state is the cc¯(uu¯dd¯) isospin partner of the X(4500) interpreted as the cc¯ss¯ state, we would generally expect a larger mass difference of MX(4500)MX(4475)200MeV rather than several MeV. The un-normal light-flavor SU(3) breaking effects make them difficult to assign in the scenario of tetraquark states.

Based on the predictions of the QCD sum rules, we can assign the X(4140), X(4274), X(4500), X(4685) and X(4700) as the color 3¯3-type tetraquark states with the positive parity tentatively, see Tab.10 in Section 3.1.1, and assign the X(4630) as the color 3¯3-type tetraquark state with the negative parity tentatively, see Tab.23 in Section 3.1.3.

1.5 X(4350)

In 2010, the Belle Collaboration measured the process γγϕJ/ψ for the ϕJ/ψ mass distributions and observed a narrow peak of 8.83.2+4.2 events with a significance of 3.2σ [117]. The mass and width are (4350.65.1+4.6±0.7)MeV and (13.39.1+17.9±4.1)MeV respectively. However, the X(4350) is not confirmed by other experiments.

1.6 X0(2900), X1(2900), Tcs¯(2900)

In 2020, the LHCb Collaboration reported a narrow peak in the DK+ invariant mass spectrum in the decays B±D+DK± [101, 102]. The peak could be reasonably parameterized in terms of two Breit−Wigner resonances:

X0(2900):JP=0+,M=2866±7±2MeV,Γ=57±12±4MeV,X1(2900):JP=1,M=2904±5±1MeV,Γ=110±11±4MeV.

They are the first exotic hadrons with fully open flavor, the valence quarks are udc¯s¯ [101, 102]. The narrow peak can be assigned as the color 3¯3-type tetraquark state with the JP=0+ [118-121], its radial/orbital excitation [122], non-tetraquark state [123], DK¯ molecular state [121, 124-128], triangle singularity [129], etc.

In 2023, the LHCb Collaboration observed the tetraquark candidates Tcs¯0/++(2900) with the spin-parity JP=0+ in the processes B+DDs+π+ and B0D¯0Ds+π with the significance larger than 9σ [130, 131]. The measured Breit−Wigner masses and widths are

Tcs¯0(2900):M=2.892±0.014±0.015GeV,Γ=0.119±0.026±0.013GeV,Tcs¯++(2900):M=2.921±0.017±0.020GeV,Γ=0.137±0.032±0.017GeV,

respectively, and they belong to a new type of open-charm tetraquark states with the c and s¯ quarks. The X¯0(2900), Tcs¯0(2900) and Tcs¯++(2900) can be accommodated in the light flavor SU(3)F symmetry sextet.

We can assign the X0(2900), Tcs¯0(2900) and Tcs¯++(2900) as the color 3¯3-type tetraquark states with the JP=0+ tentatively based on the QCD sum rules, see Section 6.1.

1.7 X(5568)

In 2016, the D0 Collaboration observed a narrow structure X(5568) in the decays X(5568)Bs0π± with significance of 5.1σ [132]. The mass and natural width are 5567.8±2.91.9+0.9MeV and 21.9±6.42.5+5.0MeV, respectively. The Bs0π± systems consist of two quarks and two antiquarks of four different flavors, just like the Tcs¯(2900) with the JP=0+ observed 7 years later in the Ds+π+/Ds+π mass spectrum by the LHCb Collaboration [130, 131]. The D0 Collaboration fitted the Bs0π± systems with the S-wave Breit-Wigner parameters, the favored assignments are JP=0+, but the assignments JP=1+ cannot be excluded according to decays X(5568)Bsπ+Bs0π+γ, where the low-energy photon is not detected. It can be assigned as a usb¯d¯ tetraquark state with the JPC=0++ [133-137], however, the X(5568) is not confirmed by the LHCb, CMS, ATLAS and CDF Collaborations [138-141].

1.8 X(6600), X(6900), X(7300)

In 2020, the LHCb Collaboration reported evidences of two fully-charm tetraquark candidates in the J/ψJ/ψ mass spectrum [142]. They observed a broad structure above the J/ψJ/ψ threshold ranging from 6.2 to 6.8 GeV and a narrow structure at about 6.9 GeV with the significance of larger than 5σ. In addition, they also observed some vague structures around 7.2 GeV.

In 2023, the ATLAS Collaboration observed statistically significant excesses in the J/ψJ/ψ channel, which are consistent with a narrow resonance at about 6.9GeV and a broader structure at much lower mass. And they also observed a statistically significant excess at about 7.0GeV in the J/ψψ channel [143].

In 2024, the CMS Collaboration observed three resonant structures in the J/ψJ/ψ mass spectrum with the masses 663838+4331+16MeV, 684728+4420+48MeV and 713425+4815+41MeV, respectively [144]. While in the no-interference model, the measured Breit−Wigner masses and widths are [144]

X(6600):M=6552±10±12MeV,Γ=12426+32±33MeV,X(6900):M=6927±9±4MeV,Γ=12221+24±18MeV,X(7300):M=728718+20±5MeV,Γ=9540+59±19MeV.

The two-meson pairs J/ψJ/ψ, J/ψψ, ψψ, hchc, χc0χc0, χc1χc1 and χc2χc2 lie at 6194MeV, 6783MeV, 7372MeV, 7051MeV, 6829MeV, 7021MeV and 7112MeV, respectively [97], it is difficult to assign the X(6600), X(6900) and X(7300) as the charmonium−charmonium molecular states without introducing coupled channel effects.

We can assign the X(6600), X(6900) and X(7300) as the color 3¯3-type tetraquark states tentatively based on the QCD sum rules, see Tab.41 in Section 3.3.

1.9 Y(4260/4230), Y(4360/4320), Y(4390), Y(4500), Y(4660), Y(4710/4750/4790)

In 2005, the BaBar Collaboration studied the initial-state radiation process e+eγISRπ+πJ/ψ and observed the Y(4260) in the π+πJ/ψ mass spectrum, the measured mass and width are (4259±86+2)MeV and (88±234+6)MeV, respectively [145]. Subsequently the Y(4260) was confirmed by the Belle and CLEO Collaborations [146, 147], the Belle Collaboration also observed an evidence for a very broad structure Y(4008) in the π+πJ/ψ mass spectrum.

In 2007, the Belle Collaboration studied the initial-state radiation process e+eγISRπ+πψ, and observed the Y(4360) and Y(4660) in the π+πψ mass spectrum at 4361±9±9MeV with a width of (74±15±10)MeV and (4664±11±5)MeV with a width of (48±15±3)MeV, respectively [148, 149], then the Y(4660) was confirmed by the BaBar Collaboration [150].

In 2008, the Belle Collaboration studied the initial-state radiation process e+eγISRΛc+Λc, observed a clear peak Y(4630) in the Λc+Λc mass spectrum just above the Λc+Λc threshold, and determined the mass and width to be (46347+88+5)MeV and (9224+4021+10)MeV, respectively [151]. Thereafter, the Y(4660) and Y(4630) are taken as the same particle according to the uncertainties of the masses and widths, for example, in The Review of Particle Physics [97].

In 2014, the BESIII Collaboration searched for the production of e+eωχcJ with J=0,1,2, and observed a resonance in the ωχc0 cross section, the measured mass and width are 4230±8±6MeV and 38±12±2MeV, respectively [152].

In 2016, the BESIII Collaboration measured the cross sections of the process e+eπ+πhc, and observed two structures, the Y(4220) has a mass of 4218.4±4.0±0.9MeV and a width of 66.0±9.0±0.4MeV respectively, and the Y(4390) has a mass of 4391.6±6.3±1.0MeV and a width of 139.5±16.1±0.6MeV respectively [153].

Also in 2016, the BESIII Collaboration precisely measured the cross section of the process e+eπ+πJ/ψ and observed two resonant structures, which agree with the Y(4260) and Y(4360), respectively. The first resonance has a mass of 4222.0±3.1±1.4MeV and a width of 44.1±4.3±2.0MeV, while the second one has a mass of 4320.0±10.4±7.0MeV and a width of 101.419.7+25.3±10.2MeV [154].

In 2022, the BESIII Collaboration observed two resonant structures in the K+KJ/ψ mass spectrum, one is the Y(4230) and the other is the Y(4500), which was observed for the first time with the Breit−Wigner mass and width 4484.7±13.3±24.1MeV and 111.1±30.1±15.2MeV, respectively [155].

In 2023, the BESIII Collaboration observed three enhancements in the DD0π+ mass spectrum in the Born cross sections of the process e+eD0Dπ+, the first and third resonances are the Y(4230) and Y(4660), respectively, while the second resonance has the Breit-Wigner mass and width 4469.1±26.2±3.6MeV and 246.3±36.7±9.4MeV, respectively, and is roughly compatible with the Y(4500) [156].

Also in 2023, the BESIII Collaboration observed three resonance structures in the Ds+Ds mass spectrum, the two significant structures are consistent with the ψ(4160) and ψ(4415), respectively, while the third structure is new, and has the Breit−Wigner mass and width 4793.3±7.5MeV and 27.1±7.0MeV, respectively, therefore is named as Y(4790) [157].

Also in 2023, the BESIII Collaboration observed a new resonance Y(4710) in the K+KJ/ψ mass spectrum with a significance over 5σ, the measured Breit−Wigner mass and width are 470815+17±21MeV and 12623+27±30MeV, respectively [158].

In 2024, the BESIII Collaboration measured the Born cross sections for the processes e+eωχc1 and ωχc2, and observed the well established ψ(4415) in the ωχc2 mass spectrum [159]. In addition, they observed a new resonance in the ωχc1 mass spectrum, and measured the mass and width as 4544.2±18.7±1.7MeV and 116.1±33.5±1.7MeV, respectively, which are also roughly compatible with the Y(4500).

Also in 2024, the BESIII Collaboration studied the processes e+eωX(3872) and γX(3872), and observed that the relatively large cross section for the e+eωX(3872) process is mainly due to the enhancement about 4.75 GeV, which maybe indicate a potential structure in the e+eωX(3872) cross section [160]. If the enhancement is confirmed in the future by enough experimental data, there maybe exist another Y state, the Y(4750).

We should bear in mind, in 2023, the BESIII Collaboration studied the process e+eΛc+Λc at twelve center-of-mass energies from 4.6119 to 4.9509 GeV, determined the Born cross sections and effective form-factors with unprecedented precision, and obtained flat cross sections about 4.63 GeV, which does not indicate the resonant structure Y(4630) [161].

The charmonium-like candidates Y(4260/4230), Y(4360/4320), Y(4390), Y(4500), Y(4660) and Y(4710/4750/4790) with the JPC=1 overwhelm the accommodating capacity of the traditional cc¯ model, some of them should be multiquark states.

Based on the predictions of the QCD sum rules, we can assign the Y(4260/4230), Y(4360/4320), Y(4390) and Y(4750) as the color 3¯3-type tetraquark states with an explicit P-wave between the diquark and antidiquark tentatively, see Tab.33 in Section 3.1.4, and assign the Y(4360), Y(4390), Y(4500), Y(4660), Y(4710) and Y(4790) as the color 3¯3-type tetraquark states with an implicit P-wave in the diquark or antidiquark tentatively, see Tab.22 and Tab.23 in Section 3.1.3.

1.10 Zc(3900/3885), Zc(4020/4025), Zcs(3985/4000), Zcs(4123), Zcs(4220)

In 2013, the BESIII Collaboration studied the process e+eπ+πJ/ψ at a center-of-mass energy of 4.260GeV, and observed a structure Zc±(3900) in the π±J/ψ mass spectrum with a mass of (3899.0±3.6±4.9)MeV and a width of (46±10±20)MeV [162], at the same time, the Belle Collaboration studied the process e+eγISRπ+πJ/ψ using initial-state radiation, and observed a structure Zc±(3900) in the π±J/ψ mass spectrum with a mass of (3894.5±6.6±4.5)MeV and a width of (63±24±26)MeV [163]. Then this structure was confirmed by the CLEO Collaboration [164].

Also in 2013, the BESIII Collaboration studied the process e+e(DD¯)±π at s=4.26GeV, and observed a structure Zc±(4025) near the (DD¯)± threshold in the π recoil mass spectrum [165]. The measured mass and width are (4026.3±2.6±3.7)MeV and (24.8±5.6±7.7)MeV, respectively [165]. Slightly later, the BESIII Collaboration studied the process e+eπ+πhc at s from 3.90GeV to 4.42GeV, and observed a distinct structure Zc(4020) in the π±hc mass spectrum, the measured mass and width are (4022.9±0.8±2.7)MeV and (7.9±2.7±2.6)MeV, respectively [166].

In 2014, the BESIII Collaboration studied the process e+eπDD¯ at s=4.26GeV, and observed a distinct charged structure Zc(3885) in the (DD¯)± mass spectrum [167]. The measured mass and width are (3883.9±1.5±4.2)MeV and (24.8±3.3±11.0)MeV, respectively, and the angular distribution of the πZc(3885) system favors the assignment JP=1+ [167].

We tentatively identify the Zc(3900) and Zc(3885) as the same particle according to the uncertainties of the masses and widths [60]. In 2017, the BESIII Collaboration established the spin-parity of the Zc(3900) to be JP=1+ [168].

In 2021, the BESIII Collaboration observed an excess near the DsD0 and DsD0 thresholds in the K+ recoil-mass spectrum with the significance of 5.3 σ in the processes e+eK+(DsD0+DsD0) [169]. The Breit-Wigner mass and width of the new structure Zcs(3985) were measured as 3985.22.0+2.1±1.7MeV and 13.85.2+8.1±4.9MeV, respectively.

The Zc(3885) and Zcs(3985) have similar production modes,

e+eZc(3885)π+(DD¯)π+,e+eZcs(3985)K+(DsD0+DsD0)K+,

and they should be cousins and have similar properties.

Also in 2021, the LHCb Collaboration reported two new exotic states with the valence quarks cc¯us¯ in the J/ψK+ mass spectrum in the decays B+J/ψϕK+ [115]. The most significant state Zcs+(4000) has a mass of 4003±614+4MeV, a width of 131±15±26MeV, and the spin-parity JP=1+, while the broader state Zcs+(4220) has a mass of 4216±2430+43MeV, a width of 233±5273+97MeV, and the spin-parity JP=1+ or 1 (with a 2σ difference in favor of the first hypothesis) [115]. Considering the large difference between the widths, the Zcs(3985) and Zcs(4000) are unlikely to be the same particle.

In 2023, the BESIII Collaboration reported an excess of the Zcs(4123)DsD0 candidate at a mass of (4123.5±0.7±4.7)MeV with a significance of 2.1σ in the process e+eK+DsD0+c.c. [170]. The Zcs(4123) is consistent with the tetraquark state with the valence quarks cc¯su¯, spin-parity-charge-conjugation JPC=1+, a mass 4.11±0.08GeV and a width 22.71±1.65MeV predicted in previous work based on the QCD sum rules [171].

The charmonium-like states Zc(3900/3885), Zc(4020/4025), Zcs(3985/4000), Zcs(4123) and Zcs(4220) have non-zero electric charge, and are excellent candidates for the tetraquark (molecular) states [171, 172].

Based on the predictions of the QCD sum rules, we can assign the Zc(3900/3885), Zc(4020), Zcs(3985/4000) and Zcs(4123) as the color 3¯3-type tetraquark states, see Tab.9 and Tab.11 in Section 3.1.1, or 11-type tetraquark states, see Tab.44 in Section 4.1.

1.11 Z1(4050), Z1(4250), Zc(4100)

In 2008, the Belle Collaboration reported the first observation of two resonance-like structures Z1+(4050) and Z2+(4250) exceeding 5σ in the π+χc1 mass spectrum near 4.1GeV in the exclusive decays B¯0Kπ+χc1 [173]. The Breit−Wigner masses and widths are M1=4051±1441+20MeV, Γ1=8217+2122+47MeV, M2=424829+4435+180MeV and Γ2=17739+5461+316MeV, respectively. However, the BaBar Collaboration observed no evidence for the Z1+(4050) and Z2+(4250) states in the π+χc1 mass spectrum in the exclusive decays B¯0π+χc1K and B+π+χc1KS0 [174].

In 2018, the LHCb Collaboration observed an evidence for the ηcπ resonant structure Zc(4100) with the significance larger than 3σ in a Dalitz plot analysis of the B0ηcK+π decays, the measured mass and width are 4096±2022+18MeV and 152±5835+60MeV respectively [175]. The assignments JP=0+ and 1 are both consistent with the experimental data. However, the Zc(4100) is not confirmed by other experiments until now.

1.12 Zc(4430), Zc(4600)

In 2007, the Belle Collaboration observed a distinct peak in the π±ψ mass spectrum in the decays BKπ±ψ, the mass and width are (4433±4±2)MeV and (4513+1813+30)MeV, respectively [176]. In 2009, the Belle Collaboration observed a signal for the decay Z(4430)+π+ψ from a Dalitz plot analysis of the decays BKπ+ψ [177]. In 2013, the Belle Collaboration performed a full amplitude analysis of the decays B0ψK+π to reach the favored assignments JP=1+ [178].

In 2014, the LHCb Collaboration analyzed the B0ψπK+ decays by performing a four-dimensional fit of the amplitude, and provided the first independent confirmation of the Z(4430) resonance and established its spin-parity JP=1+. The measured Breit−Wigner mass and width are (4475±725+15)MeV and (172±1334+37)MeV, respectively [179], which excludes the possibility of assigning the Zc(4430) as the DD1 molecular state with the spin-parity JP=0 [180], although it lies near the DD1 threshold.

The Okubo−Zweig−Iizuka supper-allowed decays

Zc(3900)J/ψπ,Zc(4430)ψπ

are expected to take place easily, and the energy gaps have the relation MZcMZc=mψmJ/ψ, the Zc(4430) can be assigned as the first radial excitation of the Zc(3900) [56, 181, 182], which was proposed before the JP of the Zc(3900) were determined by the BESIII Collaboration [168].

In 2019, the LHCb Collaboration performed an angular analysis of the weak decays B0J/ψK+π, examined the m(J/ψπ) versus the m(K+π) plane, and observed two possible resonant structures in the vicinity of the energies m(J/ψπ)=4200MeV and 4600MeV, respectively [183], the structure Zc(4600) has not been confirmed by other experiments yet. According to the mass gaps MZc(4600)MZc(4020)MZc(4430)MZc(3900), we can tentatively assign the Zc(4600) as the first radial excitation of the Zc(4020) [184, 185].

Based on the predictions of the QCD sum rules, we can assign the Zc(4430) and Zc(4600) as the first radial excitations of the color 3¯3-type tetraquark states, see Tab.9 in Section 3.1.1 and Tab.17 in Section 3.1.2.

1.13 Zc(4200), Zc¯s¯(4600), Zc¯s¯(4900), Zc¯s¯(5200)

In 2014, the Belle Collaboration analyzed the decays B¯0Kπ+J/ψ and observed a resonance Zc(4200) in the J/ψπ+ mass spectrum with a statistical significance more than 6.2σ, the measured mass and width are 419629+3113+17MeV and 37070+70132+70MeV, respectively, the preferred assignment is JP=1+ [186].

In 2019, the LHCb Collaboration performed an angular analysis of the decays B0J/ψK+π, examined the m(J/ψπ) versus the m(K+π) plane, and observed two structures in the vicinity of the energies m(J/ψπ)=4200MeV and 4600MeV, respectively [183].

In 2024, the LHCb Collaboration performed the first full amplitude analysis of the decays B+ψ(2S)K+π+π, and they developed an amplitude model with 53 components comprising 11 hidden-charm exotic states, for example, the Zc(4200) and Zc(4430) in the ψ(2S)π+ mass spectrum with the JP=1+; the Zc¯s¯(4600) and Zc¯s¯(4900) in the ψ(2S)K(892) mass spectrum with the JP=1+, which might be the radial excitations of the Zc¯s¯(4000) in the scenario of tetraquark states with the valence quarks cc¯ds¯; the Zc¯s¯(4000), Zc(4055) and Zc¯s¯(5200) are effective descriptions of generic partial wave-functions with the JP=1+, 1 and 1, respectively [116]. The spin-parity of the Zc(4200) is determined to be 1+ for the first time with a significance exceeding 5σ.

We group the Zc(4200) with the Zc¯s¯(4600), Zc¯s¯(4900) and Zc¯s¯(5200) together into one-subsection as its assignment is still an open problem, and we would like to revisit this subject to discuss the possible assignment based on the QCD sum rules in Section 3.1.

1.14 Zb(10610), Zb(10650), Y(10750)

In 2011, the Belle Collaboration reported the first observation of the Zb(10610) and Zb(10650) in the π±Υ(1,2,3S) and π±hb(1,2P) mass spectra associated with a single charged pion in the Υ(5S) decays, the quantum numbers IG(JP)=1+(1+) are favored [187]. Subsequently, the Belle Collaboration updated the measured parameters MZb(10610)=(10607.2±2.0)MeV, MZb(10650)=(10652.2±1.5)MeV, ΓZb(10610)=(18.4±2.4)MeV and ΓZb(10650)=(11.5±2.2)MeV, respectively [188]. In 2013, the Belle Collaboration observed the Υ(5S)Υ(1,2,3S)π0π0 decays for the first time, and obtained the neutral Zb0(10610) in a Dalitz analysis of the decays to the final states Υ(2,3S)π0 [189].

In 2019, the Belle Collaboration observed a resonance structure Y(10750) in the e+eΥ(nS)π+π (n=1,2,3) cross sections [190]. The Breit−Wigner mass and width are 10752.7±5.91.1+0.7MeV and 35.511.3+17.63.3+3.9MeV, respectively. The Y(10750) is observed in the Υ(nS)π+π mass spectrum with n=1,2,3, its quantum numbers are JPC=1.

The BelleII Collaboration confirmed the Y(10750) in the processes e+eωχb1(1P), ωχb2(1P) [191], π+πΥ(1S), π+πΥ(2S) [192], and observed no evidence in the processes e+eωηb(1S), ωχb0(1P) [193], π+πΥ(3S) [192].

Based on the predictions of the QCD sum rules, we can assign the Zb(10610) and Zb(10650) as the color 3¯3-type or 11-type tetraquark states tentatively, see Section 3.1.1 and Section 4.1, and assign the Y(10750) as the color 3¯3-type tetraquark state with an explicit P-wave between the diquark and antidiquark tentatively, see Section 3.1.4.

1.15 Tcc(3875)

In 2021, the LHCb Collaboration formally announced observation of the exotic state Tcc+(3875) just below the D0D+ threshold [194, 195]. The Breit−Wigner mass and width are δMBW=273±61±514+11KeV below the D0D+ threshold and ΓBW=410±165±4338+18KeV [194, 195]. The exotic state Tcc+(3875) is consistent with the ground isoscalar tetraquark state with the valence quarks ccu¯d¯ and spin-parity JP=1+, and exploring the DD mass spectrum disfavors interpreting the Tcc+(3875) as an isovector state. The observation of the Tcc+(3875) is a great breakthrough beyond the Ξcc++ for hadron physics, and it is the first doubly-charmed tetraquark candidate with the typical quark configuration ccu¯d¯.

Based on the predictions of the QCD sum rules, we can assign the Tcc(3875) as the color 3¯3-type or 11-type tetraquark state tentatively, see Section 3.2 and Section 4.2.

1.16 Pc(4312), Pc(4380), Pc(4440), Pc(4457), Pc(4337), Pcs(4338), Pcs(4459)

In 2015, the LHCb Collaboration observed two exotic structures Pc(4380) and Pc(4450) in the J/ψp mass spectrum in the Λb0J/ψKp decays [196]. The Pc(4380) has a mass of 4380±8±29MeV and a width of 205±18±86MeV, while the Pc(4450) has a mass of 4449.8±1.7±2.5MeV and a width of 39±5±19MeV. The preferred spin-parity assignments of the Pc(4380) and Pc(4450) are JP=32 and 52+, respectively [196].

In 2019, the LHCb Collaboration studied the Λb0J/ψKp decays with a data sample, which is an order of magnitude larger than that previously analyzed, and observed a narrow pentaquark candidate Pc(4312) in the J/ψp mass spectrum. Furthermore, the LHCb Collaboration confirmed the pentaquark structure Pc(4450), and observed that it consists of two narrow overlapping peaks Pc(4440) and Pc(4457) [197]. The measured masses and widths are

Pc(4312):M=4311.9±0.70.6+6.8MeV,Γ=9.8±2.74.5+3.7MeV,Pc(4440):M=4440.3±1.34.7+4.1MeV,Γ=20.6±4.910.1+8.7MeV,Pc(4457):M=4457.3±0.61.7+4.1MeV,Γ=6.4±2.01.9+5.7MeV.

In 2021, the LHCb Collaboration reported an evidence of a hidden-charm pentaquark candidate Pcs(4459) with the strangeness S=1 in the J/ψΛ mass spectrum with a significance of 3.1σ in the ΞbJ/ψKΛ decays [198], the Breit−Wigner mass and width are

Pcs(4459):M=4458.8±2.91.1+4.7MeV,Γ=17.3±6.55.7+8.0MeV,

and the spin-parity have not been determined yet up to now.

In 2022, the LHCb Collaboration observed an evidence for a structure Pc(4337) in the J/ψp and J/ψp¯ systems in the Bs0J/ψpp¯ decays with a significance about 3.13.7σ depending on the JP hypothesis [199], the Breit−Wigner mass and width are

Pc(4337):M=43374+72+2MeV,Γ=2912+2614+14MeV.

Its existence still needs confirmation and its spin-parity is not measured yet.

In 2023, the LHCb Collaboration observed an evidence for a new structure Pcs(4338) in the J/ψΛ mass distribution in the BJ/ψΛp¯ decays [200], the measured Breit−Wigner mass and width are

Pcs(4338):M=4338.2±0.7±0.4MeV,Γ=7.0±1.2±1.3MeV,

and the favored spin-parity is JP=12.

The Pcs(4338) and Pcs(4459) are observed in the J/ψΛ mass spectrum, they have the isospin I=0, as the strong decays conserve isospin. The Pc(4312), Pc(4380), Pc(4440), Pc(4457), Pcs(4459) and Pcs(4338) lie slightly below or above the thresholds of the charmed meson-baryon pairs D¯Σc, D¯Σc, D¯Σc, D¯Σc, D¯Ξc (D¯Ξc, D¯Ξc, D¯Ξc) and D¯Ξc, respectively. It is difficult to identify the Pc(4337) as the molecular state without resorting to the help of large coupled-channel effects due to lacking nearby meson-baryon thresholds. Or the Pc(4312) and Pc(4337) are the same particle, such a possibility cannot be excluded at the present time.

Based on the predictions of the QCD sum rules, we can assign the Pc(4312), Pc(4337), Pc(4380), Pc(4440), Pc(4457) and Pcs(4459) as the color 3¯3¯3¯-type pentaquark states tentatively, see Tab.50 and Tab.52 in Section 5.1, and assign the Pc(4312), Pc(4380), Pc(4440), Pc(4457), Pcs(4338) and Pcs(4459) as the color 11-type pentaquark states, see Tab.54 in Section 5.2.

1.17 d(2380)

In 2014, the scientists in the WASA-at-COSY Collaboration and SAID data analysis center performed exclusive and kinematically complete high-statistics measurements of the polarized np scattering through the quasifree process dpnp+pspectator in the energy region of the narrow resonance-like structure d with the I(JP)=0(3+), and confirmed their (WASA-at-COSY Collaboration) early observation of the d(2380) in the double-pionic fusion channels, they produced a resonance pole in the 3D33G3 coupled partial waves at 2380±10i40±5 MeV [201, 202], — in accordance with the ΔΔ dibaryon resonance [203-207]. And we will revisit this subject at the end of Section 4.2.

2 Theoretical foundations

In this section, we would like to review the typical theoretical methods and related possible assignments concisely, then focus on the QCD sum rules in the subsequent sub-sections, see Sections 2.2, 2.3, and 2.4.

2.1 Typical theoretical methods and possible assignments

There have been tremendous progresses on the hadron spectrum containing two heavy quarks experimentally since the observation of the X(3872). It is surprising that many resonant structures lie around thresholds of a pair of heavy hadrons. A natural conjecture is that they are possible deuteron-like two-particle bound states bound via attractive interactions induced by one-pion exchange or one-boson exchange [63, 65, 67, 72, 76, 208-215], it is only a possibility. In the heavy quark limit, the Q3q¯3¯ mesons and Q3[qq]3¯ baryons have the antiquark-diquark symmetry, q¯3¯[qq]3¯, therefore the Qq¯qQ¯ and Q[qq]qQ¯ systems could be analyzed in the same theoretical scheme, in this sub-section, we would like to focus on the tetraquark systems. Someone maybe wonder: are they threshold cusps, triangle singularities or genuine resonances? As there always exist threshold cusps at the S-wave thresholds or triangle singularities near the thresholds. Firstly, let us see the outcomes based on the (non)relativistic effective field theory.

2.1.1 Threshold cusps, triangle singularities or genuine resonances

Not all peaks in the invariant mass distributions are genuine resonances, they often arise due to the nearby kinematical singularities of the transition amplitudes in the complex energy plane. Those singularities (or Landau singularities) occur when the intermediate particles are on the mass-shell. The simplest case is the cusp at the normal two-body threshold, there always exists a cusp at the S-wave threshold of two particles coupling to the final states, while a more complicated case is the so-called triangle singularity. They maybe produce observable effects if the involved interactions are strong enough, sometimes, even mimic the behavior of a resonance. It is important to distinguish kinematic singularities from genuine resonances. We would like to give an example concerning the exotic states Y(4260) and Zc(3900) to illustrate their possible assignments in the scenarios of threshold cusps, triangle singularities and genuine resonances.

The threshold cusp is determined by masses of the involved particles, how strong the cusp depends on detailed dynamics and the cusp could be rather dramatic if there is a nearby pole, thus it plays an important role in studying the exotic states [216]. For example, the X(3872), Zc(3900), Zc(4020), Zb(10610) and Zb(10650) lie near the D0D¯0, DD¯, DD¯, BB¯ and BB¯ thresholds, respectively, their quantum numbers are the same as the corresponding S-wave meson pairs although the JPC of the Zc(4020) have not been fully determined yet [97].

The X(3872) was assigned to be a threshold cusp by Bugg [89], subsequently, he realized that the very narrow line shapes in the J/ψρ and D0D¯0 channels could not be fitted with only a threshold cusp, and a resonance or virtual state pole was necessary [217].

In a modified threshold cusp model [218, 219], see the Feynman diagram shown in Fig.1 as an example, both the inelastic (J/ψπ,hcπ) and elastic (DD¯, DD¯) decay modes were considered for the Zc(3900) and Zc(4020), analogous discussions are applied to the Zb(10610) and Zb(10650). A Gaussian form-factor was chosen for all the vertices including the tree-level ones. Then the experimental data for the J/ψπ and DD¯ mass spectra for the Zc(3900) and the DD¯ and hcπ mass spectra for the Zc(4020) could be fitted very good, exotic resonances are not required to account for the experimental data. However, the fitting quality depends crucially on the cutoff parameter in the Gaussian form-factor.

The triangle singularity is determined by the masses of the intermediate particles plus the invariant masses of the external ones, therefore the triangle singularities are sensitive to the kinematic variables, the peak position and peak shape change according to the variations of the external energies. More precisely, the triangle singularities are determined by the scalar triangle loop integral, which does not depend on the orbital angular momentum for each vertex, however, sharp triangle singularity peaks are constrained to the S-wave internal particles, as momentum power factor weakens the singular behavior in other cases [216].

In the initial single-pion emission (ISPE) mechanism, the triangle diagrams contribute to threshold cusps, see Fig.2 for a typical Feynman diagram. This mechanism was suggested firstly to study the exotic structures Zb(10610) and Zb(10650), Chen and Liu introduced a dipole form-factor to accompany the exchanged B-meson propagator and took account of the BB¯, BB¯, BB¯ and BB¯ triangle loop diagrams, and produced sharp cusps right around the Zb(10610) and Zb(10650) structures in the Υ(1S,2S,3S)π and hb(1P,2P)π mass spectra, but observed no cusp at the BB¯ threshold [220]. Similarly, Chen, Liu and Matsuki [221] took account of the DD¯, DD¯, DD¯ and DD¯ triangle loop diagrams and the intermediate f0(600) and f0(980) to study the decays Y(4260)J/ψππ, and observed two peaks, the Zc(3900) and its reflection. And they studied other processes with possible triangle singularities [222].

In Ref. [223], Dong, Guo and Zou showed that the threshold cusp appears as a peak only for channels with attractive interaction, and the cusp’s width is inversely proportional to the reduced mass for the relevant threshold. There should be threshold structures at any threshold of a qQ¯ and Qq¯ (Qqq) pair, which have attractive interaction at threshold, in the invariant mass distribution of a QQ¯ state and a qq¯ (qqq) state coupling to the qQ¯ and Qq¯ (Qqq) pair, and the structure becomes more pronounced if there is a near-threshold pole.

In Ref. [224], Liu and Li supposed the Y(4260) as a D1D¯ molecular state to study its decays, see Fig.3 as an example, and observe that under special kinematic configurations, the triangle singularity maybe occur in the re-scattering amplitude, which can change the threshold behavior significantly. Obvious threshold enhancements or narrow cusp structures appear quite naturally without introducing a genuine resonance, but cannot exclude existence of a genuine resonance, such a mechanism also works for the pentaquark structures [225, 226].

If the Y(4260) have a large D1D¯ molecular component, the decays Y(4260)J/ψπ+π can occur through the re-scattering process [227-229], see Fig.3. The singularity regions provide an ideal environment for forming bound states or resonances. Although the Y(4260) lies slightly below the D1(2420)D¯ threshold, the triangle singularity in the J/ψπ± invariant mass distribution of the re-scattering amplitude is still near the physical boundary and can influence the J/ψπ± invariant mass distribution around the DD¯ threshold significantly [227]. Despite the importance of the triangle diagram contribution, it is insisted that a Zc(3900) resonance was still needed in order to fit to the narrow peak observed in experiments [224, 227, 230, 231]. The diagrams similar to Fig.3 also play an important role in the hidden-bottom sector [232, 233].

The resonance pole can be incorporated by constructing a unitarized coupled-channel scattering T-matrix by fitting to the experimental data, the best fit still demands the T-matrix to have a resonant or virtual pole near the DD¯ threshold, which can be interpreted as the Zc(3900) [234]. The molecule assignment provides a natural explanation for the resonance-like structure Zc(3900) in the Y(4260) decays [229, 234], the kinematical threshold cusp cannot produce a narrow peak in the invariant mass distribution in the elastic channel in contrast with a genuine S-matrix pole [235].

In a similar scenario, Szczepaniak suggested that the Zc(3900) peak could be attributed to the D0(2300)D¯D loop instead of the D1(2420)D¯D loop, which is in the physical region by neglecting the width of the D0(2300) [236]. The triangle singularities can produce enhancement potentially in the amplitude consistent with the experimental data qualitatively.

In Ref. [237], Chen, Du and Guo performed a unified description of the π+π and J/ψπ± mass distributions for the e+eJ/ψπ+π and the D0D mass distribution for the e+eD0Dπ+ at s= 4.23 and 4.26 GeV. They take account of the open-charm meson loops containing triangle singularities, the J/ψπ-DD¯ coupled-channel interaction respecting unitarity, and the strong ππ-KK¯ final-state interaction using dispersion relations, which lead to a precise determination of the pole mass and width (3880.7±1.7±22.4) MeV and (35.9±1.4±15.3) MeV, respectively, and indicate the molecular and non-molecular components are of similar importance for the structure Zc(3900).

Precisely measuring the near threshold structures plays an important role in diagnosing the heavy-hadron interactions, therefore understanding the puzzling hidden-charm and hidden-bottom structures. Furthermore, it is important to search for the resonant structures in processes free of triangle singularities, such as the photo-production and pion-induced production processes in the e+e and pp collisions [238-240]. For a recent review on the production of the exotic hadrons in the pp and nuclear collisions, see Ref. [241].

2.1.2 Dynamical generated resonances and molecular states

If we take the traditional heavy mesons as the elementary degrees of freedom, then we construct the heavy meson effective Lagrangian according to the chiral symmetry, hidden-local symmetry and heavy quark symmetry [242-244]. It is easy to obtain the two-meson scattering amplitudes V. Then we have three choices:

Firstly, we unitarize the amplitudes by taking account of the intermediate two-meson loops with the coupled channel effects through the Bethe-Salpeter or Lippmann-Schwinger equation with on-shell factorization [245, 246],

T=V+VGV+VGVGV+=[1VG]1V,

where the G is the loop function, see Fig.4 for a diagrammatical representation. Then we explore the analytical properties of the full amplitudes T, and try to find the poles in the complex Riemann sheets, such as the bound states, virtual states and resonances. Such discussions are applied to the baryon-meson systems directly.

Bound states appear as poles on the physical sheet, and only appear on the real s-axis below the lowest threshold by causality. Virtual states also appear on the real s-axis, however, on the unphysical Riemann sheet. Resonances appear as poles on an unphysical Riemann sheet close to the physical one with non-zero imaginary part, and they appear in conjugate pairs. For example, the loosely bound states X(3872), Y(3940), Zc(3900), Zb(10610), Zb(10650) [70, 75, 247-254], the hidden-charm pentaquark resonances [255-259]. We usually apply Weinberg’s compositeness condition to estimate the hadronic molecule components [260-266].

Secondly, we take the scattering amplitudes V as interaction kernels, solve the quasi-potential Bethe-Salpeter or Lippmann−Schwinger equation with the coupled channel effects directly, then explore the analytical properties of the full amplitudes [267-273], or obtain the bound energies directly to estimate the bound states [76, 274-278].

Thirdly, we reduce the scattering amplitudes to interaction potentials in the momentum space in terms of the Breit approximation and introduce monopole form-factors associated with the exchanged particles. Generally, we should introduce form-factors in each interaction vertex, which reflects the off-shell effect of the exchanged meson and the structure effect, because the components of the molecular states and exchanged mesons are not point particles. Then we perform the Fourier transformation to obtain the potential in the coordinate space, finally we solve the Schrödinger equation directly to obtain the binding energy [63, 279-289].

2.1.3 QQ¯ states with coupled channel effects

In the famous GI model, the charmonium χc1(23P1) has the mass about 3953MeV [86], which is about 100MeV above the X(3872) lying near the D0D¯0 threshold, the strong coupling to the nearby threshold maybe lead to some DD molecular configuration. Furthermore, pure charmonium assignment cannot interpret the high γψ decay rate [290].

We can extend the constituent quark models to include the meson−meson Fock components, and write the physical charmonium (bottomonium) states in terms of |Ψ,

|Ψ=αcα|ψα+βχβ(P)|ϕ1ϕ2β,

where the |ψα are the QQ¯ eigenstates, the ϕi are Qq¯ or qQ¯ eigenstates, |ϕ1ϕ2β are the two-meson state with β quantum numbers, and the χβ(P) is the relative wave function between the two mesons. Then we solve the Schrodinger equation directly [291-300].

2.1.4 Hybrid and tetraquark states in Born−Oppenheimer approximation

Due to the large ratio of the mass of a nucleus to that of the electron, the electrons respond almost instantaneously to the motion of the nuclei. The energy of the electrons combined with the repulsive Coulomb energy of the nuclei defines a Born−Oppenheimer potential. Accordingly, due to the large ratio of the heavy-quark mass mQ to the energy scale ΛQCD associated with the gluon field, the gluons respond almost instantaneously to the motion of the heavy quarks Q and Q¯ [301-306].

In the static limit, the Q and Q¯ serve as two color-sources separated by a distance r, the ground-state flavor-singlet Born−Oppenheimer potential VΣg+(r) is defined by the minimal energy of the gluonic configurations, whose small and large r limiting behaviors are qualitatively compatible with the simple phenomenological Cornell potential, thus the VΣg+(r) describes the traditional heavy quarkonium states. The excited Born-Oppenheimer potentials VΓ(r) are defined as the minimal energies of the excited configurations for the gluon and light-quark fields with the quantum numbers Γ [301-303].

The hybrid states Q¯GQ are energy levels of a heavy quark pair QQ¯ in the excited flavor-singlet Born-Oppenheimer potentials, the hybrid potentials. Juge, Kuti and Morningstar [307] calculated many flavor-singlet potentials using the quenched lattice QCD. There have been some works on the lowest lying hybrid potentials using lattice QCD with two flavors of dynamical Wilson fermions [308, 309]. At large r, the hybrid potential VΓ(r) is a flux-tube extending between the Q and Q¯. At small r, the hybrid potential approaches the repulsive color-Coulomb potential between a Q and Q¯ in a color-octet state. In the limit r0, the Q and Q¯ sources reduce to a single local color-octet QQ¯ source. The energy levels of the flavor-singlet gluon and light-quark field configurations bound to a static color-octet source are called static hybrid mesons. The most effective pictorial representation of the hybrid states is the flux-tube model. Lattice QCD simulations show that two static quarks Q and Q¯ at large distances are confined by approximately cylindrical regions of the color fields [307-309].

The Q¯Qq¯q tetraquark states are energy levels in the Born-Oppenheimer potentials with nonsinglet (excited singlet) flavor quantum numbers, the tetraquark potentials, which are distinguished by the quantum numbers, uu¯±dd¯, ud¯, du¯, sq¯, ss¯, etc. At large r, the minimal-energy configuration consists of two static mesons localized near the Q and Q¯ sources. At small r, the minimal-energy configuration is the flavor-singlet Σg+ potential accompanied by one or two pions (two or three pions), depending on the quantum numbers Γ. There have been works on the energies of static adjoint mesons using the quenched lattice QCD [310]. The static adjoint mesons are energy levels of the light-quark and gluon fields with nonsinglet flavor quantum numbers bound to a static color-octet source.

The heavy quark motion is restored by solving the Schrodinger equation in each of those potentials, and many X, Y and Z mesons could be assigned as the bound states with the Born−Oppenheimer potentials [301-306].

2.1.5 Tetraquarks in diquark models

If we take the quarks in color triplet 3 as the basic constituents, then we could construct the hadrons according to the SU(3) symmetry. For the traditional mesons,

q3q¯3¯[q¯q]1.

For the traditional baryons,

q3q3q3[qqq]1.

For the tetraquark molecular states,

q3q¯3¯q3q¯3¯[q¯q]1[q¯q]1[q¯q]8[q¯q]8[q¯qq¯q]1[q¯qq¯q]1,

and we usually call the color 11 type structures as the molecular states. For the tetraquark states,

q3q3q¯3¯q¯3¯[qq]3¯[q¯q¯]3[qq]6[q¯q¯]6¯[qqq¯q¯]1[qqq¯q¯]1,

and we usually call the color 3¯3 type structures as the tetraquark states. For the pentaquark molecular states,

q3q3q3q3q¯3¯[qqq]1[q¯q]1[qqqqq¯]1.

For the pentaquark states,

q3q3q3q3q¯3¯[qq]3¯[qq]3¯[q¯]3¯[qqqqq¯]1.

If we take the viewpoint of the quantum field theory, the scattering amplitude for one-gluon exchange is proportional to

(λa2)ij(λa2)kl=Nc+14Nc(δijδklδilδkj)+Nc14Nc(δijδkl+δilδkj),

where the λa is the Gell-Mann matrix, the i, j, k, m and l are color indexes, the Nc is the color number, and Nc=3 in the real world. The negative sign in front of the antisymmetric antitriplet 3¯ indicates the interaction is attractive, which favors formation of the diquarks in color antitriplet, while the positive sign in front of the symmetric sextet 6 indicates the interaction is repulsive, which disfavors formation of the diquarks in color sextet.

In this sub-sub-section, we would like to focus on the tetraquark states, as the extension to the pentaquark states is straightforward. Now we define the color factor,

C^iC^j=λia2λja2,

where the subscripts i and j denote the quarks, C^iC^j=23 and 13 for the 3¯ and 6 diquark [qq], respectively, and C^iC^j=43 and 16 for the 1 and 8 quark-antiquark q¯q, respectively. If we define C^12C^34=(C^1+C^2)(C^3+C^4), then C^12C^34=43 and 103 for the 3¯3 and 66¯ type tetraquark states. It is feasible to take both the 3¯3 and 66¯ diquark configurations to explore the tetraquark states, while the preferred or usually chosen configuration is of the 3¯3 type.

The color-spin Hamiltonian can be written as [311]

H=ijκijSiSjλia2λja2,

the color factor λia2λja2 can be absorbed into the chromomagnetic couplings κij after taking matrix elements between the 3¯3 type tetraquark states.

In 2004, Maiani et al. [55] introduced the simple spin-spin Hamiltonian,

H=2m[cq]+2(κcq)3¯(ScSq+Sc¯Sq¯)+2κqq¯(SqSq¯)+2κcq¯(ScSq¯+Sc¯Sq)+2κcc¯(ScSc¯),

to study the hidden-charm tetraquark states in the diquark model, where the mcq is the charmed diquark mass. They took the X(3872) with the JPC=1++ as the basic input and predicted a mass spectrum for the 3¯3 type hidden-charm tetraquark states with the JPC=0++, 1+ and 2++. Maiani et al. [312] assigned the JPC=1+ charged resonance in the Y(4260)π+πJ/ψ decays as the X(3882) or Z(3882) according to the BESIII and Belle data, however, there is no evidence for the lower resonance X(3754) or Z(3754). In fact, in the Type-I diquark model, see the Hamiltonian in Eq. (22) [55], the predicted masses 3754MeV and 3882MeV for the JPC=1+ states are smaller than that of the tetraquark candidates Zc(3900) and Zc(4020) observed later, respectively [162, 163, 165-167].

In 2005, Maiani et al. [313] assigned the Y(4260) to be the first orbital excitation of the [cs][c¯s¯] state by including the spin-orbit interaction, and obtained a crucial prediction that the Y(4260) should decay predominantly in the DsD¯s channel. The decay model Y(4260)DsD¯s has not been observed yet up to now. In 2009, Drenska, Faccini and Polosa [314] studied the [cs][c¯s¯] tetraquark states with the JPC=0++, 0+, 0, 1++, 1+, 1+ and 1 by computing the mass spectrum and decay modes.

In 2014, Maiani et al. [56] restricted the dominant spin−spin interactions to the ones within each diquark, and simplify the effective Hamiltonian,

H=2m[cq]+2κcq(ScSq+Sc¯Sq¯),

which could describe the hierarchy of the masses of the X(3872), Zc(3900), Zc(4020) very well in the scenario of tetraquark states, furthermore, they introduced a spin−orbit interaction to interpret the Y states,

H=M00+BcL222aLS+2κcq(ScSq+Sc¯Sq¯),

where the M00, Bc and a are parameters to be fitted experimentally. Then they sorted the tetraquark states in terms of |Sqc,Sq¯c¯;S,L;J, where the L is the angular momentum between the diquark and antidiquark, Sqc=Sq+Sc, Sq¯c¯=Sq¯+Sc¯, S=Sqc+Sq¯c¯, J=S+L, and assigned the Y(4008), Y(4260), Y(4290/4220) and Y(4630) to be the tetraquark states |0,0;0,1;1, 12(|1,0;1,1;1+|0,1;1,1;1), |1,1;0,1;1 and |1,1;2,1;1, respectively. The effective Hamiltonian, see Eq. (23), is referred to as the Type-II diquark model. Then the mass spectrum of the [cs][c¯s¯] tetraquark states was explored [315], and applied to study the LHCb’s J/ψϕ resonances [316].

In 2017, Maiani, Polosa and Riquer introduced a hypothesis that the diquarks and antidiquarks in tetraquarks are separated by a potential barrier to answer the long standing questions challenging the diquark-antidiquark model of exotic resonances [317].

In 2018, Ali et al. [318] analyzed the P-wave hidden-charm tetraquark states in the diquark model using an effective Hamiltonian incorporating the dominant spin−spin, spin−orbit and tensor interactions,

H=2mD+BD2L2+2aYLS+bY(3S1nS2nS1S2)+2κcq(SqSc+Sq¯Sc¯),

where n=rr, the S1 and S2 are the spins of the 3¯ diquark D ([qc]) and 3 antidiquark D¯ ([q¯c¯]), respectively, the mD, BY, aY and bY are parameters to be fitted experimentally. And their updated analysis indicate that it is favorable to assign the Y(4220), Y(4330), Y(4390), Y(4660) as the tetraquark states |0,0;0,1;1, 12(|1,0;1,1;1+|0,1;1,1;1), |1,1;0,1;1 and |1,1;2,1;1, respectively.

In 2021, Maiani, Polosa and Riquer [319] suggested that the Zcs(3985) and Zcs(4003) are two different particles, and there exist two SU(3)f nonets with the JPC=1++ and 1+, respectively, thus they could assign the X(3872), Zc(3900), Zcs(3985), Zcs(4003) and X(4140) consistently.

Again, let us turn to the chromomagnetic interaction model, see Eq. (21), and choose the 3¯3 plus 66¯ configurations and 11 plus 88 configurations as two independent representations (or basis) respectively to explore the mass spectrum of the exotic states and their decay channels, and have obtained many successful descriptions [320-324].

In the dynamical diquark picture, Brodsky, Hwang and Lebed [57] assumed that the DD¯ pair forms promptly at the production point, and rapidly separates due to the kinematics of the production process, as the diquark and antidiquark are colored objects, they cannot separate asymptotically far apart; they create a color flux tube or string between them. If sufficient energy is available, the string would break to create an additional qq¯ pair, and rearrange into a baryon-antibaryon pair, for example, the ΛcΛ¯c pair. The overlap of the wave-functions between the quark and antiquark is suppressed greatly, due to the large spatial separation between the diquark and antidiquark pair, therefore, the transition rate is suppressed and leads to small exotic widths [57]. The exotic mass spectrum is calculated in this picture [325-331].

In the relativized quark model, the Hamiltonian can be written as

H=i=14(pi2+mi2)1/2+i<jVijconf+i<jVijoge,

where the Vijconf is the linear confining potential, the Vijoge is the one-gluon exchange potential including a Coulomb and a hyperfine term. Then the 3¯3 type configurations or both the 3¯3 and 66¯ type configurations are taken into account to solve the Schrodinger equations to obtain the mass spectrum [332-341].

In the quasipotential approach, Ebert et al. [59, 342-344] took the 3¯3-type configurations to study the hidden-charm (hidden-bottom, charm-bottom or fully-heavy) tetraquark mass spectrum by solving the Schrödinger type equations, where an effective one-gluon exchange potential plus a linear confining potential are adopted.

In the constituent quark model, all possible quark configurations satisfying the Pauli principle are explored by solving the Schrodinger equation with the potential kernel containing the confinement plus one-gluon-exchange plus (or not plus) one-meson-exchange interactions [345-352], while in the color flux-tube model, a multi-body interacting confinement potential instead of a two-body interacting confinement potential is chosen [353, 354].

2.1.6 Tetraquark states with lattice QCD

Lattice QCD provides rather accurate and reliable calculations for the hadrons which lie well below strong-decay threshold and do not decay strongly, the physical information is commonly extracted from the discrete energy spectrum. The physical system with specified quantum numbers is created from the vacuum |0 using an operator Oj at time t=0, then this system propagates for a time t before being annihilated by an operator Oi. The spectral decomposition is performed to express the correlators Cij(t) in terms of the energies En and overlaps Zjn of the eigenstates |n,

Cij(t)=0|Oi(t)Oj(0)|0=nZinZjneEnt,

Zin0|Oi|n. The correlators Cij(t) are calculated on the lattice and their time-dependence is used to extract the En and Zni [355, 356]. The lattice QCD has been applied extensively to study the exotic states [357-364].

In the energy region near or above the strong decay thresholds, the masses of the bound states and resonances are inferred from the finite-volume scattering matrix of one-channel elastic or multiple-channel inelastic scattering. Various approaches with varying degrees of mathematical rigour have been used in the simulations [365]. The simplest example is a one-channel elastic scattering with the partial wave l, where the scattering matrix S(p) satisfying unitarity SS=1 is parameterized in terms of the phase shift δl(p),

S(p)=e2iδl(p)=1+2iT(p),T(p)=1cot(δl(p))i,

the phase shift δl(p) for the S-wave scattering is extracted using the well-established and rigorous Luscher’s relation [366-368], which applies for the elastic scattering below and above threshold. The phase shifts δ(p) provide copious information about the masses of resonances and bound states. In the vicinity of a hadronic resonance with a mass mR and a width Γ, the cross section has a Breit−Wigner-type shape with the value δ(s=mR2)=π2,

T(p)=sΓ(p)smR2+isΓ(p)=1cotδ(p)i.

Below and above threshold, the pcotδ(p) can be expanded by the effective range approximation,

pcotδ(p)=1a+12rp2,

where the a is the scattering length and the r is the effective range. On the other hand, the bound state (B) is realized when the scattering amplitude T(p) has a pole at the value pB=i|pB|,

T=1cot(δ(pB))i=,

the location of this shallow bound state can be obtained by parameterizing δ(p) near the threshold and finding the pB which satisfies cot(δ(pB))=i. Most of the exotic candidates are above several two-hadron thresholds, and have more than one decay channels, which requires determining the scattering matrix for the coupled-channel nonelastic scattering matrix elements [369, 370, 371, 372, 373].

The HALQCD Collaboration use the HALQCD approach to extract the coupled-channel scattering matrix [374, 375], the HALQCD approach is based on the lattice determination of the potential V(r) between different channels, then employs the Nambu-Bethe-Salpeter equation to extract the masses of the bound states.

2.2 Multiquark states with the QCD sum rules

The QCD sum rules were introduced by Shifman, Vainstein and Zakharov in 1979 to study the conventional mesons [376, 377], then they were extended to study the conventional baryons by Ioffe [378]. The QCD sum rules are analytic and fully relativistic, and approach the bound state problem in QCD from short distances and move to longer distances step by step by including the non-perturbative effects so as to extract information on the hadronic properties.

In the past years, the QCD sum rules have been applied widely to study the hadronic properties, such as the masses of the quarks; masses and decay constants of the light and heavy mesons and baryons; form-factors of the mesons and baryons; valence quark distributions and spin structure functions of the nucleons; structure functions of the photon, pseudoscalar, vector and axialvector mesons; hadronic matrix elements for the K0K¯0, BdB¯d, BsB¯s mixing; strong coupling constants and magnetic moments of the mesons and baryons; parameters of the effective field theories; spectroscopy and properties of the exotic states; hadrons in the nuclear matter; properties of hadronic matter at high temperature and density [46, 47, 379-388]. Especially since 2007, the QCD sum rules have been applied extensively to study the X, Y, Z, T and P states, which are the typical multiquark candidates [46, 47, 58, 60, 81]. For the early works on the exotic states, we can consult Refs. [389-400].

In this sub-section, we would like to illustrate the general procedure of the QCD sum rules for the masses of the conventional hadrons and multiquark states concisely.

At the beginning point, let us write down the general two-point vacuum correlation functions,

Π(p)=id4xeipx0|T{J(x)J(0)}|0,

where the J(x) are the local currents consist of quark-gluon fields with specified quantum numbers, and the T denotes the time-ordering operation. For the conventional mesons and baryons, the currents J(x) have been explored extensively [381], for the multiquark states, the currents J(x) can be constructed straightforwardly.

At the large squared momentum region P2=p2ΛQCD2, the integral in Eq. (32) is dominated by small spatial distances and time intervals,

t|x|1P2r,

to avoid fast exponential oscillating, where the hadron size r1ΛQCD. If we set ΛQCD=(200300)MeV, then the hadron size r0.71.0fm. Therefore, at the condition of large hadron size, say r1fm, the local currents J(x) are questionable to interpolate the corresponding hadrons.

Now let us take it for granted that the exotic states have the size r1fm, just like the conventional mesons and baryons, could be interpolated by the local currents J(x) tacitly. For example, the charge radii of the π±, K± and p are r2=0.659±0.004fm, 0.560±0.031fm and 0.8409±0.0004fm respectively from the Particle Data Group [97]. We extend the QCD sum rules on the conventional mesons and baryons to study the multiquark states directly with a simple replacement of the interpolating currents, and would like to come back to this subject again in Section 2.3.

A Lorentz invariant vacuum average can be expressed as

0|T{J(x)J(0)}|0=dτexp(iτx2)f(τ),

where the f(τ) is a function. Then

Π(p2)=idτd4xexp(iτx2)exp(iP2/4τ)f(τ).

The dominant contributions to the Π(p2) come from the region,

x21τ1P2.

In the limit P2, we reach the light-cone x20, which is a necessary but not yet sufficient condition for the short-distance dominance, we have to constrain |x|1P2.

Now we focus on the quark-gluon degrees of freedom and calculate the correlation functions using Wilson’s operator product expansion to separate the physics of short and long distances,

Π(p2)=nCn(p2,μ)On(μ),

where the Cn(p2,μ) are the Wilson’s coefficients encoding short-distance contributions, the On(μ) are vacuum expectations of the local operators with dimension n. The short-distance contributions at p2>μ2 are encoded in the coefficients Cn(p2,μ), the long-distance contributions at p2<μ2 are absorbed into the vacuum condensates On(μ) [385]. If μΛQCD, the Wilson coefficients Cn(p2,μ) depend only on short-distance dynamics, the vacuum condensates On(μ) embody the long-distance effects. The lowest condensate is vacuum expectation of the unit operator O0(μ) associated with the perturbative contributions. The vacuum condensates with dimensions n=3, 4, 5, 6, are quark condensate q¯q, gluon condensate αsGGπ, mixed condensate q¯gsσGq, four-quark condensate q¯q2, , which parameterize the non-perturbative effects or soft gluons and quarks. We can consult Refs. [381, 385] for the basic techniques in performing the operator product expansion.

If there exist m heavy quark lines and n light-quark lines in the correlation functions Π(p2), each heavy quark line emits a gluon and each light quark line contributes a quark-antiquark pair, we obtain a quark-gluon operator,

GμνGαβq¯qq¯qq¯qq¯q,

which is of dimension 2m+3n, we should perform the operator product expansion up to the vacuum condensates of dimension 2m+3n at least [60, 61, 81-83]. For example, m=3 and n=4, we should calculate the vacuum condensates of dimension 18, see Fig.5. When m=n=1, we obtain the conventional Qq¯ mesons, it is obvious that we should calculate the mixed condensate q¯gsσGq.

Then we obtain the Källen-Lehmann representation through dispersion relation at the quark-gluon degrees of freedom,

ΠQCD(p2)=1πΔ2dsImΠQCD(s)sp2,

where the Δ2 denotes the thresholds.

At the hadron degrees of freedom, we insert a complete set of intermediate hadronic states |n with the same quantum numbers as the currents J(x) into the correlation functions Π(p2), and take account of the current-hadron couplings to obtain the analytical expressions, again we obtain the Källen−Lehmann representation through dispersion relation,

ΠH(p2)=1πΔ2dsImΠH(s)sp2,

where

ρH(s)=ImΠH(s)π=n|0|J(0)|n|2δ(sMn2),

the subscript H denotes the hadron side.

According to the Quark−Hadron duality, we introduce the continuum threshold parameters s0, and match the QCD side with hadron side of the correlation functions Π(p2),

1πΔ2s0dsImΠH(s)sp2=1πΔ2s0dsImΠQCD(s)sp2.

An important point is the choice of the continuum threshold s0, which is a physical parameter that should be determined from the hadronic spectrum. Then we perform the Borel transformation,

Π(T2)=B[Π(P2)]limP2,nP2/n=T2(P2)n+1(n)!(ddP2)nΠ(P2),

with P2=p2 to obtain the QCD sum rules,

1πΔ2s0dsImΠH(s)exp(sT2)=1πΔ2s0dsImΠQCD(s)exp(sT2),

where the T2 is the Borel parameter. Some typical and useful examples of the Borel transformation are given in the Appendix. If only the ground state is taken, then

ImΠH(s)π=λH2(sMH2),

we obtain the QCD sum rules,

λH2exp(MH2T2)=1πΔ2s0dsImΠQCD(s)exp(sT2),

where the MH is the mass of the ground state of the conventional hadron or multiquark state, the λH is the pole residue.

It is obvious that the Borel transformation wipes out any eventual subtraction terms in the correlation functions and suppresses the continuum contributions exponentially, therefore, it improves the convergent behavior of the dispersion integral. Furthermore, it suppresses the higher-dimensional operators in the operator product expansion factorially, which contain inverse powers of the P2, see Eq. (481), thus justifies truncation of the operator product expansion and favors a good convergent behavior.

Finally, we eliminate the pole residue λH to obtain the QCD sum rules for the ground state mass,

MH2=ddτΔ2s0dsImΠQCD(s)exp(τs)Δ2s0dsImΠQCD(s)exp(τs)τ=1T2.

In the QCD sum rules, we choose some phenomenological inputs which limit the accuracy of this method to be around 10%20% [388].

2.3 Are QCD sum rules reliable to study multiquark states

Any color singlet four-quark and five-quark currents J(x) can be written as J(x)=JAi(x)JBi(x), where the JAi(x) and JBi(x) are color singlet clusters with i=1, 2, 3, , for example,

Jμ(x)=εijkεimn2{ujT(x)Cγ5ck(x)d¯m(x)γμCc¯nT(x)ujT(x)Cγμck(x)d¯m(x)γ5Cc¯nT(x)}=122{ic¯iγ5cd¯γμuic¯γμcd¯iγ5u+c¯ud¯γμγ5cc¯γμγ5ud¯cic¯γνγ5cd¯σμνu+ic¯σμνcd¯γνγ5uic¯σμνγ5ud¯γνc+ic¯γνud¯σμνγ5c},

for the four-quark current [60, 401], and

J(x)=εilaεijkεlmnujT(x)Cγ5dk(x)umT(x)Cγ5cn(x)Cc¯aT(x),=14Sudγ5cc¯u+14Sudγλγ5cc¯γλu+18Sudσλτγ5cc¯σλτu+14Sudγλcc¯γλγ5u+i4Sudcc¯iγ5u+14Sudγ5uc¯c14Sudγλγ5uc¯γλc18Sudσλτγ5uc¯σλτc14Sudγλuc¯γλγ5ci4Suduc¯iγ5c,

for the five-quark current [402], where the components SudΓc=εijkuiTCγ5djΓck.

According to the pioneer works [403-405], in the coordinate space, we write the two-hadron-reducible contributions as

ΠRE(x)=0|T{JAi(x)JAj(0)}|00|T{JBi(x)JBj(0)}|0,

and the two-hadron-irreducible contributions as

ΠIR(x)=0|T{J(x)J(0)}|00|T{JAi(x)JAj(0)}|00|T{JBi(x)JBj(0)}|0.

At the phenomenological side of the correlation functions, we can write the two-hadron-reducible contributions as

ΠRE(p2)=|λAB|2i(2π)4d4qiq2MA2i(pq)2MB2,

where the couplings

0|JAi(0)|A(q)0|JBi(0)|B(pq)=λAB,

and the λAB could be estimated phenomenologically [403-406].

In the correlation functions for the color singlet-singlet type currents [407, 408], Lucha, Melikhov and Sazdjian assert that the Feynman diagrams can be divided into factorizable and nonfactorizable diagrams in the color space, the contributions at the order O(αsk) with k1, which are factorizable in the color space, are exactly canceled out by the meson-meson scattering states at the hadron side, the nonfactorizable diagrams, if have a Landau singularity, begin to make contributions to the tetraquark (molecular) states, the tetraquark (molecular) states begin to receive contributions at the order O(αs2), see Fig.6.

In Ref. [409], we examine the assertion of Lucha, Melikhov and Sazdjian in details and use two examples for the currents Jμ(x) and Jμν(x) to illustrate that the Landau equation is of no use in the QCD sum rule for the tetraquark molecular states, where

Jμ(x)=12[u¯(x)iγ5c(x)c¯(x)γμd(x)+u¯(x)γμc(x)c¯(x)iγ5d(x)],Jμν(x)=12[s¯(x)γμc(x)c¯(x)γνγ5s(x)s¯(x)γνγ5c(x)c¯(x)γμs(x)].

Firstly, we cannot assert that the factorizable Feynman diagrams in color space are exactly canceled out by the meson-meson scattering states, because the meson-meson scattering state and tetraquark molecular state both have four valence quarks, which can be divided into two color-neutral clusters. We cannot distinguish which Feynman diagrams contribute to the meson-meson scattering state or tetraquark molecular state based on the two color-neutral clusters [409].

Secondly, the quarks and gluons are confined objects, they cannot be put on the mass-shell, it is questionable to assert that the Landau equation is applicable for the quark-gluon bound states [410].

If we insist on applying the Landau equation to study the Feynman diagrams, we should choose the pole masses rather than the MS¯ masses to warrant mass poles. As the tetraquark (molecular) states begin to receive contributions at the order O(αs2) [407, 408], it is reasonable to take the pole masses m^Q as

m^Q=mQ(mQ)[1+43αs(mQ)π+f(αs(mQ)π)2+g(αs(mQ)π)3].

See Refs. [411, 412] for the explicit expressions of the f and g. If the Landau equation is applicable for the tetraquark (molecular) states, it is certainly applicable for the traditional charmonium and bottomonium states. In the case of the c-quark (b-quark), the pole mass m^c=1.67±0.07GeV (m^b=4.78±0.06GeV) from the Particle Data Group [411], the Landau singularity appears at the s-channel s=p2=2m^c=3.34±0.14GeV>Mηc/MJ/ψ (2m^b=9.56±0.12GeV>Mηb/MΥ). It is unreliable that the masses of the charmonium (bottomonium) states lie below the threshold 2m^c (2m^b) for the ηc and J/ψ (ηb and Υ) [409].

Thirdly, the nonfactorizable Feynman diagrams which have the Landau singularities begin to appear at the order O(αs0/αs1) rather than at the order O(αs2), and make contributions to the tetraquark molecular states, if the assertion (only nonfactorizable Feynman diagrams which have Landau singularities make contributions to the tetraquark molecular states) of Lucha, Melikhov and Sazdjian is right.

The nonfactorizable contributions appear at the order O(αs) due to the operators q¯gsGqq¯gsGq, which come from the Feynman diagrams shown in Fig.7. If we insist on choosing the pole mass and applying the landau equation to study the diagrams, we obtain a sub-leading Landau singularity at the s-channel s=p2=4m^c2. From the operators q¯gsGqq¯gsGq, we obtain the vacuum condensate q¯gsσGq2, where the gs2 is absorbed into the vacuum condensate. The nonfactorizable Feynman diagrams appear at the order O(αs0) or O(αs1), not at the order O(αs2) asserted in Refs. [407, 408].

In fact, for the triply-heavy dibaryon-type currents η(x) and ημ(x),

η(x)=JcT(x)Cγ5Jcc(x),ημ(x)=JcT(x)CγμJcc(x),Jc(x)=εijkqiT(x)Cγαqj(x)γαγ5ck(x),Jcc(x)=εijkciT(x)Cγαcj(x)γαγ5qk(x),

even in the lowest order Feynman diagrams, there are both connected and disconnected contributions in the color space [413], see Fig.8. From the first diagram in Fig.8, we can obtain both connected and disconnected Feynman diagrams, the connected contributions appear due to the quark-gluon operators q¯gsGqq¯gsGq [413], which are of the order O(αs1) and come from the Feynman diagrams shown in Fig.9.

Fourthly, the Landau equation serves as a kinematical equation in the momentum space, and does not depend on the factorizable and nonfactorizable properties of the Feynman diagrams in the color space.

In the leading order, the factorizable Feynman diagrams shown in Fig.10 can be divided into two color-neutral clusters, however, in the momentum space, they are nonfactorizable diagrams, the basic integrals are of the form,

d4qd4kd4l1(p+qk+l)2mc21q2mq21k2mq21l2mc2.

If we choose the pole masses, there is a Landau singularity at s=p2=(m^u+m^d+m^c+m^c)2, which is just a signal of a four-quark intermediate state. We cannot assert that it is a signal of a two-meson scattering state or a tetraquark molecular state, because the meson−meson scattering state and tetraquark molecular state both have four valence quarks, q, q¯, c and c¯, which form two color-neutral clusters.

Fifthly, only formal QCD sum rules for the tetraquark (molecular) states are obtained based on the assertion of Lucha, Melikhov and Sazdjian in Refs. [407, 408], no feasible QCD sum rules are obtained up to now.

Sixthly, we carry out the operator product expansion in the deep Euclidean space, p2, then obtain the physical spectral densities at the quark-gluon level through dispersion relation [376, 377],

ρQCD(s)=1πImΠ(s+iϵ)ϵ0,

where the Π(s) denotes the correlation functions. The Landau singularities require that the squared momentum p2=(m^u+m^d+m^c+m^c)2 in the Feynman diagrams, see Fig.10 and Eq. (57), it is questionable to perform the operator product expansion.

Seventhly, we choose the local four-quark or five-quark currents, while the traditional mesons and baryons are spatial extended objects and have mean spatial sizes r20, for example, r2E,Σc++=0.48fm, r2M,Σc++=0.83fm, r2M,Σc0=0.81fm from the lattice QCD, where the subscripts E and M stand for the electric and magnetic radii, respectively [414], r2M,Σc++=0.77fm, r2M,Σc0=0.52fm, r2M,Σc+=0.81fm, r2M,Ξc+=0.55fm, r2M,Ξc0=0.79fm from the self-consistent SU(3) chiral quark-soliton model [415], r2D+=0.43fm, r2D0=0.55fm from the light-front quark model [416], r2J/ψ=0.41fm, r2χc2=0.71fm from the screened potential model [88]. Local currents couple potentially to the compact objects having the average spatial sizes as that of the typical heavy mesons and baryons, not to the two-particle scattering states with average spatial size r21.0fm, which is too large to be interpolated by the local currents [417, 418].

Now we take a short digression to give a short notice. In the QCD sum rules, as we choose the local currents, the four-quark and five-quark states are all compact objects, they are 3¯3-type, 66¯-type, 11-type or 88-type tetraquark states, and 3¯3¯3¯-type or 11-type pentaquark states, although we usually call the 11-type states as the molecular states.

Now, let us write down the correlation functions Πμν(p) and Πμναβ(p) for the two currents shown in Eq. (54),

Πμν(p)=id4xeipx0|T{Jμ(x)Jν(0)}|0,

Πμναβ(p)=id4xeipx0|T{Jμν(x)Jαβ(0)}|0.

If we assume the four-quark currents couple potentially both to the two-meson scattering states and molecular states, then we can express the Jμ(x) and Jμν(x) in terms of the heavy meson fields,

Jμ(x)=12fDmD2mcfDmD[D0(x)Dμ(x)+Dμ0(x)D(x)]+12fDmD2mcfD0[D0(x)iμD0(x)+iμD00(x)D(x)]+λZZc,μ(x)+,

Jμν(x)=12fDsmDsfDs1mDs1[Ds,μ+(x)Ds1,ν(x)Ds1,ν+(x)Ds,μ(x)]12fDsmDsfDs[Ds,μ+(x)νDs(x)νDs+(x)Ds,μ(x)]+12fDs0fDs1mDs1[iμDs0+(x)Ds1,ν(x)Ds1,ν+(x)iμDs0(x)]12fDs0fDs[iμDs0+(x)νDs(x)νDs+(x)iμDs0(x)]λ~XεμναβiαXcβ(x)λ~X+[iμXc,ν+(x)iνXc,μ+(x)]+,

where we have taken the standard definitions for the decay constants of the traditional mesons and pole residues of the tetraquark molecular states,

0|Jμ(0)|Zc(p)=λZεμ(p),

0|Jμν(0)|Xc(p)=λ~Xεμναβεα(p)pβ,0|Jμν(0)|Xc+(p)=λ~X+[εμ(p)pνεν(p)pμ],

λX±=λ~X±MX±, the Zc, X and X+ have the JPC=1+, 1+ and 1++, respectively, the εμ(p) are the polarization vectors, we introduce the superscripts to denote the parity.

It is straightforward to obtain the hadronic representation,

Πμν(p)=Π(p2)(gμν+pμpνp2)+,

Πμναβ(p)=Π(p2)(gμαgνβgμβgναgμαpνpβp2gνβpμpαp2+gμβpνpαp2+gναpμpβp2)

+Π+(p2)(gμαpνpβp2gνβpμpαp2+gμβpνpαp2+gναpμpβp2),

where

Π(p2)=λZ2MZ2p2+ΠTW(p2)+,Π(p2)=PμναβΠμναβ(p)=λX2MX2p2+ΠTW(p2)+,Π+(p2)=P+μναβΠμναβ(p)=λX+2MX+2p2+,

we project out the components Π(p2) and Π+(p2) by introducing the operators Pμναβ and P+μναβ respectively,

Pμναβ=16(gμαpμpαp2)(gνβpνpβp2),P+μναβ=16(gμαpμpαp2)(gνβpνpβp2)16gμαgνβ,

ΠTW(p2)=λDD216π2Δ12s0ds1sp2λ(s,mD2,mD2)s[1+λ(s,mD2,mD2)12smD2]+λDD0216π2Δ22s0ds1sp2λ(s,mD2,mD02)sλ(s,mD2,mD02)12s+,

ΠTW(p2)=λDsDs1216π2Δ32s0ds1sp2λ(s,mDs2,mDs12)s[1+λ(s,mDs2,mDs12)12smDs2λ(s,mDs2,mDs12)12smDs12]+λDsDs216π2Δ42s0ds1sp2λ(s,mDs2,mDs2)sλ(s,mDs2,mDs2)12s+λDs0Ds1216π2Δ52s0ds1sp2λ(s,mDs02,mDs12)sλ(s,mDs02,mDs12)12s+,

where λDD2=fD2mD4fD2mD2mc2, Δ12=(mD+mD)2, λDD02=fD2mD4fD02mc2, Δ22=(mD+mD0)2, λDsDs12=fDs2mDs2fDs12mDs12, Δ32=(mDs+mDs1)2, λDsDs2=fDs2mDs2fDs2, Δ42=(mDs+mDs)2, λDs0Ds12=fDs02fDs12mDs12, Δ52=(mDs0+mDs1)2 and λ(a,b,c)=a2+b2+c22ab2bc2ca.

The traditional hidden-flavor mesons have the normal quantum numbers, JPC=0+, 0++, 1, 1+, 1++, 2, 2+, 2++, . The component Π(p2) receives contributions with the exotic quantum numbers JPC=1+, the component Π+(p2) receives contributions with the normal quantum numbers JPC=1++. We choose the component Π(p2) with the exotic quantum numbers JPC=1+ and discard the component Π+(p2) with the normal quantum numbers JPC=1++. Thereafter, we will neglect the superscript in the Xc for simplicity.

According to the assertion of Lucha, Melikhov and Sazdjian [407, 408], all contributions of the order O(αsk) with k1 are exactly canceled out by the two-meson scattering states, we set

Π(p2)=ΠTW(p2)+,Π(p2)=ΠTW(p2)+,

at the hadron side [409]. Then let us take the quark−hadron duality below the continuum threshold s0 and perform Borel transformation with respect to the variable P2=p2 to obtain the QCD sum rules:

ΠTW(T2)=λDD216π2Δ12s0dsλ(s,mD2,mD2)s[1+λ(s,mD2,mD2)12smD2]exp(sT2)+λDD0216π2Δ22s0dsλ(s,mD2,mD02)sλ(s,mD2,mD02)12sexp(sT2)=κ4mc2s0dsρZ,QCD(s)exp(sT2),

ΠTW(T2)=λDsDs1216π2Δ32s0dsλ(s,mDs2,mDs12)s[1+λ(s,mDs2,mDs12)12smDs2λ(s,mDs2,mDs12)12smDs12]exp(sT2)+λDsDs216π2Δ42s0dsλ(s,mDs2,mDs2)sλ(s,mDs2,mDs2)12sexp(sT2)+λDs0Ds1216π2Δ52s0dsλ(s,mDs02,mDs12)sλ(s,mDs02,mDs12)12sexp(sT2)=κ4mc2s0dsρX,QCD(s)exp(sT2),

the explicit expressions of the QCD spectral densities ρZ,QCD(s) and ρX,QCD(s) are given in Ref. [409]. We introduce the parameter κ to measure the deviations from 1, if κ1, we could get the conclusion tentatively that the two-meson scattering states can saturate the QCD sum rules. Then we differentiate Eqs. (72) and (73) with respect to 1T2, and obtain two additional QCD sum rules,

dΠTW(T2)d(1/T2)=κdd(1/T2)4mc2s0dsρZ,QCD(s)exp(sT2),

dΠTW(T2)d(1/T2)=κdd(1/T2)4mc2s0dsρX,QCD(s)exp(sT2).

Thereafter, we will denote the QCD sum rules in Eqs. (74) and (75) as the QCDSR I, and the QCD sum rules in Eqs. (72) and (73) as the QCDSR II.

On the other hand, if the two-meson scattering states cannot saturate the QCD sum rules, we have to introduce the tetraquark molecular states to saturate the QCD sum rules,

λZ/X2exp(MZ/X2T2)=4mc2s0dsρZ/X,QCD(s)exp(sT2),

then we differentiate Eq. (76) with respect to 1T2, and obtain two QCD sum rules for the masses of the tetraquark molecular states,

MZ/X2=dd(1/T2)4mc2s0dsρZ/X,QCD(s)exp(sT2)4mc2s0dsρZ/X,QCD(s)exp(sT2).

In Fig.11, we plot the values of the κ with variations of the Borel parameters T2 with the continuum threshold parameters s0=4.40GeV and 5.15GeV for the u¯cc¯d and s¯cc¯s two-meson scattering states, respectively. From Fig.11, we can see explicitly that the values of the κ increase monotonically and quickly with the increase of the Borel parameters T2, no platform appears, which indicates that the QCD sum rules in Eqs. (72) and (73) obtained according to the assertion of Lucha, Melikhov and Sazdjian are unreasonable. Reasonable QCD sum rules lead to platforms flat enough or not flat enough, rather than no evidence of platforms, the two-meson scattering states cannot saturate the QCD sum rules.

We saturate the hadron side of the QCD sum rules with the tetraquark molecular states alone, and study the QCD sum rues shown in Eqs. (76) and (77). In Fig.12, we plot tetraquark molecule masses with variations of the Borel parameters T2. From the Fig.12, we observe that there appear Borel platforms in the Borel windows indeed, the relevant results are shown explicitly in Tab.1, we adopt the energy scale formula μ=MX/Z24×(1.84GeV)2 to choose the best energy scales of the QCD spectral densities. The tetraquark molecular states alone can satisfy the QCD sum rules [409]. We obtain the prediction MZ=3.89±0.09GeV and MX=4.67±0.08GeV, which are consistent with the Zc(3900) observed by the BESIII and Belle Collaborations [162, 163] and the X(4630) observed one-year latter by the LHCb Collaboration [115]. If we have taken account of the light-flavor SU(3) breaking effects of the energy scale formula, the fit between the theoretical calculation and experimental measurement would be better for the X(4630).

The local currents do suppress but do not forbid the couplings between the four-quark currents and two-meson scattering states, as the overlaps of the wave-functions are very small [417], furthermore, the quantum field theory does not forbid the couplings between the four-quark currents and two-meson scattering states if they have the same quantum numbers. We study the contributions of the intermediate meson−meson scattering states DD¯, J/ψπ, J/ψρ, etc besides the tetraquark molecular state Zc to the correlation function Πμν(p) as an example,

Πμν(p)=λ^Z2p2M^Z2ΣDD(p2)ΣJ/ψπ(p2)ΣJ/ψρ(p2)+g~μν(p)+,

where g~μν(p)=gμν+pμpνp2. We choose the bare quantities λ^Z and M^Z to absorb the divergences in the self-energies ΣDD¯(p2), ΣJ/ψπ(p2), ΣJ/ψρ(p2), etc. The renormalized energies satisfy the relation p2MZ2Σ¯DD(p2)Σ¯J/ψπ(p2)Σ¯J/ψρ(p2)+=0, where the overlines above the self-energies denote that the divergent terms have been subtracted. As the tetraquark molecular state Zc is unstable, the relation should be modified, p2MZ2ReΣ¯DD(p2)ReΣ¯J/ψπ(p2)ReΣ¯J/ψρ(p2)+=0, and ImΣ¯DD(p2)ImΣ¯J/ψπ(p2)ImΣ¯J/ψρ(p2)+=p2Γ(p2). The renormalized self-energies contribute a finite imaginary part to modify the dispersion relation,

Πμν(p)=λZ2p2MZ2+ip2Γ(p2)g~μν(p)+.

If we assign the Zc(3900) to be the DD¯+DD¯ tetraquark molecular state with the JPC=1+ [81, 82, 83], the physical width ΓZc(3900)(MZ2)=28.2±2.6MeV from the Particle Data Group [411].

We take account of the finite width effect by the simple replacement of the hadronic spectral density,

λZ2δ(sMZ2)λZ21πMZΓZ(s)(sMZ2)2+MZ2ΓZ2(s),

where

ΓZ(s)=ΓZMZss(MD+MD)2MZ2(MD+MD)2.

Then the hadron sides of the QCD sum rules in Eq. (76) and Eq. (77) undergo the following changes,

λZ2exp(MZ2T2)λZ2(mD+mD)2s0ds1πMZΓZ(s)(sMZ2)2+MZ2ΓZ2(s)exp(sT2)=(0.780.79)λZ2exp(MZ2T2),

λZ2MZ2exp(MZ2T2)λZ2(mD+mD)2s0dss1πMZΓZ(s)(sMZ2)2+MZ2ΓZ2(s)exp(sT2)=(0.800.81)λZ2MZ2exp(MZ2T2),

for the value s0=4.40GeV. We can absorb the numerical factors 0.780.79 and 0.800.81 into the pole residue with the simple replacement λZ0.89λZ safely, the intermediate meson-loops cannot affect the mass MZ significantly, but affect the pole residue remarkably, which are consistent with the fact that we obtain the masses of the tetraquark molecular states from a fraction, see Eq. (77).

We obtain the conclusion confidently that it is reliable to study the multiquark states with the QCD sum rules, the contaminations from the two-particle scattering states play a tiny role [409].

2.4 Energy scale dependence of the QCD sum rules

In calculating the Feynman diagrams, we usually adopt the dimensional regularization to regularize the divergences, and resort to wave-function, quark-mass and current renormalizations to absorb the ultraviolet divergences, and resort to the vacuum condensate redefinitions to absorb the infrared divergences. Thus, the correlation functions Π(p2) are free of divergences. And we expect to calculate the Π(p2) at any energy scale μ at which perturbative calculations are feasible, and the physical quantities are independent on the specified energy scale μ. Roughly speaking, the correlation functions Π(p2) are independent on the energy scale approximately,

ddμΠ(p2)=0,

at least the bare Π(p2) are independent on the energy scale.

We write down the correlation functions Π(p2) for the hidden-charm (or hidden-bottom) four-quark currents, the most commonly chosen currents in studying the X, Y and Z states, in the Källen−Lehmann representation,

Π(p2)=4mQ2(μ)s0dsρQCD(s,μ)sp2+s0dsρQCD(s,μ)sp2.

In fact, there are subtraction terms neglected at the right side of Eq. (85), which could be deleted after performing the Borel transformation. The Π(p2) should be independent on the energy scale we adopt to perform the operator product expansion, but which does not mean

ddμ4mQ2(μ)s0dsρQCD(s,μ)sp20,

due to the two features inherited from the QCD sum rules:

● Perturbative corrections are neglected, even in the QCD sum rules for the traditional mesons, we cannot take account of the radiative corrections up to arbitrary orders, for example, we only have calculated the radiative corrections up to the order O(αs2) for the pseudoscalar D/B mesons up to now [419-421]. The higher dimensional vacuum condensates are factorized into lower dimensional ones based on the vacuum saturation, for example,

ψ¯Γψψ¯Γψ=1144ψ¯ψ2[Tr(Γ)Tr(Γ)Tr(ΓΓ)],

where ψ=u, d or s, Tr=TrDTrC, the subscripts denote the Dirac spinor and color spaces, respectively, therefore the energy scale dependence of the higher dimensional vacuum condensates is modified.

● Truncations s0 which are physical quantities determined by the experimental data set in, the correlations between the thresholds 4mQ2(μ) and continuum thresholds s0 are unknown. Quark−hadron duality is just an assumption.

After performing the Borel transformation, we obtain the integrals

4mQ2(μ)s0dsρQCD(s,μ)exp(sT2),

which are sensitive to the Q-quark mass mQ(μ), in other words, the energy scale μ. Variations of the energy scale μ can lead to changes of integral ranges 4mQ2(μ)s0 of the variable ds besides the QCD spectral densities ρQCD(s,μ), therefore changes of the Borel windows and predicted hadron masses and pole residues. The strong fine-structure αs=gs24π appears even in the tree-level,

q¯γμtaqDηgsGλτa=gηλgμτgητgμλ274παsq¯q2,

where ta=λa2. Thus we have to deal with the energy scale dependence of the QCD sum rules.

Let us take a short digression and perform some phenomenological analysis. We can describe the heavy four-quark systems QQ¯qq¯ by a double-well potential with two light quarks qq¯ lying in the two wells, respectively. In the heavy quark limit, the Q-quark serves as a static well potential and attracts with the light quark q to form a heavy diquark DqQi in color antitriplet,

q+QDqQi,

or attracts with the light antiquark q¯ to form a meson-like color-singlet cluster (meson-like color-octet cluster),

q¯+Qq¯Q(q¯λaQ),

the Q¯-quark serves as another static well potential and attracts with the light antiquark q¯ to form a heavy antidiquark Dq¯Q¯i in color triplet,

q¯+Q¯Dq¯Q¯i,

or attracts with the light quark q to form a heavy meson-like color-singlet cluster (meson-like color-octet-cluster),

q+Q¯Q¯q(Q¯λaq).

Then

DqQi+Dq¯Q¯i3¯3typetetraquarkstates,q¯Q+Q¯q11typetetraquarkstates,q¯λaQ+Q¯λaq88typetetraquarkstates,

the two heavy quarks Q and Q¯ stabilize the four-quark systems qq¯QQ¯, just as in the case of the (μe+)(μ+e) molecule in QED [57].

We can also describe the hidden-charm (or hidden-bottom) five-quark systems qq1q2QQ¯ by a double-well potential. In the heavy quark limit, the Q-quark (Q¯-quark) serves as a static well potential, the diquark Dq1q2j and quark q lie in the two wells, respectively,

q1+q2+Q¯Dq1q2j+Q¯kTq1q2Q¯i,q+QDqQi,

or

q1+q2+QDq1q2j+Qjq1q2Q,q+Q¯qQ¯,

where the Tq1q2Q¯i denotes the heavy triquark in the color triplet. Then

DqQi+Tq1q2Q¯i3¯3¯3¯typepentaquarkstates,q1q2Q+Q¯q11typepentaquarkstates.

Now we can obtain the conclusion tentatively that the heavy tetraquark states are characterized by the effective heavy quark masses MQ (or constituent quark masses) and the virtuality V=MX/Y/Z2(2MQ)2 (or bound energy not as robust) [81, 422, 423, 424].

In summary, the QCD sum rules have three typical energy scales μ2, T2, V2. It is natural to set the energy scales as,

μ2=V2=O(T2),

and we obtain the energy scale formula [424],

μ=MX/Y/Z/P2(2MQ)2,

which works very well for the tetraquark states and pentaquark states. It can improve the convergence of the operator product expansion remarkably and enhance the pole contributions remarkably.

We usually set the small u and d quark masses to be zero, and take account of the light-flavor SU(3) breaking effects by introducing an effective s-quark mass Ms, thus we reach the modified energy scale formula,

μ=MX/Y/Z2(2MQ)2κMs,

to choose the suitable energy scales of the QCD spectral densities [61, 82, 83, 172, 425], where the κ=0, 1 and 2 denote the numbers of the s-quarks, the MQ and Ms have universal values to be commonly used elsewhere.

We can rewrite the energy scale formula in Eq. (99) in the following form,

MX/Y/Z/P2=μ2+Constants,

where the Constants have the values 4MQ2 [426]. As we cannot obtain energy scale independent QCD sum rules, we conjecture that the predicted multiquark masses and the pertinent energy scales of the QCD spectral densities have a Regge-trajectory-like relation, see Eq. (101), where the Constants are free parameters and fitted by the QCD sum rules. Direct calculations have proven that the Constants have universal values and work well. We take account of the light-flavor SU(3) breaking effects, and write down the modified energy scale formula [418],

MX/Y/Z/P2=(μ+κMs)2+Constants.

In Ref. [60], we take the X(3872) and Zc(3900) as the hidden-charm tetraquark with the JPC=1++ and 1+, respectively, and explore the energy scale dependence of the QCD sum rules for the exotic states for the first time. In Fig.13, we plot the mass of the Zc(3900) with variations of the Borel parameter T2 and energy scale μ for the continuum threshold parameter s0=4.4GeV. From the figure, we can see clearly that the mass decreases monotonously with increase of the energy scale. The energy scale μ=1.4GeV is the lowest energy scale to reproduce the experimental data [162, 163].

There are three schemes to choose the input parameters at the QCD side of the QCD sum rules:

Scheme I. We take the energy scale formula and its modifications, see Eqs. (99) and (100), to choose the energy scales of the QCD spectral densities in a consistent way.

Scheme II. We take the MS¯ (modified-minimal-subtraction) masses for the heavy quarks mQ(mQ), and take the light quark masses and vacuum condensates at the energy scale μ=1GeV.

Scheme III. We take all the input parameters at the energy scale μ=1GeV.

The Scheme II is adopted in most QCD sum rules [46, 47, 58, 180, 427-461] ([462-474]), where the MS¯ masses mQ(mQ) are usually smaller than (or equal to) the values from the Particle Data Group (with much smaller pole contributions) [97].

The Scheme III was adopted in early works of Wang and his collaborators in 2009−2011 [475-486], where many elegant four-quark currents were constructed originally.

In Ref. [487], we study the pentaquark molecular states in the three schemes in details and examine their advantages and shortcomings, in Schemes I and III, we truncate the operator product expansion up to the vacuum condensates of D=13, while in Scheme II, we truncate the operator product expansion up to the vacuum condensates of D=10, which is commonly adopted in this case.

We write down the correlation functions Π(p), Πμν(p) and Πμναβ(p) firstly,

Π(p)=id4xeipx0|T{J(x)J¯(0)}|0,Πμν(p)=id4xeipx0|T{Jμ(x)J¯ν(0)}|0,Πμναβ(p)=id4xeipx0|T{Jμν(x)J¯αβ(0)}|0,

where the currents J(x)=JD¯Σc(x), Jμ(x)=JμD¯Σc(x), JμD¯Σc(x), Jμν(x)=JμνD¯Σc(x),

JD¯Σc(x)=c¯(x)iγ5u(x)εijkuiT(x)Cγαdj(x)γαγ5ck(x),JμD¯Σc(x)=c¯(x)iγ5u(x)εijkuiT(x)Cγμdj(x)ck(x),JμD¯Σc(x)=c¯(x)γμu(x)εijkuiT(x)Cγαdj(x)γαγ5ck(x),JμνD¯Σc(x)=c¯(x)γμu(x)εijkuiT(x)Cγνdj(x)ck(x)+(μν).

We separate the contributions of the molecular states with the JP=12±, 32± and 52± unambiguously, then we introduce the weight functions sexp(sT2) and exp(sT2) to obtain the QCD sum rules at the hadron side,

4mc2s0ds[sρj,H1(s)±ρj,H0(s)]exp(sT2)=2Mλj2exp(M2T2),

where the s0 are the continuum threshold parameters and the T2 are the Borel parameters, the ρj,H1(s) and ρj,H0(s) with the j=12, 32 and 52 are hadronic spectral densities, the λj± are the pole residues.

We perform the operator product expansion to obtain the analytical QCD spectral densities ρj,QCD1(s) and ρj,QCD0(s) through dispersion relation, then we take the quark-hadron duality below the continuum thresholds s0 and introduce the weight functions sexp(sT2) and exp(sT2) to obtain the QCD sum rules:

2Mλj2exp(M2T2)=4mc2s0ds[sρj,QCD1(s)+ρj,QCD0(s)]exp(sT2).

For the technical details, one can consult Ref. [487] or Section 5.1.

We differentiate Eq. (106) with respect to τ=1T2, then eliminate the pole residues to obtain the molecule masses,

M2=ddτ4mc2s0ds[sρQCD1(s)+ρQCD0(s)]exp(τs)4mc2s0ds[sρQCD1(s)+ρQCD0(s)]exp(τs),

where ρQCD1(s)=ρj,QCD1(s) and ρQCD0(s)=ρj,QCD0(s).

We show the Borel windows T2, continuum threshold parameters s0, energy scales of the QCD spectral densities and pole contributions of the ground states explicitly in Tab.2. In the Scheme III, the pole contributions are less than 25%, which are too small, and the Scheme III could be abandoned. On the other hand, the convergent behaviors of the operator product expansion have relation Scheme I > Scheme II > Scheme III.

At last, we take account of all uncertainties of the input parameters, and obtain the masses and pole residues of the molecular states, which are shown explicitly in Tab.3 and Fig.14 and Fig.15.

In Fig.14 and Fig.15, we plot the masses at much larger ranges of the Borel parameters than the Borel windows. The predicted masses in the Scheme I decrease monotonously and quickly with increase of the Borel parameters at the region T22.0GeV2, then reach small platforms and increase slowly with increase of the Borel parameters. The predicted masses in the Scheme II increase monotonously and quickly with increase of the Borel parameters at the region T2<2.6GeV2, then increase slowly and steadily with increase of the Borel parameters. It is obvious that the flatness of the platforms have relation Scheme I > Scheme II.

Both the predictions in the Schemes I and II support assigning the Pc(4312) as the D¯Σc molecular state with the JP=12, assigning the Pc(4380) as the D¯Σc molecular state with the JP=32, assigning the Pc(4440/4457) as the D¯Σc molecular state with the JP=32 or the D¯Σc molecular state with the JP=52. However, if we take account of the vacuum condensates of the dimensions of D=11 and 13 in the Scheme II, we would obtain a mass about 200MeV larger than the corresponding one given in Tab.3, so we prefer the Scheme I [487].

3 3 ¯3 type tetraquark states

3.1 Hidden-heavy tetraquark states

The scattering amplitude for one-gluon exchange is proportional to,

(λa2)ij(λa2)kl=13(δijδklδilδkj)+16(δijδkl+δilδkj),

the negative (positive) sign in front of the antisymmetric antitriplet 3¯ (symmetric sextet 6) indicates the interaction is attractive (repulsive), which favors (disfavors) formation of the diquarks in color 3¯ (6). We usually construct the 3¯3 type color-singlet four-quark currents J3¯3 to interpolate the tetraquark states,

J3¯3=εkijεkmnqiTCΓqjq¯mΓCq¯nT,

where the q and q are the quarks, the Γ and Γ are the Dirac γ-matrixes. The color factor has the relation,

εkijεkmn=δimδjnδinδjm.

With the simple replacement,

δimδjnδinδjmδimδjn+δinδjm,

we obtain the corresponding 66¯ type currents J66¯,

J66¯=qiTCΓqjq¯iΓCq¯jT+qiTCΓqjq¯jΓCq¯iT.

If the J66¯ type currents satisfy the Fermi−Dirac statistics, the quantum field theory does not forbid their existence. In fact, in the potential quark models, we usually take both the 3¯3 and 66¯ diquark configurations, see Section 2.1.5. The 66¯ type currents are also widely used in literatures [34, 42, 121, 134, 185, 437, 488-493].

Or we construct the 88 type currents directly to interpolate the exotic states [455, 459, 494-497], although the one-gluon exchange induced interaction is repulsive in this channel, see Section 2.1.5. In Ref. [496], we take the Zc(4200) as the 88 type axial-vector molecule-like state, and construct the 88 type axial-vector current to study its mass and width with the QCD sum rules, the numerical results support assigning the Zc(4200) as the 88 type molecule-like state with the JPC=1+. Furthermore, we discuss the possible assignments of the Zc(3900), Zc(4200) and Z(4430) as the 3¯3 type tetraquark states with the JPC=1+.

3.1.1 Tetraquark states with positive parity

The one-gluon-exchange induced attractive interactions favor formation of the diquarks in the color antitriplet, flavor antitriplet and spin singlet [311], while the most favored configurations are the scalar and axialvector diquark states [498-504]. The QCD sum rules indicate that the heavy-light scalar and axialvector diquark states have almost degenerate masses [498, 499, 500], while the masses of the light axialvector diquark states lie about (150200)MeV above that of the light scalar diquark states [501-504], if they have the same valence quarks.

The diquarks εijkqjTCΓqk in the 3¯ have five structures in the Dirac spinor space, where CΓ=Cγ5, C, Cγμγ5, Cγμ and Cσμν for the scalar, pseudoscalar, vector, axialvector and tensor diquarks, respectively. In the non-relativistic quark model, a P-wave changes the parity by contributing a factor ()L= with the angular momentum L=1. The Cγ5 and Cγμ diquark states have the spin-parity JP=0+ and 1+, respectively, the corresponding C and Cγμγ5 diquark states have the spin-parity JP=0 and 1, respectively, the effects of the P-waves are embodied in the underlined γ5 in the Cγ5γ5_ and Cγμγ5_. The tensor diquark states have both the JP=1+ and 1 components, we project out the 1+ and 1 components explicitly, and introduce the symbols A~ and V~ to represent them respectively. We would like to give an example on the heavy-light tensor diquarks to illustrate how to perform the projection.

Under parity transformation P^, the tensor diquarks have the properties,

P^εijkqjT(x)Cσμνγ5Qk(x)P^1=εijkqjT(x~)Cσμνγ5Qk(x~),P^εijkqjT(x)CσμνQk(x)P^1=εijkqjT(x~)CσμνQk(x~),

where the four vectors xμ=(t,x) and x~μ=(t,x). We introduce the four vector tμ=(1,0) and project out the 1+ and 1 components explicitly [505],

P^εijkqjT(x)Cσμνtγ5Qk(x)P^1=+εijkqjT(x~)Cσμνtγ5Qk(x~),P^εijkqjT(x)Cσμνvγ5Qk(x)P^1=εijkqjT(x~)Cσμνvγ5Qk(x~),P^εijkqjT(x)CσμνvQk(x)P^1=+εijkqjT(x~)CσμνvQk(x~),P^εijkqjT(x)CσμνtQk(x)P^1=εijkqjT(x~)CσμνtQk(x~),

where σμνt=i2[γμt,γνt], σμνv=i2[γμv,γνt], γμv=γttμ, γμt=γμγttμ. We can also introduce the P-wave explicitly in the Cγ5 and Cγμ diquarks and obtain the vector diquarks εijkqjTCγ5μqk and tensor diquarks εijkqjTCγμνqk, where the derivative μ=μμ embodies the P-wave effects.

We can also adopt the covariant derivative with the simple replacement μDμ=μigsGμ, then the four-quark currents are gauge covariant, however, gluonic components are introduced, we have to deal with both valence quarks and gluons.

Now let us construct the 3¯3-type four-quark currents to interpolate the hidden-charm tetraquark states with the JPC=0++, 1+, 1++ and 2++,

JSS(x)=εijkεimnujT(x)Cγ5ck(x)d¯m(x)γ5Cc¯nT(x),JAA(x)=εijkεimnujT(x)Cγμck(x)d¯m(x)γμCc¯nT(x),JA~A~(x)=εijkεimnujT(x)Cσμνvck(x)d¯m(x)σvμνCc¯nT(x),JVV(x)=εijkεimnujT(x)Cγμγ5ck(x)d¯m(x)γ5γμCc¯nT(x),JV~V~(x)=εijkεimnujT(x)Cσμνtck(x)d¯m(x)σtμνCc¯nT(x),JPP(x)=εijkεimnujT(x)Cck(x)d¯m(x)Cc¯nT(x),

J,μSA(x)=εijkεimn2[ujT(x)Cγ5ck(x)d¯m(x)γμCc¯nT(x)ujT(x)Cγμck(x)d¯m(x)γ5Cc¯nT(x)],J,μνAA(x)=εijkεimn2[ujT(x)Cγμck(x)d¯m(x)γνCc¯nT(x)ujT(x)Cγνck(x)d¯m(x)γμCc¯nT(x)],J,μνSA~(x)=εijkεimn2[ujT(x)Cγ5ck(x)d¯m(x)σμνCc¯nT(x)ujT(x)Cσμνck(x)d¯m(x)γ5Cc¯nT(x)],J,μA~A(x)=εijkεimn2[ujT(x)Cσμνγ5ck(x)d¯m(x)γνCc¯nT(x)ujT(x)Cγνck(x)d¯m(x)γ5σμνCc¯nT(x)],J,μV~V(x)=εijkεimn2[ujT(x)Cσμνck(x)d¯m(x)γ5γνCc¯nT(x)+ujT(x)Cγνγ5ck(x)d¯m(x)σμνCc¯nT(x)],J,μνVV(x)=εijkεimn2[ujT(x)Cγμγ5ck(x)d¯m(x)γ5γνCc¯nT(x)ujT(x)Cγνγ5ck(x)d¯m(x)γ5γμCc¯nT(x)],J,μPV(x)=εijkεimn2[ujT(x)Cck(x)d¯m(x)γ5γμCc¯nT(x)+ujT(x)Cγμγ5ck(x)d¯m(x)Cc¯nT(x)],

J+,μSA(x)=εijkεimn2[ujT(x)Cγ5ck(x)d¯m(x)γμCc¯nT(x)+ujT(x)Cγμck(x)d¯m(x)γ5Cc¯nT(x)],J+,μνSA~(x)=εijkεimn2[ujT(x)Cγ5ck(x)d¯m(x)σμνCc¯nT(x)+ujT(x)Cσμνck(x)d¯m(x)γ5Cc¯nT(x)],J+,μV~V(x)=εijkεimn2[ujT(x)Cσμνck(x)d¯m(x)γ5γνCc¯nT(x)ujT(x)Cγνγ5ck(x)d¯m(x)σμνCc¯nT(x)],J+,μA~A(x)=εijkεimn2[ujT(x)Cσμνγ5ck(x)d¯m(x)γνCc¯nT(x)+ujT(x)Cγνck(x)d¯m(x)γ5σμνCc¯nT(x)],J+,μPV(x)=εijkεimn2[ujT(x)Cck(x)d¯m(x)γ5γμCc¯nT(x)ujT(x)Cγμγ5ck(x)d¯m(x)Cc¯nT(x)],J+,μνAA(x)=εijkεimn2[ujT(x)Cγμck(x)d¯m(x)γνCc¯nT(x)+ujT(x)Cγνck(x)d¯m(x)γμCc¯nT(x)],J+,μνVV(x)=εijkεimn2[ujT(x)Cγμγ5ck(x)d¯m(x)γ5γνCc¯nT(x)+ujT(x)Cγνγ5ck(x)d¯m(x)γ5γμCc¯nT(x)],

where the subscripts ± denote the positive and negative charge conjugation, respectively, the superscripts or subscripts P, S, A(A~) and V(V~) denote the pseudoscalar, scalar, axialvector and vector diquark and antidiquark operators, respectively [61]. With the simple replacements,

uq,d¯s¯,

where q=u, d, we obtain the corresponding currents for the cc¯qs¯ states [172]. Again, with the simple replacements,

us,d¯s¯,

we obtain the corresponding currents for the cc¯ss¯ states [425, 506, 507].

We introduce the symbols,

J(x)=JSS(x),JAA(x),JA~A~(x),JVV(x),JV~V~(x),JPP(x),Jμ(x)=J,μSA(x),J,μA~A(x),J,μV~V(x),J,μPV(x),J+,μSA(x),J+,μV~V(x),J+,μA~A(x),J+,μPV(x),

Jμν(x)=J,μνAA(x),J,μνSA~(x),J,μνVV(x),J+,μνSA~(x),J+,μνAA(x),J+,μνVV(x),

for simplicity.

Under parity transformation P^, the currents have the properties,

P^J(x)P^1=+J(x~),P^Jμ(x)P^1=Jμ(x~),P^JμνSA~(x)P^1=JSA~μν(x~),P^JμνAA/VV(x)P^1=+JAA/VVμν(x~),

where xμ=(t,x) and x~μ=(t,x), and we have neglected other superscripts and subscripts.

The currents J(x), Jμ(x) and Jμν(x) have the symbolic quark constituent c¯cd¯u with the isospin I=1 and I3=1, other currents in the isospin multiplets can be constructed analogously, for example, we write down the isospin singlet current for the JSS(x) directly,

JSSI=0(x)=εijkεimn2[ujT(x)Cγ5ck(x)u¯m(x)γ5Cc¯nT(x)+djT(x)Cγ5ck(x)d¯m(x)γ5Cc¯nT(x)].

In the isospin limit, the currents with the symbolic quark constituents c¯cd¯u, c¯cu¯d, c¯cu¯ud¯d2, c¯cu¯u+d¯d2 couple potentially to the hidden-charm tetraquark states with degenerated masses, the currents with the isospins I=1 and 0 lead to the same QCD sum rules. Thereafter, we will denote the Zc states as the isospin triplet, and the X states as the isospin singlet,

Zc:c¯cd¯u,c¯cu¯d,c¯cu¯ud¯d2,X:c¯cu¯u+d¯d2.

Accordingly, the Zcs states with the symbolic quark constituents c¯cs¯q and isospin I=12 have degenerated masses.

The currents with the symbolic quark constituents c¯cu¯ud¯d2 and c¯cu¯u+d¯d2 have definite charge conjugation. We would like to assume that the c¯cd¯u type tetraquark states have the same charge conjugation as their charge-neutral cousins.

Under charge conjugation transformation C^, the currents J(x), Jμ(x) and Jμν(x) have the properties,

C^J(x)C^1=+J(x)ud,C^J±,μ(x)C^1=±J±,μ(x)ud,C^J±,μν(x)C^1=±J±,μν(x)ud,

where we have neglected other superscripts and subscripts.

Now we write down the correlation functions Π(p), Πμν(p) and Πμναβ(p),

Π(p)=id4xeipx0|T{J(x)J(0)}|0,Πμν(p)=id4xeipx0|T{Jμ(x)Jν(0)}|0,Πμναβ(p)=id4xeipx0|T{Jμν(x)Jαβ(0)}|0.

At the hadron side, we insert a complete set of intermediate hadronic states with the same quantum numbers as the currents J(x), Jμ(x) and Jμν(x) into the correlation functions Π(p), Πμν(p) and Πμναβ(p) to obtain the hadronic representation [376, 377], and isolate the ground state hidden-charm tetraquark contributions,

Π(p)=λZ+2MZ+2p2+=Π+(p2),Πμν(p)=λZ+2MZ+2p2(gμν+pμpνp2)+=Π+(p2)(gμν+pμpνp2)+,ΠμναβAA,(p)=λ~Z+2MZ+2p2(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+λ~Z2MZ2p2(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+

=Π~+(p2)(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+Π~(p2)(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ),

ΠμναβSA~,±(p)=λ~Z2MZ2p2(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+λ~Z+2MZ+2p2(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+=Π~(p2)(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+Π~+(p2)(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ),ΠμναβAA,+(p)=λZ+2MZ+2p2(g~μαg~νβ+g~μβg~να2g~μνg~αβ3)+,=Π+(p2)(g~μαg~νβ+g~μβg~να2g~μνg~αβ3)+,

where g~μν=gμνpμpνp2. We add the superscripts ± in the ΠμναβAA,(p), ΠμναβSA~,±(p) and ΠμναβAA,+(p) to denote the positive and negative charge conjugation, respectively, and add the superscripts (subscripts) ± in the Zc± (the Π±(p2) and Π~±(p2) components) to denote the positive and negative parity, respectively. And the ΠμναβVV,±(p) and ΠμναβAA,±(p) have the same tensor structures. The pole residues λZ± are defined by

0|J(0)|Zc+(p)=λZ+,0|Jμ(0)|Zc+(p)=λZ+εμ,0|J±,μνSA~(0)|Zc(p)=λ~Zεμναβεαpβ,0|J±,μνSA~(0)|Zc+(p)=λ~Z+(εμpνενpμ),0|J,μνAA/VV(0)|Zc+(p)=λ~Z+εμναβεαpβ,0|J,μνAA/VV(0)|Zc(p)=λ~Z(εμpνενpμ),0|J+,μνAA/VV(0)|Zc+(p)=λZ+εμν,

where λZ±=λ~Z±MZ±, the εμ/α and εμν are the polarization vectors. We choose the components Π+(p2) and p2Π~+(p2) to study the hidden-charm tetraquark states.

In Tab.4, we present the quark constituents and corresponding currents explicitly.

At the QCD side, we carry out the operator product expansion for the correlation functions Π(p), Πμν(p) and Πμναβ(p). For example, we contract the quark fields in the Πμν(p) with the Wick’s theorem, and obtain the results,

Πμν(p)=iεijkεimnεijkεimn2d4xeipx{Tr[γ5Sckk(x)γ5CUjjT(x)C]Tr[γνScnn(x)γμCDmmT(x)C]+Tr[γμSckk(x)γνCUjjT(x)C]

Tr[γ5Scnn(x)γ5CDmmT(x)C]Tr[γμSckk(x)γ5CUjjT(x)C]Tr[γνScnn(x)γ5CDmmT(x)C]Tr[γ5Sckk(x)γνCUjjT(x)C]Tr[γ5Scnn(x)γμCDmmT(x)C]},

where the full quark propagators, Sij(x)=Uij(x)=Dij(x),

Sij(x)=iδijx2π2x4δijmq4π2x2δijq¯q12+iδijxmqq¯q48δijx2q¯gsσGq192+iδijx2xmqq¯gsσGq1152igsGαβatija(xσαβ+σαβx)32π2x2iδijx2xgs2q¯q27776δijx4q¯qgs2GG2764818q¯jσμνqiσμν14q¯jγμqiγμ+,

and Scij(x)=SQij(x),

SQij(x)=i(2π)4d4keikx{δijkmQgsGαβntijn4σαβ(k+mQ)+(k+mQ)σαβ(k2mQ2)2+gsDαGβλntijn(fλβα+fλαβ)3(k2mQ2)4gs2(tatb)ijGαβaGμνb(fαβμν+fαμβν+fαμνβ)4(k2mQ2)5+},fλαβ=(k+mQ)γλ(k+mQ)γα(k+mQ)γβ(k+mQ),fαβμν=(k+mQ)γα(k+mQ)γβ(k+mQ)γμ(k+mQ)γν(k+mQ),

with Q=c, Dα=αigsGαntn [60, 381, 508], and the correspond to the currents J+,μSA(x) and J,μSA(x), respectively. In Eq. (129), we retain the terms q¯jσμνqi and q¯jγμqi come from the Fierz transformation of the qiq¯j to absorb the gluons emitted from the other quark lines to form q¯jgsGαβatmnaσμνqi and q¯jγμqiDνgsGαβatmna to extract the mixed condensate q¯gsσGq and four-quark condensate gs2q¯q2, respectively [60], where q=u, d or s. The condensate gs2q¯q2 comes from the q¯γμtaqgsDηGλτa, q¯jDμDνDαqi and q¯jDμDνDαqi rather than comes from the radiative O(αs) corrections to the q¯q2.

In fact, the method of adopting the full propagators in Eq. (128) implies vacuum saturation implicitly, the factorization of the higher dimensional vacuum condensates, for example, see Eq. (87), is already performed. We calculate the higher dimensional vacuum condensates using the formula tijatmna=16δijδmn+12δjmδin rigorously.

Let us see Eq. (128) again, there are two Q-quark propagators and two q-quark propagators, if each Q-quark line emits a gluon and each q-quark line contributes a quark−antiquark pair, we obtain an operator GμνGαβu¯ud¯d (or GμνGαβq¯qs¯s or GμνGαβs¯ss¯s), which is of dimension 10, see the Feynman diagram in Fig.7 for example. We should take account of the vacuum condensates at least up to dimension 10. The higher dimensional vacuum condensates are associated with the 1T2, 1T4 or 1T6, which manifest themselves at small Borel parameter T2 and play an important role in determining the Borel windows, where they play a minor important role. Therefore, we take account of the vacuum condensates q¯q, αsGGπ, q¯gsσGq, q¯q2, gs2q¯q2, q¯qαsGGπ, q¯qq¯gsσGq, q¯gsσGq2 and q¯q2αsGGπ, which are vacuum expectations of the quark-gluon operators of the order O(αsk) with k1. We truncate the operator product expansion in such a way consistently and rigorously. The condensates gs3GGG, αsGGπ2, αsGGπq¯gsσGq have the dimensions 6, 8, 9 respectively, but they are vacuum expectations of the quark-gluon operators of the order O(αs3/2), O(αs2), O(αs3/2) respectively, and discarded. Furthermore, direct calculations indicate such contributions are tiny indeed [509].

We accomplish the integrals in Eqs. (128)−(130) sequentially by simply setting d4xd4kd4qdDxdDkdDq with D=42ϵ to regularize the divergences, such a simple scheme misses many subtraction terms, which make no contribution in obtaining the imaginary parts through p2p2+iϵ. Finally, we obtain the QCD spectral densities ρQCD(s) through dispersion relation. Now we take a short digression to give an example,

Int(p2)=+dDxdDkdDqexp[i(p+q+k)x](q2mQ2)λ(k2mQ2)τx2n=i242nπ2Γ(n)+dDkdDqΓ(α)(q2mQ2)λ(k2mQ2)τ(p+q+k)2α=i242nπ2Γ(n)iπ2Γ(τ)01dx+dDqxα1(1x)τ1Γ(α+τD2)(q2mQ2)λ(x(1x)(p+q)2(1x)mQ2)α+τD2=i242nπ6Γ(λ)Γ(τ)Γ(n)01dxxα1(1x)τ1[x(1x)]α+τD201dyyα+τD21(1y)λ1[y(1y)]α+λ+τDΓ(α+λ+τD)(p2m~Q2)α+λ+τD,

with α=D2n. We set D=42ϵ and z=x(1y), then we obtain,

Int(p2)=i242nπ6Γ(λ)Γ(τ)Γ(n)yiyfdyy1λzi1ydzz1τ(1yz)n1Γ(λ+τn2+ϵ)(p2m~Q2)λ+τn2+ϵ.

In the case λ=τ=1 and n=2, we obtain,

Int(p2)=iπ6yiyfdyzi1ydz(1yz)Γ(ϵ2)(p2m~Q2)ϵ2,

and

1πImInt(s+iϵ)=iπ62yiyfdyzi1ydz(1yz)(sm~Q2)2,

where yf=1+14mQ2/s2, yi=114mQ2/s2, zi=ymQ2ysmQ2 and m~Q2=mQ2y+mQ2z.

Then we match the hadron side with the QCD side of the components Π+(p2) and p2Π~+(p2) of the correlation functions Π(p), Πμν(p) and Πμναβ(p) below the continuum thresholds s0 and perform Borel transformation with respect to P2=p2 to obtain the QCD sum rules:

λZ+2exp(MZ+2T2)=4mc2s0dsρQCD(s)exp(sT2).

We derive Eq. (135) with respect to τ=1T2, and obtain the QCD sum rules for the masses of the hidden-charm tetraquark states Zc or X,

MZ+2=4mc2s0dsddτρQCD(s)exp(τs)4mc2s0dsρQCD(s)exp(τs).

Now let us begin to perform numerical analysis, and write down the energy-scale dependence of the input parameters,

q¯q(μ)=q¯q(1GeV)[αs(1GeV)αs(μ)]12332nf,s¯s(μ)=s¯s(1GeV)[αs(1GeV)αs(μ)]12332nf,q¯gsσGq(μ)=q¯gsσGq(1GeV)[αs(1GeV)αs(μ)]2332nf,s¯gsσGs(μ)=s¯gsσGs(1GeV)[αs(1GeV)αs(μ)]2332nf,

mc(μ)=mc(mc)[αs(μ)αs(mc)]12332nf,ms(μ)=ms(2GeV)[αs(μ)αs(2GeV)]12332nf,αs(μ)=1b0t[1b1b02logtt+b12(log2tlogt1)+b0b2b04t2],

where t=logμ2ΛQCD2, b0=332nf12π, b1=15319nf24π2, b2=285750339nf+32527nf2128π3, ΛQCD=210MeV, 292MeV and 332MeV for the flavors nf=5, 4 and 3, respectively [510, 511]. Because the c-quark is concerned, we choose the flavor number to be nf=4.

At the beginning points, we adopt the commonly-used values q¯q=(0.24±0.01GeV)3, s¯s=(0.8±0.1)q¯q, q¯gsσGq=m02q¯q, s¯gsσGs=m02s¯s, m02=(0.8±0.1)GeV2, αsGGπ=(0.012±0.004)GeV4 at the typical energy scale μ=1GeV [376, 377, 381, 385], and adopt the MS¯ (modified-minimal-subtraction) masses mc(mc)=(1.275±0.025)GeV and ms(2GeV)=(0.095±0.005)GeV from the Particle Data Group [510].

We apply the modified energy scale formula μ=MX/Y/Z2(2Mc)2κMs to choose the suitable energy scales of the QCD spectral densities [61, 172, 425], where the Mc and Ms are the effective c and s-quark masses respectively, and have universal values to be commonly used elsewhere. We adopt the updated value Mc=1.82GeV [512], and take the collective light-flavor SU(3)-breaking effects into account by introducing an effective s-quark mass Ms=0.20GeV (0.12GeV) for the S-wave (P-wave) tetraquark states [82, 83, 171, 172, 418, 425, 513-515].

The continuum threshold parameters are not completely free parameters, and cannot be determined by the QCD sum rules themselves completely. We often consult the experimental data in choosing the continuum threshold parameters. The Zc(4430) can be assigned to be the first radial excitation of the Zc(3900) according to the analogous decays,

Zc±(3900)J/ψπ±,Zc±(4430)ψπ±,

and the analogous mass gaps MZc(4430)MZc(3900)=591MeV and MψMJ/ψ=589MeV from the Particle Data Group [56, 181, 182, 411].

We tentatively choose the continuum threshold parameters as s0=MX/Z+0.60GeV and vary the continuum threshold parameters s0 and Borel parameters T2 to satisfy the following four criteria:

● Pole dominance at the hadron side;

● Convergence of the operator product expansion;

● Appearance of the Borel platforms;

● Fulfillment of the modified energy scale formula,

via trial and error.

Thereafter, such criteria are adopted for all the hidden-charm (hidden-bottom) tetraquark (molecular) states, hidden-charm (hidden-bottom) pentaquark (molecular) states, doubly-charm (doubly-bottom) tetraquark (molecular) states, doubly-charm (doubly-bottom) pentaquark (molecular) states, etc.

The pole dominance at the hadron side and convergence of the operator product expansion at the QCD side are two basic criteria, we should satisfy them to obtain reliable QCD sum rules. In the QCD sum rules for the hidden-charm tetraquark and pentaquark states, the largest power of the energy variable s in the QCD spectral densities ρQCD(s)s4 and s5, respectively, which make the integrals,

4mc2s0dss4exp(sT2),4mc2s0dss5exp(sT2),

converge more slowly compared to the traditional hadrons, it is very difficult to satisfy the two basic criteria at the same time. We define the pole contributions (PCs) by

PC=4mc2s0dsρQCD(s)exp(sT2)4mc2dsρQCD(s)exp(sT2),

while we define the contributions of the vacuum condensates D(n) of dimension n by

D(n)=4mc2s0dsρQCD,n(s)exp(sT2)4mc2s0dsρQCD(s)exp(sT2),

or

D(n)=4mc2dsρQCD,n(s)exp(sT2)4mc2dsρQCD(s)exp(sT2),

sometimes we would like to use the notation Dn in stead of D(n). Compared to the criterion in Eq. (142), the criterion in Eq. (141) is strong and leads to larger Borel parameter T2. If we only study the ground state contributions, Eq. (141) is preferred.

After trial and error, we obtain the Borel windows, continuum threshold parameters, energy scales of the QCD spectral densities, pole contributions, and contributions of the vacuum condensates of dimension 10, which are shown explicitly in Tab.5 and Tab.6.

At the hadron side, the pole contributions are about (40−60)%, the pole dominance is well satisfied. It is more easy to obtain the pole contributions (40−60)% with the help of the modified energy scale formula. In Tab.6, we provide two sets of data for the scalar tetraquark states, one set data is based on the continuum threshold parameters about s0=MX+(0.60±0.10)GeV with the central values s0<MY+0.60GeV (this criterion is also adopted for the tetraquark states with the JPC=1+, 1++ and 2++), the other set data is based on the continuum threshold parameters about s0=MX+(0.55±0.10)GeV with the central values s0<MY+0.55GeV, which will be characterized by the additional symbol in Tab.6, Tab.8 and Tab.10.

At the QCD side, the contributions of the vacuum condensates of dimension 10 are |D(10)|1% or 1% except for |D(10)|<2% for the [uc]V~[dc¯]V[uc]V[dc¯]V~ tetraquark state with the JPC=1++, the convergent behavior of the operator product expansion is very good.

In Fig.16, we plot the contributions of the vacuum condensates D(n) with variation of the Borel parameter T2 for the current J,μSA(x). From the figure, we can see clearly that the main contributions come from the D(0), D(3), D(5), D(6) and D(8), while the largest contribution comes from the D(3). Therefore, if we would like to calculate the radiative O(αs) corrections, we should calculate the radiative O(αs) corrections to the D(3), not just to the D(0). At the present time, only the radiative O(αs) corrections to the D(0) are partially calculated [128, 452, 456, 461, 516-519].

We take account of all the uncertainties of the input parameters and obtain the masses and pole residues of the scalar, axialvector and tensor hidden-charm tetraquark states, which are shown explicitly in Tab.7 and Tab.8. From Tab.5–Tab.8, we could see clearly that the modified energy scale formula μ=MX/Y/Z2(2Mc)2κMs is well satisfied. In Fig.17, we plot the masses of the [uc]S[dc¯]A[uc]A[dc¯]S and [uc]S[dc¯]A+[uc]A[dc¯]S tetraquark states having the spin-parity JP=1+ with variations of the Borel parameters at much larger ranges than the Borel widows as an example, there appear platforms in the Borel windows indeed.

In Tab.9 and Tab.10, we present the possible assignments of the ground state hidden-charm tetraquark states, and revisit the assignments based on the tetraquark scenario, there are rooms for the X(3860), X(3872), X(3915), X(3960), X(4140), X(4274), X(4500), X(4685), X(4700), Zc(3900), Zc(4020), Zc(4050), Zc(4055), Zc(4430) and Zc(4600).

It is not difficult to reproduce the mass of the X(3872) in the scenario of tetraquark state [58, 60, 61], it is a great challenge to reproduce the tiny width (1.19±0.21)MeV from the Particle Data Group consistently [97]. In Ref. [62], we study the strong decays X(3872)J/ψπ+π, J/ψω, χc1π0, D0D¯0 and D0D¯0π0 via the QCD sum rules according to rigorous quark-hadron duality, and obtain the total decay width about 1MeV, it is the first time to reproduce the tiny width of the X(3872) via the QCD sum rules. The thresholds of the J/ψρ and J/ψω are 3872.16MeV and 3879.56MeV respectively, which are larger than the value 3871.64MeV of the mass of the X(3872) from the Particle Data Group [97] and also lead to the possibility that it may be a J/ψρ or J/ψω molecular state [84], the decays X(3872)J/ψπ+π and J/ψπ+ππ take place through the virtual ρ and ω mesons, respectively, we introduce form-factors to parameterize the off-shell-ness, however, the arbitrariness in choosing the form-factors would impair the predictive ability.

As far as the Zc(3900) is concerned, it is not difficult to reproduce its mass and width via the QCD sum rules [60, 61, 520]. It is the benchmark for the 3¯3 type hidden-charm tetraquark states in the simple diquark models [56, 301, 312, 313, 521].

In Tab.9, there are enough rooms to accommodate the hc(4000) and χc1(4010) in the scenario of tetraquark states, as the central values of the predicted tetraquark masses happen to coincide with the experimental data from the LHCb Collaboration [109]. We should bear in mind that the predictions were made before the experimental observation, therefore the calculations are robust enough [61]. While in the traditional charmonium scenario, the hc and χc1 have the masses 3956MeV and 3953MeV, respectively [108], there exist discrepancies about 5060MeV.

The predictions MX=(4.11±0.09)GeV and (4.17±0.09)GeV for the 1S [sc]S[sc¯]A+[sc]A[sc¯]S and [sc]S[sc¯]A~+[sc]A~[sc¯]S tetraquark states with the JPC=1++ respectively support assigning the X(4140) and X(4685) as the 1S and 2S [sc]S[sc¯]A+[sc]A[sc¯]S or [sc]S[sc¯]A~+[sc]A~[sc¯]S tetraquark states respectively. In Refs. [507, 522], we obtain the prediction MX=4.14±0.10GeV for the 1S [sc]V~[sc¯]V[sc]V[sc¯]V~ tetraquark state with the JPC=1++ in the old scheme, which supports assigning the X(4140) as the [sc]V~[sc¯]V[sc]V[sc¯]V~ tetraquark state.

The prediction MX=4.29±0.09GeV for the 1S [sc]V~[sc¯]V[sc]V[sc¯]V~ tetraquark state supports assigning the X(4274) as the [sc]V~[sc¯]V[sc]V[sc¯]V~ tetraquark state with the JPC=1++. The calculations in Refs. [507, 522] are updated, as where the light-flavor SU(3) mass-breaking effects in the energy scale formula are not taken into account.

In Ref. [523], the predictions support assigning the X(4274) as the [sc]A[s¯c¯]V[sc]V[s¯c¯]A tetraquark state with a relative P-wave between the diquark and antidiquark pairs, such an assignment does not suffer from shortcomings in sense of treating scheme. The X(4274) maybe have two important Fock components.

The X(4140) and X(4274) are also play an important role in establishing the hidden-charm tetraquark states, especially the X(4140). There have been several possible assignments of the X(4140), such as the tetraquark state [315, 316, 332, 333, 425, 507, 522, 524-527], hybrid state [477, 478, 528] or rescattering effect [529].

In Tab.10, we assign the X(3960) and X(4500) as the 1S and 2S [cs]S[c¯s¯]S tetraquark states with the JPC=0++, the X(4700) as the 1S [cs]V[c¯s¯]V tetraquark state with the JPC=0++; or assign the X(4700) as the 2S [cs]A[c¯s¯]A or [cs]A~[c¯s¯]A~ tetraquark state with the JPC=0++. In Ref. [530], we assign the X(4500) as the [sc]V^[sc¯]V^ tetraquark state with the JPC=0++, where the V^ denotes the diquark with an explicit P-wave. In Ref. [527], Chen et al. assigned the X(4500) and X(4700) as the D-wave tetraquark states with the quark content csc¯s¯ and JP=0+: the X(4500) consists of one D-wave diquark and one S-wave antidiquark, with the antisymmetric color structure 3¯3; the X(4700) consists of similar diquarks but with the symmetric color structure 66¯.

However, there is no room for the X(4350), we should introduce mixing effect to interpret its nature [484].

In fact, the predictions of the QCD sum rules have arbitrariness, and depend on the interpolating currents, truncations of the operator product expansion, convergent behavior, pole contributions, input parameters, etc. Therefore, the predictions maybe quite different, we have to perform systematic investigations with uniform criterion to outcome the shortcomings.

We suggest to study the two-body strong decays to diagnose those hidden-charm tetraquark states [61, 425], for example,

Zc±(1+)π±J/ψ,π±ψ,π±hc,ρ±ηc,(DD¯)±,(DD¯)±,(DD¯)±,Zc±(0++)π±ηc,π±χc1,ρ±J/ψ,ρ±ψ,   (DD¯)±,(DD¯)±,

Zc±(1++)π±χc1,ρ±J/ψ,ρ±ψ,(DD¯)±,(DD¯)±,(DD¯)±,Zc±(2++)π±ηc,π±χc1,ρ±J/ψ,ρ±ψ,(DD¯)±,(DD¯)±,

Zc0(1+)π0J/ψ,π0ψ,π0hc,ρ0ηc,(DD¯)0,(DD¯)0,(DD¯)0,Zc0(0++)π0ηc,π0χc1,ρ0J/ψ,ρ0ψ,(DD¯)0,(DD¯)0,Zc0(1++)π0χc1,ρ0J/ψ,ρ0ψ,(DD¯)0,(DD¯)0,(DD¯)0,Zc0(2++)π0ηc,π0χc1,ρ0J/ψ,ρ0ψ,(DD¯)0,(DD¯)0,

X(1+)ηJ/ψ,ηψ,ηhc,ωηc,(DD¯)0,(DD¯)0,(DD¯)0,X(0++)ηηc,ηχc1,ωJ/ψ,ωψ,(DD¯)0,(DD¯)0,

X(1++)ηχc1,ωJ/ψ,ωψ,(DD¯)0,(DD¯)0,(DD¯)0,X(2++)ηηc,ηχc1,ωJ/ψ,ωψ,(DD¯)0,(DD¯)0,

X(1+)ηJ/ψ,ηψ,ηhc,ϕηc,DsD¯s,DsD¯s,DsD¯s,X(0++)ηηc,ηχc1,ϕJ/ψ,ϕψ,DsD¯s,DsD¯s,X(1++)ηχc1,ϕJ/ψ,ϕψ,DsD¯s,DsD¯s,DsD¯s,X(2++)ηηc,ηχc1,ϕJ/ψ,ϕψ,DsD¯s,DsD¯s.

The LHCb Collaboration observed the hc(4000) and χc1(4010) in the D±D mass spectrum [109], see the two-body strong decays of the Zc0 shown in Eq. (144).

As the cc¯qs¯ tetraquark states are concerned, we present the predictions based on the direct calculations plus light-flavor SU(3)-breaking effects in Tab.11 [171, 172]. In Ref. [171], we tentatively assign the Zc(4020/4025) as the AA¯-type hidden-charm tetraquark state with the JPC=1+ according to the analogous properties of the Zc(3900/3885) and Zcs(3985/4000), and study the AA¯-type tetraquark states without strange, with strange and with hidden-strange via the QCD sum rules in a consistent way. Then we study the hadronic coupling constants of the tetraquark states without strange and with strange via the QCD sum rules based on rigorous quark-hadron duality, and obtain the total decay widths

,

ΓZcs=22.71±1.65(or±6.60)MeV,ΓZc=29.57±2.30(or±9.20)MeV,

and suggest to search for the Zcs state in the mass spectrum of the hcK, J/ψK, ηcK, DD¯s, DsD¯. Slightly later, the BESIII Collaboration observed an evidence for the Zcs(4123) [170].

With the simple replacement,

cb,

we obtain the corresponding QCD sum rules for the hidden-bottom tetraquark states. And we would like to present the results from the QCD sum rules in Ref. [531] directly, see Tab.12 and Tab.13.

In calculations, we use the energy scale formula μ=MX/Y/Z2(2Mb)2 with the effective b-quark mass Mb=5.17GeV to determine the ideal energy scales of the QCD spectral densities [532], and choose the continuum threshold parameters s0=Zb+0.55±0.10GeV as a constraint to extract the masses and pole residues from the QCD sum rules. The predicted masses 10.61±0.09GeV and 10.62±0.09GeV for the 1+ tetraquark states support assigning the Zb(10610) and Zb(10650) to be the hidden-bottom tetraquark states with the JPC=1+, more theoretical and experimental works are still needed to assign the Zb(10610) and Zb(10650) unambiguously according to the partial decay widths.

In the diquark-model, the Zb±(10610) and Zb±(10650) are also assigned as the SA¯AS¯-type and AA¯-type hidden-bottom tetraquark states, respectively [533, 534]. And we could extend this subsection to study the Bc-like tetraquark states [458, 535, 536, 537].

3.1.2 Tetraquark states with the first radial excitations

In this sub-section, we would like to study the first radial excitations of the hidden-charm tetraquark states and make possible assignments of the exotic states. In Ref. [538], Maior de Sousa and Rodrigues da Silva suggested a theoretical scheme to study the double-pole QCD sum rules, and study the quarkonia ρ(1S,2S), ψ(1S,2S), Υ(1S,2S) and ψt(1S,2S), the predicted hadron masses are not good enough, they adopt the experimental masses except for the ψt(1S,2S) to study the decay constants. In Ref. [182], we extend this scheme to study the Zc(3900) and Zc(4430) as the ground state and its first radial excitation, respectively, and observed that it is impossible to reproduce the experimental masses at the same energy scale, just as in the case of the ρ(1S,2S), ψ(1S,2S) and Υ(1S,2S) [538], we should resort to the energy scale formula, see Eq. (99), to choose the optimal energy scales independently.

We adopt the correlation functions Πμν(p) and Πμναβ(p) [184], see Eq. (125), and set Jμ(x)=Jμ1(x), Jμ2(x), Jμ3(x), and Jμ1(x)=J,μSA(x), Jμ2(x)=J,μA~A(x), Jμ3(x)=J,μV~V(x) and Jμν(x)=J,μνAA(x), see Eq. (116).

At the hadron side, if we only take account of the ground state hidden-charm tetraquark states, we obtain the QCD sum rules:

λZ2exp(MZ2T2)=4mc2s0dsρQCD(s)exp(sT2),

MZ2=4mc2s0dsddτρQCD(s)eτs4mc2s0dsρQCD(s)eτsτ=1T2.

Thereafter, we will refer the QCD sum rules in Eq. (149) and Eq. (150) as QCDSR I.

If we take into account the contributions of the first radially excited tetraquark states Zc in the hadronic representation, we can obtain the QCD sum rules,

λZ2exp(MZ2T2)+λZ2exp(MZ2T2)=4mc2s0dsρQCD(s)exp(sT2),

where the s0 is continuum threshold parameter, then we introduce the notations τ=1T2, Dn=(ddτ)n, and resort to the subscripts 1 and 2 to represent the ground state Zc and first radially excited state Zc respectively for simplicity. We rewrite the QCD sum rules as

λ12exp(τM12)+λ22exp(τM22)=ΠQCD(τ),

here we introduce the subscript QCD to represent the QCD representation. We derive the QCD sum rules in Eq. (152) in regard to τ to obtain

λ12M12exp(τM12)+λ22M22exp(τM22)=DΠQCD(τ).

From Eqs. (152) and (153), we obtain the QCD sum rules,

λi2exp(τMi2)=(DMj2)ΠQCD(τ)Mi2Mj2,

where the indexes ij. Let us derive the QCD sum rules in Eq. (154) in regard to τ to obtain

Mi2=(D2Mj2D)ΠQCD(τ)(DMj2)ΠQCD(τ),Mi4=(D3Mj2D2)ΠQCD(τ)(DMj2)ΠQCD(τ).

The squared masses Mi2 satisfy the equation,

Mi4bMi2+c=0,

where

b=D3D0D2DD2D0DD,c=D3DD2D2D2D0DD,DjDk=DjΠQCD(τ)DkΠQCD(τ),

the indexes i=1,2 and j,k=0,1,2,3. Finally we solve above equation analytically to obtain two solutions [182, 538],

M12=bb24c2,

M22=b+b24c2.

Thereafter, we will denote the QCD sum rules in Eq. (151) and Eqs. (158) and (159) as QCDSR II. In calculations, we observe that if we specify the energy scales of the spectral densities in the QCD representation, only one solution satisfies the energy scale formula μ=MX/Y/Z2(2Mc)2 in the QCDSR II, we have to abandon the other solution, i.e., the mass M1 (MZ). It is the unique feature of our works [182, 184], which is in contrast to Ref. [538].

The Okubo−Zweig−Iizuka supper-allowed decays,

ZcJ/ψπ,Zcψπ,Zcψπ,

are expected to take place easily. The energy gaps maybe have the relations MZMZ=mψmJ/ψ and MZMZ=mψmψ. The charmonium masses are mJ/ψ=3.0969GeV, mψ=3.686097GeV and mψ=4.039GeV from the Particle Data Group [411], mψmJ/ψ=0.59GeV, mψmJ/ψ=0.94GeV, we can choose the continuum threshold parameters to be s0=MZ+0.59GeV and s0=MZ+0.95GeV tentatively and vary the continuum threshold parameters and Borel parameters to satisfy the four criteria in Section 3.1.1.

After trial and error, we obtain the continuum threshold parameters, Borel windows, best energy scales, and contributions of the ground states for the QCDSR I, see Tab.14. In Tab.14, we write the continuum threshold parameters as s0=21.0±1.0GeV2 rather than as s0=(4.58±0.11GeV)2 for the [uc]A~[d¯c¯]A[uc]A[d¯c¯]A~ and [uc]A[d¯c¯]A tetraquark states to remain the same form as in Ref. [512]. Again we obtain the parameters for the QCDSR II using trial and error, see Tab.15. We take the energy scale formula μ=MX/Y/Z2(2Mc)2 to obtain the ideal energy scales of the spectral densities [182, 184].

From Tab.14 and Tab.15, we can see clearly that the contributions of the single-pole terms are about (40−60)% for the QCDSR I, the contributions of the two-pole terms are about (70−80)% for the QCDSR II, which satisfy the pole dominance very well. In the QCDSR II, the contributions of the ground states are about (30−45)%, which are much less than the ground state contributions in the QCDSR I, we prefer the QCDSR I for the ground states.

We take account of all the uncertainties of the parameters, and obtain the masses and pole residues, which are shown in Tab.16 and Tab.17. From those Tables, we can see that the ground state tetraquark masses from the QCDSR I and the radially excited tetraquark masses from the QCDSR II satisfy the energy scale formula μ=MX/Y/Z2(2Mc)2, where the updated value of the effective c-quark mass Mc=1.82GeV is adopted [512]. In Tab.17, we also present the central values of the ground state masses and pole residues extracted from the QCDSR II at the same energy scales. If we examine Tab.17, we would observe the ground state masses cannot satisfy the energy scale formula, and should be discarded.

For example, in Fig.18, we plot the ground state masses from the QCDSR I and the first radially excited tetraquark masses from the QCDSR II with variations of the Borel parameters for the [uc]S[d¯c¯]A[uc]A[d¯c¯]S and [uc]A~[d¯c¯]A[uc]A[d¯c¯]A~ states. From the figure, we observe that there indeed appear very flat platforms in the Borel windows.

We present the possible assignments explicitly in Tab.9. The predicted mass MZ=4.47±0.09GeV for the 2S [uc]S[d¯c¯]A[uc]A[d¯c¯]S tetraquark state exhibits very good agreement with the experimental data 4475±725+15MeV from the LHCb Collaboration [179], which is in favor of assigning the Zc(4430) as the first radial excitation of the [uc]S[d¯c¯]A[uc]A[d¯c¯]S tetraquark state with the JPC=1+ [182, 184].

The predicted mass MZ=4.60±0.09GeV for the 2S [uc]A~[d¯c¯]A[uc]A[d¯c¯]A~ tetraquark state and MZ=4.58±0.09GeV for the 2S [uc]A[d¯c¯]A tetraquark state both exhibit very good agreement with the experimental data 4600MeV from the LHCb Collaboration [183], and the predicted mass MZ=4.66±0.10GeV for the 1S [uc]V~[d¯c¯]V+[uc]V[d¯c¯]V~ tetraquark state is also compatible with the experimental data. In summary, there are three tetraquark state candidates with the JPC=1+ for the Zc(4600).

The scheme of the QCDSR II and its modification have been applied extensively to study the 2S tetraquark states, such as the Zc(4430) [182, 184, 185, 464, 539], Zc(4600) [184, 185], X(4500) (as the 2S state of the X(3915)) [540, 541], X(4685) (as the 2S state of the X(4140)) [522], and Λc(2S) and Ξc(2S) [542].

3.1.3 Tetraquark states with negative parity

In Section 3.1.1, we study the hidden-heavy tetraquark states with the positive parity, in this subsection, we would like to study the hidden-heavy tetraquark states with the negative parity. Again we choose the diquark operators without explicit P-waves as the elementary building blocks, for detailed discussions about the diquark operators, see the beginning of Section 3.1.1. Compared to the vector tetraquark states, it is easy to analyze the pseudoscalar tetraquark states, and we would like to present the results directly.

Again, let us adopt the correlation functions Πμν(p) and Πμναβ(p) defined in Eq. (125), and write down the currents

Jμ(x)=J,μPA(x),J+,μPA(x),J,μSV(x),J+,μSV(x),J,μV~A(x),J+,μV~A(x),J,μA~V(x),J+,μA~V(x),Jμν(x)=J,μνSV~(x),J+,μνSV~(x),J,μνPA~(x),J+,μνPA~(x),J,μνAA(x),

J,μPA(x)=εijkεimn2[ujT(x)Cck(x)d¯m(x)γμCc¯nT(x)ujT(x)Cγμck(x)d¯m(x)Cc¯nT(x)],J+,μPA(x)=εijkεimn2[ujT(x)Cck(x)d¯m(x)γμCc¯nT(x)+ujT(x)Cγμck(x)d¯m(x)Cc¯nT(x)],J,μSV(x)=εijkεimn2[ujT(x)Cγ5ck(x)d¯m(x)γ5γμCc¯nT(x)+ujT(x)Cγμγ5ck(x)d¯m(x)γ5Cc¯nT(x)],

J+,μSV(x)=εijkεimn2[ujT(x)Cγ5ck(x)d¯m(x)γ5γμCc¯nT(x)ujT(x)Cγμγ5ck(x)d¯m(x)γ5Cc¯nT(x)],

J,μV~A(x)=εijkεimn2[ujT(x)Cσμνck(x)d¯m(x)γνCc¯nT(x)ujT(x)Cγνck(x)d¯m(x)σμνCc¯nT(x)],J+,μV~A(x)=εijkεimn2[ujT(x)Cσμνck(x)d¯m(x)γνCc¯nT(x)+ujT(x)Cγνck(x)d¯m(x)σμνCc¯nT(x)],J,μA~V(x)=εijkεimn2[ujT(x)Cσμνγ5ck(x)d¯m(x)γ5γνCc¯nT(x)+ujT(x)Cγνγ5ck(x)d¯m(x)γ5σμνCc¯nT(x)],J+,μA~V(x)=εijkεimn2[ujT(x)Cσμνγ5ck(x)d¯m(x)γ5γνCc¯nT(x)ujT(x)Cγνγ5ck(x)d¯m(x)γ5σμνCc¯nT(x)],

J,μνSV~(x)=εijkεimn2[ujT(x)Cγ5ck(x)d¯m(x)σμνCc¯nT(x)ujT(x)Cσμνck(x)d¯m(x)γ5Cc¯nT(x)],J+,μνSV~(x)=εijkεimn2[ujT(x)Cγ5ck(x)d¯m(x)σμνCc¯nT(x)+ujT(x)Cσμνck(x)d¯m(x)γ5Cc¯nT(x)],J,μνPA~(x)=εijkεimn2[ujT(x)Cck(x)d¯m(x)γ5σμνCc¯nT(x)ujT(x)Cσμνγ5ck(x)d¯m(x)Cc¯nT(x)],J+,μνPA~(x)=εijkεimn2[ujT(x)Cck(x)d¯m(x)γ5σμνCc¯nT(x)+ujT(x)Cσμνγ5ck(x)d¯m(x)Cc¯nT(x)],

J,μνAA(x)=εijkεimn2[ujT(x)Cγμck(x)d¯m(x)γνCc¯nT(x)ujT(x)Cγνck(x)d¯m(x)γμCc¯nT(x)],

the subscripts ± denote the positive and negative charge conjugation, respectively, the superscripts P, S, A(A~) and V(V~) denote the pseudoscalar, scalar, axialvector and vector diquark and antidiquark operators, respectively [424, 505, 512, 543, 544]. With the simple replacement,

us,d¯s¯,

we reach the corresponding currents for the cc¯ss¯ states [424, 544, 545].

Under parity transformation P^, the current operators Jμ(x) and Jμν(x) have the properties,

P^Jμ(x)P^1=+Jμ(x~),P^J~μν(x)P^1=J~μν(x~),P^J,μνAA(x)P^1=+JAA,μν(x~),

according to the properties of the diquark operators,

P^εijkqjT(x)CΓck(x)P^1=εijkqjT(x~)Cγ0Γγ0ck(x~),P^εijkq¯j(x)ΓCc¯kT(x)P^1=εijkq¯j(x~)γ0Γγ0Cc¯kT(x~),

where J~μν(x)=J,μνSV~(x), J+,μνSV~(x), J,μνPA~(x), J+,μνPA~(x), xμ=(t,x) and x~μ=(t,x). For Γ=1, γ5, γμ, γμγ5, σμν, σμνγ5, we obtain γ0Γγ0=1, γ5, γμ, γμγ5, σμν, σμνγ5. And we rewrite Eq. (167) in more explicit form,

P^Ji(x)P^1=Ji(x~),P^J~ij(x)P^1=J~ij(x~),P^J,0iAA(x)P^1=JAA,0i(x~),

P^J0(x)P^1=+J0(x~),P^J~0i(x)P^1=+J~0i(x~),P^J,ijAA(x)P^1=+JAA,ij(x~),

where i, j=1, 2, 3.

Under charge conjugation transformation C^, the currents Jμ(x) and Jμν(x) have the properties,

C^J±,μ(x)C^1=±J±,μ(x)ud,C^J±,μν(x)C^1=±J±,μν(x)ud.

The currents Jμ(x) and Jμν(x) have the symbolic quark structure c¯cd¯u with the isospin I=1 and I3=1. In the isospin limit, the currents with the symbolic quark structures

c¯cd¯u,c¯cu¯d,c¯cu¯ud¯d2,c¯cu¯u+d¯d2

couple potentially to the hidden-charm tetraquark states with degenerated masses, and the currents with the isospin I=1 and 0 lead to the same QCD sum rules. Only the currents with the symbolic quark structures c¯cu¯ud¯d2 and c¯cu¯u+d¯d2 have definite charge conjugation, again we assume that the tetraquark states c¯cd¯u have the same charge conjugation as their neutral partners.

According to the current-hadron duality, we obtain the hadronic representation,

Πμν(p)=λY2MY2p2(gμν+pμpνp2)+=Π(p2)(gμν+pμpνp2)+,Π~μναβ(p)=λ~Y2MY2p2(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+λ~Z+2MZ+2p2(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+=Π~(p2)(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+Π~+(p2)(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ),ΠμναβAA(p)=λ~Z+2MZ+2p2(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+λ~Y2MY2p2(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+=Π~+(p2)(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+Π~(p2)(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ),

where the pole residues are defined by

0|Jμ(0)|Yc(p)=λYεμ,0|J~μν(0)|Yc(p)=λ~Yεμναβεαpβ,0|J~μν(0)|Zc+(p)=λ~Z+(εμpνενpμ),0|J,μνAA(0)|Zc+(p)=λ~Z+εμναβεαpβ,0|J,μνAA(0)|Yc(p)=λ~Y(εμpνενpμ),

λZ+=λ~Z+MZ+, λY=λ~YMY, the εμ/α are the polarization vectors, we add the superscripts/subscripts ± to denote the positive and negative parity, respectively, and add the wide-tilde and superscript in the Πμναβ(p) to denote the currents J~μν(x) and J,μνAA(x), respectively. We choose the components Π(p2) and p2Π~(p2) to explore the negative parity hidden-charm tetraquark states [543, 545].

At the QCD side, we calculate the vacuum condensates up to dimension 10 and take into account the vacuum condensates q¯q, αsGGπ, q¯gsσGq, q¯q2, gs2q¯q2, q¯qαsGGπ, q¯qq¯gsσGq, q¯gsσGq2 and q¯q2αsGGπ, which are vacuum expectations of the quark-gluon operators of the order O(αsk) with k1 [543, 545].

We match the hadron side with the QCD side of the components Π(p2) and p2Π~(p2) below the continuum thresholds s0 with the help of the spectral representation, and perform Borel transformation with respect to P2=p2 to obtain the QCD sum rules:

λY2exp(MY2T2)=4mc2s0dsρQCD(s)exp(sT2),

the explicit expressions of the QCD spectral densities ρQCD(s) are neglected for simplicity.

We derive Eq. (175) with respect to τ=1T2, and obtain the QCD sum rules for the masses of the vector hidden-charm tetraquark states Yc,

MY2=4mc2s0dsddτρQCD(s)exp(τs)4mc2s0dsρQCD(s)exp(τs).

With a simple replacement cb, we obtain the corresponding QCD sum rules for the hidden-bottom tetraquark states directly.

We perform the same procedure as in the previous subsections, and obtain the Borel windows, continuum threshold parameters, suitable energy scales of the QCD spectral densities and pole contributions, which are shown explicitly in Tab.18 and Tab.19. From the Tables, we can see clearly that the pole contributions are about (40−60)%, the central values are larger than 50%, the pole dominance is well satisfied. In calculations, we observe that the main contributions come from the perturbative terms, the higher dimensional condensates play a minor important role and the contributions |D(10)|1%, the operator product expansion converges very good.

We take account of all the uncertainties of the relevant parameters and obtain the masses and pole residues of the hidden-charm (and hidden-strange) tetraquark states with the JPC=1 and 1+, which are shown explicitly in Tab.20 and Tab.21.

In Tab.22 and Tab.23, we present the possible assignments of the hidden-charm tetraquark states with the JPC=1 and 1+ obtained in Refs. [543, 545]. Considering the large uncertainties, it is possible to assign the X(4630) as the [sc]S[sc¯]V~+[sc]V~[sc¯]S state with the JPC=1+, which has a mass 4.68±0.09GeV, see Tab.23. In Ref. [409], we prove that it is feasible and reliable to study the multiquark states in the framework of the QCD sum rules, and obtain the prediction for the mass of the DsD¯s1Ds1D¯s molecular state with the exotic quantum numbers JPC=1+, MX=4.67±0.08GeV, which was obtained before the LHCb data and is compatible with the LHCb data. The X(4630) maybe have two important Fock components.

After Ref. [543] was published, the Y(4500) was observed by the BESIII Collaboration [155, 156, 159]. At the energy about 4.5GeV, we obtain three hidden-charm tetraquark states with the JPC=1, the [uc]V~[uc¯]A+[dc]V~[dc¯]A[uc]A[uc¯]V~[dc]A[dc¯]V~, [uc]A~[uc¯]V+[dc]A~[dc¯]V+[uc]V[uc¯]A~+[dc]V[dc¯]A~ and [uc]S[uc¯]V~+[dc]S[dc¯]V~[uc]V~[uc¯]S[dc]V~[dc¯]S tetraquark states have the masses 4.53±0.07GeV, 4.48±0.08GeV and 4.50±0.09GeV, respectively [543]. In Ref. [546], we study the two-body strong decays systematically, i.e., we obtain thirty QCD sum rules for the hadronic coupling constants based on rigorous quark−hadron duality, then obtain the partial decay widths, therefore the total widths approximately, which are compatible with the experimental data of the Y(4500) from the BESIII Collaboration, see Section 7.1 for details. In Ref. [547], we take the Y(4500) as the [uc]A~[uc¯]V+[dc]A~[dc¯]V+[uc]V[uc¯]A~+[dc]V[dc¯]A~ tetraquark state, and study the three-body strong decay Y(4500)DD0π+ with the light-cone QCD sum rules, see Section 7.2 for details.

If only the mass is concerned, the Y(4660) can be assigned as the [sc]A~[sc¯]V+[sc]V[sc¯]A~, [sc]S[sc¯]V~[sc]V~[sc¯]S, [uc]P[uc¯]A+[dc]P[dc¯]A[uc]A[uc¯]P[dc]A[dc¯]P or [uc]A[uc¯]A+[dc]A[dc¯]A tetraquark state, see Tab.22-Tab.23. In other words, the Y(4660) maybe have several important Fock components, we have to study the strong decays in details to diagnose its nature. For example, if we assign the Y(4660) as the [sc]A~[sc¯]V+[sc]V[sc¯]A~ or [sc]S[sc¯]V~[sc]V~[sc¯]S state, then the strong decays Y(4660)J/ψf0(980), ηcϕ, χc0ϕ, DsD¯s, DsD¯s, DsD¯s, DsD¯s, J/ψπ+π and ψπ+π are Okubo-Zweig-Iizuka super-allowed, considering the intermediate process f0(980)π+π. Up to now, only the decays Y(4660)J/ψπ+π, ψ2(3823)π+π, Λc+Λc and Ds+Ds1 have been observed [97], which cannot exclude the assignments Y(4660)=[uc]P[uc¯]A+[dc]P[dc¯]A[uc]A[uc¯]P[dc]A[dc¯]P or [uc]A[uc¯]A+[dc]A[dc¯]A, as the decay Y(4660)Ds+Ds1 can take place through the re-scattering mechanism. We can investigate or search for the neutral Yc tetraquark states with the JPC=1 and 1+ through the two-body or three-body strong decays,

Yc(1)χc0ρ/ω,J/ψπ+π,J/ψKK¯,ηcρ/ω,χc1ρ/ω,Yc(1+)J/ψρ/ω,hcρ/ω.

From Tab.22 and Tab.23, we observe that there is no room for the Y(4260/4220). In Ref. [544], we choose the currents,

Jμ1(x)=J,μPA(x)ud¯ss¯,Jμ4(x)=J,μSV(x)ud¯uu¯+dd¯2

to interpolate the Y states, and fit the correlation functions,

ΣY=Y(4220),Y(4360),Y(4390),Y(4660)λY2exp(MY2T2)=4mc2s0dsρQCD1(s)exp(sT2),

ΣY=Y(4220),Y(4360),Y(4390),Y(4660)λY2exp(MY2T2)=4mc2s0dsρQCD4(s)exp(sT2),

by taking the λY as free parameters. We obtain the best values, which are shown in Tab.24, at the pertinent energy scales μ=2.4GeV for the current Jμ4(x) and μ=2.9GeV for the current Jμ1(x), the values of the pole residue λY(4220) are very small. Without introducing explicit P-waves, we cannot produce the experimental mass of the Y(4260/4220) in the scenario of tetraquark state.

The Y(4660) has been studied extensively via the QCD sum rules [424, 431, 440, 444, 485, 512, 543-545, 548], however, no definite conclusion can be obtained, more works are still needed to decipher its structure.

Now let us turn to the pseudoscalar tetraquark states and write down the local currents,

J(x)=JAV+(x),JAV(x),JPS+(x),JPS(x),JTT+(x),JTT(x),

JAV+(x)=εijkεimn2[qjT(x)Cγμck(x)q¯m(x)γ5γμCc¯nT(x)qjT(x)Cγμγ5ck(x)q¯m(x)γμCc¯nT(x)],JAV(x)=εijkεimn2[qjT(x)Cγμck(x)q¯m(x)γ5γμCc¯nT(x)+qjT(x)Cγμγ5ck(x)q¯m(x)γμCc¯nT(x)],JPS+(x)=εijkεimn2[qjT(x)Cck(x)q¯m(x)γ5Cc¯nT(x)+qjT(x)Cγ5ck(x)q¯m(x)Cc¯nT(x)],JPS(x)=εijkεimn2[qjT(x)Cck(x)q¯m(x)γ5Cc¯nT(x)qjT(x)Cγ5ck(x)q¯m(x)Cc¯nT(x)],JTT+(x)=εijkεimn2[qjT(x)Cσμνck(x)q¯m(x)γ5σμνCc¯nT(x)+qjT(x)Cσμνγ5ck(x)q¯m(x)σμνCc¯nT(x)],JTT(x)=εijkεimn2[qjT(x)Cσμνck(x)q¯m(x)γ5σμνCc¯nT(x)qjT(x)Cσμνγ5ck(x)q¯m(x)σμνCc¯nT(x)],

with q, q=u, d, s, the superscripts ± symbolize the positive and negative charge-conjugation, respectively, the subscripts P, S, V, A and T stand for the pseudoscalar, scalar, vector, axialvector and tensor diquark operators, respectively [515].

Under parity transformation P^, the J(x) have the property,

P^J(x)P^1=J(x~).

Under charge-conjugation transformation C^, the J(x) have the property,

C^J±(x)C^1=±J±(x)qq,

and we can prove that the current JTT(x)=0 through performing the Fierz-transformation. Again, we take the isospin limit mu=md, the four-quark currents with the symbolic quark structures,

c¯cd¯u,c¯cu¯d,c¯cu¯ud¯d2,c¯cu¯u+d¯d2,

couple potentially to the pseudoscalar tetraquark states with degenerated masses. And the four-quark currents with the symbolic quark structures,

c¯cu¯s,c¯cd¯s,c¯cs¯u,c¯cs¯d,

also couple potentially to the pseudoscalar tetraquark states with degenerated masses according to the isospin symmetry. And we obtain the QCD sum rules routinely.

In Tab.25, we present the Borel windows, continuum threshold parameters, energy scales of the QCD spectral densities and pole contributions. From the Table, we can see distinctly that the pole contributions are about (40−60)% at the hadron side, while the central values are larger than 50%, the pole dominance criterion is satisfied very good. On the other hand, the higher vacuum condensates play a minor important role, the operator product expansion converges very well.

We take all the uncertainties of the parameters into account and acquire the masses and pole residues, see Tab.26. From Tab.25 and Tab.26, we can see distinctly that the modified energy scale formula μ=MX/Y/Z2(2Mc)2kms(μ) with k=0, 1 or 2 is satisfied, where we subtract the small s-quark mass approximately to account for the small light-flavor SU(3) mass-breaking effects, which is slightly different from Eq. (100).

As can be seen distinctly from Tab.26 that the lowest mass of the pseudoscalar hidden-charm tetraquark state with the symbolic quark constituents cc¯ud¯ is about 4.56±0.08GeV, which is much larger than the value 4239±1810+45MeV from the LHCb Collaboration [179]. In 2014, the LHCb Collaboration provided the first independent confirmation of the existence of the Zc(4430) in the ψπ mass spectrum and established its spin-parity to be JP=1+ [179]. Furthermore, the LHCb Collaboration observed a weak evidence for an additional resonance, the Zc(4240), in the ψπ mass spectrum with the preferred spin-parity JP=0 and the Breit−Wigner mass 4239±1810+45MeV and width 220±47074+108MeV, respectively with large uncertainties [179]. If the Zc(4240) is confirmed by further experiments in the future, it is an excellent candidate for the hidden-charm tetraquark state with the JPC=0, and we should revisit the QCD sum rules for the discrepancy.

In Ref. [437], Chen and Zhu studied the hidden-charm tetraquark states with the symbolic quark constituents cc¯ud¯ with the QCD sum rules, and obtained the ground state masses 4.55±0.11GeV for the tetraquark states with the JPC=0, the masses 4.55±0.11GeV, 4.67±0.10GeV, 4.72±0.10GeV for the tetraquark states with the JPC=0+. The present predictions are consistent with their calculations, again, we should bear in mind that their interpolating currents and schemes in treating the operator product expansion and input parameters at the QCD side differ from the present work remarkably. Any current with the same quantum numbers and same quark structure as a Fock state in a hadron couples potentially to this hadron, so we can construct several currents to interpolate a hadron, or construct a current to interpolate several hadrons.

From Tab.26, we can see explicitly that the central values of the masses of the JPC=0+ tetraquark states with the symbolic quark constituents ucd¯c¯, ucs¯c¯, scs¯c¯ are about 4.564.58GeV, 4.614.62GeV and 4.664.67GeV, respectively, the central values of the masses of the JPC=0 tetraquark states with the symbolic quark constituents ucd¯c¯, ucs¯c¯ and scs¯c¯ are about 4.58GeV, 4.63GeV and 4.67GeV, respectively. We obtain the conclusion tentatively that the currents JAV+(x), JPS+(x) and JTT+(x) (JAV(x) and JPS(x)) couple potentially to three (two) different pseudoscalar tetraquark states with almost degenerated masses, or to one pseudoscalar tetraquark state with three (two) different Fock components. As the currents with the same quantum numbers couple potentially to the pseudoscalar tetraquark states with almost degenerated masses, the mixing effects cannot improve the predictions remarkably if only the tetraquark masses are concerned. All in all, we obtain reasonable predictions for the masses of the pseudoscalar tetraquark states without strange, with strange and with hidden-strange, the central values are about 4.564.58GeV, 4.614.63GeV and 4.664.67GeV, respectively.

The following two-body strong decays of the pseudoscalar hidden-charm tetraquark states,

Zc(0χc1ρ,ηcρ,J/ψa1(1260),J/ψπ,DD¯0+h.c.,DD¯1+h.c.,DD¯+h.c.,Zc(0+)χc0π,ηcf0(500),J/ψρ,DD¯0+h.c.,DD¯1+h.c.,DD¯+h.c.,Zcs(0)χc1K,ηcK,J/ψK1,J/ψK,DsD¯0+h.c.,DD¯s0+h.c.,DsD¯1+h.c.,DD¯s1+h.c.,DsD¯+h.c.,DD¯s+h.c.,Zcs(0+)χc0K,ηcK0(700),J/ψK,DsD¯0+h.c.,DD¯s0+h.c.,DsD¯1+h.c.,DD¯s1+h.c.,DsD¯+h.c.,DD¯s+h.c.,Zcss(0)χc1ϕ,ηcϕ,J/ψf1,J/ψη,DsD¯s0+h.c.,DsD¯s1+h.c.,DsD¯s+h.c.,Zcss(0+)χc0η,ηcf0(980),J/ψϕ,DsD¯s0+h.c.,DsD¯s1+h.c.,DsD¯s+h.c.,

can take place through the Okubo−Zweig−Iizuka super-allowed fall-apart mechanism, we suggest to search for the pseudoscalar hidden-charm tetraquark states in those channels.

The QCD sum rules obtained in this sub-section can be extended directly to study the tetraquark states in the bottom sector with the simple replacements cb and c¯b¯.

3.1.4 Tetraquark states with an explicit P-wave

In the type-II diquark model [56], Maiani et al assign the Y(4008), Y(4260), Y(4290/4220) and Y(4630) as four tetraquark states with the L=1 based on the effective spin-spin and spin-orbit interactions, see Eq. (24). In Ref. [318], A. Ali et al incorporate the dominant spin-spin, spin-orbit and tensor interactions, see Eq. (25), and observe that the preferred assignments of the tetraquark states with the L=1 are the Y(4220), Y(4330), Y(4390), Y(4660). In the diquark model, the quantum numbers of the Y states are shown explicitly in Tab.27, where the L is the angular momentum between the diquark and antidiquark, S=Sqc+Sq¯c¯, J=S+L, and L=1 denotes the explicit P-wave.

We take the isospin limit, and construct the interpolating currents according to the quantum numbers shown in Tab.27,

Jμ1(x)=εijkεimn2ujT(x)Cγ5ck(x)μd¯m(x)γ5Cc¯nT(x),Jμ2(x)=εijkεimn2ujT(x)Cγαck(x)μd¯m(x)γαCc¯nT(x),Jμ3(x)=εijkεimn2[ujT(x)Cγμck(x)αd¯m(x)γαCc¯nT(x)+ujT(x)Cγαck(x)αd¯m(x)γμCc¯nT(x)],Jμν(x)=εijkεimn22[ujT(x)Cγ5ck(x)μd¯m(x)γνCc¯nT(x)+ujT(x)Cγνck(x)μd¯m(x)γ5Cc¯nT(x)ujT(x)Cγ5ck(x)νd¯m(x)γμCc¯nT(x)ujT(x)Cγμck(x)νd¯m(x)γ5Cc¯nT(x)],

where μ=μμ embodies the explicit P-wave.

Under charge conjugation transformation C^, the currents Jμ(x) and Jμν(x) have the property,

C^Jμ(x)C^1=Jμ(x),C^Jμν(x)C^1=Jμν(x),

the currents have definite charge conjugation.

We choose the currents Jμ(x)=Jμ1(x), Jμ2(x), Jμ3(x) and Jμν(x) and resort to the correlation functions in Eq. (125) to study the vector tetraquark states using the modified energy scale formula,

μ=MX/Y/Z2(2Mc+0.5GeV)2=MX/Y/Z2(4.1GeV)2,

to determine the ideal energy scales of the QCD spectral densities, and reexamine the possible assignments of the Y states [549, 550]. The numerical results are shown explicitly in Tab.28 and Tab.29.

The predicted mass MY=4.24±0.10GeV of the |0,0;0,1;1 tetraquark state is in excellent agreement with the experimental data MY(4220)=4222.0±3.1±1.4MeV from the BESIII Collaboration [154], or MY(4260)=4230.0±8.0MeV from the Particle Data Group [411], which supports assigning the Y(4260/4220) as the Cγ5μγ5C type vector tetraquark state. We obtain the vector hidden-charm tetraquark state with the lowest mass up to now.

There have been other possible assignments for the Y(4260) states, such as the hybrid states [303, 551, 552, 553, 554], molecular state [227, 229, 280, 555, 556], baryonium states [557, 558], hadro-charmonium state [559], and interference effect [560, 561].

From Tab.22 and Tab.29, we can see explicitly that there are no rooms for the Y(4008) and Y(4750) in the hidden-charm tetraquark scenario. If we assign the Y(4220/4230/4260) as the ground state, then we could assign the Y(4750) as its first radial excitation according to the mass gap MY(4750)MY(4260)=0.51GeV [562].

We carry out the calculations routinely to obtain two QCD sum rules,

λY2exp(MY2T2)=4mc2s0dsρQCD(s)exp(sT2),

λY2exp(MY2T2)+λY2exp(MY2T2)=4mc2s0dsρQCD(s)exp(sT2),

where the s0 and s0 correspond to the ground states Y and first radial excitations Y, respectively [562].

We adopt the QCDSR II, see Eq. (151) and Eqs. (158) and (159), to study the radially excited states, and obtain the Borel windows, continuum threshold parameters, suitable energy scales and pole contributions, which are shown explicitly in Tab.30 and Tab.31. From the tables, we can see explicitly that the pole contributions of the 1P states (the 1P plus 2P states) are about (40−60)% ((67−85)%), the pole dominance is satisfied very well. On the other hand, the contributions from the highest dimensional condensates play a minor important role, |D(10)|<3% or 1% (<1% or 1%) for the 1P states (the 1P plus 2P states), the operator product expansion converges very good and better than that in our previous work [550], see Tab.28.

The predicted masses and pole residues are presented in Tab.32. From Tab.31 and Tab.32, we can see explicitly that the modified energy scale formula, see Eq. (189), can be well satisfied, and the relations s0=MY+0.500.55±0.10GeV and s0=MY+(0.40±0.10)GeV are hold, see Eq. (160).

In Fig.19, we plot the masses of the 1P and 2P hidden-charm tetraquark states with the JPC=1. From the figure, we can see explicitly that there appear flat platforms in the Borel windows, the uncertainties come from the Borel parameters are rather small.

In Tab.33, we present the possible assignments of the vector tetraquark states [562]. From the table, we can see explicitly that there is a room to accommodate the Y(4750), i.e., the Y(4220/4260) and Y(4750) can be assigned as the ground state and first radial excited state of the Cγ5μγ5C type tetraquark states with the JPC=1, respectively [562].

We can study the corresponding hidden-bottom tetraquark states with the simple replacement cb in Eq. (187). In Ref. [563], we observe that the Y(10750) observed by the Belle Collaboration [190] can be assigned as the Cγ5μγ5C type hidden-bottom tetraquark state with the JPC=1.

3.2 Doubly heavy tetraquark states

In 2016, the LHCb Collaboration observed the doubly-charmed baryon state Ξcc++ in the Λc+Kπ+π+ mass spectrum and measured the mass, but did not determine the spin [564]. The Ξcc++ maybe have the spin 12 or 32, we can take the diquark εijkciTCγμcj as basic constituent to construct the current

J(x)=εijkciT(x)Cγμcj(x)γ5γμuk(x),

or

Jμ(x)=εijkciT(x)Cγμcj(x)uk(x),

to study it with the QCD sum rules [565]. The observation of the doubly-charmed baryon state Ξcc++ has led to a renaissance on the doubly-heavy tetraquark spectroscopy [566-568]. For a QQq¯q¯ system, if the two Q-quarks are in long separation, the gluon exchange induced force between them would be screened by the two q¯-quarks, then a loosely Qq¯Qq¯ type bound state is formed. If the two Q-quarks are in short separation, the QQ pair forms a compact point-like color source in heavy quark limit, and attracts a q¯q¯ pair, which serves as another compact point-like color source, then an exotic QQq¯q¯ type tetraquark state is formed. The existence and stability of the QQq¯q¯ tetraquark states have been extensively discussed in early literatures based on the potential models [569-574] and heavy quark symmetry [575].

In Ref. [576], we choose the currents Jμ(x) and ημ(x) to study the doubly heavy tetraquark states with the JP=1+, where

Jμ(x)=εijkεimnQjT(x)CγμQk(x)u¯m(x)γ5Cs¯nT(x),ημ(x)=εijkεimnQjT(x)CγμQk(x)u¯m(x)γ5Cd¯nT(x),

Q=c,b, again, we adopt the correlation functions Πμν(p) in Eq. (125). The tetraquark states are spatial extended objects, not point-like objects, however, we choose the local currents to interpolate them and take all the quarks and antiquarks as the color sources, and neglect the finite size effects.

We rewrite the current Jμ(x) as

Jμ(x)=QjT(x)CγμQk(x)[u¯j(x)γ5Cs¯kT(x)u¯k(x)γ5Cs¯jT(x)]=12[QjT(x)CγμQk(x)QkT(x)CγμQj(x)][u¯j(x)γ5Cs¯kT(x)u¯k(x)γ5Cs¯jT(x)],

according to the identity εijkεimn=δjmδknδjnδkm in the color space. The current Jμ(x) is of 3¯3 type in the color space, we can also construct the current J~μ(x) satisfying the Fermi−Dirac statistics,

J~μ(x)=12[QjT(x)Cγ5Qk(x)+QkT(x)Cγ5Qj(x)][u¯j(x)γμCs¯kT(x)+u¯k(x)γμCs¯jT(x)],

which is of 66¯ type in the color space, and differs from the corresponding current constructed in Ref. [577] slightly.

The color factor defined in Eq. (20) has the values, C^iC^j=23 and 13 for the 3¯ and 6 diquark [qq], respectively. If we define C^12C^34=(C^1+C^2)(C^3+C^4), then C^12C^34=43 and 103 for the 3¯3 and 66¯ type tetraquark states. The one-gluon exchange induced attractive (repulsive) interaction favors (disfavors) formation of the 3¯ (6) diquark state QjTCγμQkQkTCγμQj (QjTCγ5Qk+QkTCγ5Qj), while the 66¯-type tetraquark states are expected to have much smaller masses than that of the 3¯3-type tetraquark states according to the C^12C^34. Furthermore, the color magnetic interaction, see Eq. (21), leads to mixing between the 66¯ and 3¯3-type tetraquark states. In Ref. [577], Du et al. obtained degenerate masses for the 66¯ and 3¯3-type tetraquark states based on the QCD sum rules. We should study this subject further.

After tedious but straightforward calculations, we obtain the QCD sum rules for the doubly-heavy tetraquark states [576], which are named as the T states in The Review of Particle Physics [97]. According to the energy scale formula in Eq. (99), we suggest an energy scale formula,

μ=MT2(2MQ)2κms(μ),

to determine the optimal energy scales of the QCD spectral densities.

There was no experimental candidate for the doubly heavy tetraquark state when performing the calculations [576]. After careful examinations, we choose the effective heavy quark masses Mc=1.84GeV and Mb=5.12GeV, and take account of the SU(3) breaking effect by subtracting the κms(μ). In Tab.34, we present the Borel windows T2, continuum threshold parameters s0, optimal energy scales μ, pole contributions of the ground states, where the same parameters as the ones in the QCD sum rules for the Zc(3900) are chosen, see the last line.

We take into account all the uncertainties of the relevant parameters, and obtain the values of the masses and pole residues of the ZQQ, which are shown explicitly in Tab.34 and Fig.20. From Fig.20, we can see explicitly that there appear platforms in the Borel windows shown in Tab.34 indeed. And we suggested to search for the ZQQ states in the Okubo−Zweig−Iizuka super-allowed two-body strong decays

Zccu¯d¯D0D+,D+D0,Zccu¯s¯D0Ds+,Ds+D0,

and weak decays through bcc¯s at the quark level,

Zbbu¯d¯B¯0B,BB¯0γJ/ψKJ/ψK¯0,Zbbu¯s¯B¯s0B,BB¯s0γJ/ψϕJ/ψK.

In 2021, the LHCb Collaboration observed the exotic state Tcc+(3875) just below the D0D+ threshold [194, 195]. The Breit−Wigner mass and width are δMBW=273±61±514+11KeV below the D0D+ threshold and ΓBW=410±165±4338+18KeV [194, 195]. While the Particle Data Group fit the Breit−Wigner mass and width to be 3874.83±0.11MeV and 0.410±0.1650.057+0.047MeV, respectively [97]. The prediction MZccu¯d¯=3.90±0.09GeV is in excellent agreement with the LHCb data.

Before the LHCb data, several theoretical groups had made predictions for the Tcc masses [276, 285, 336, 342, 345, 351, 354, 360, 361, 566, 567, 576-596], the predicted masses differ from each other in one way or the other.

In Ref. [578], we resort to the correlation functions Π(p) and Πμναβ(p) in Eq. (125) and currents

J(x)=Ju¯d¯;0(x),Ju¯s¯;0(x),Js¯s¯;0(x),Jμν(x)=Jμν;1(x),Jμν;2(x),Jμν;1/2(x)=Jμν;u¯d¯;1/2(x),Jμν;u¯s¯;1/2(x),Jμν;s¯s¯;1/2(x),

where

Ju¯d¯;0(x)=εijkεimncjT(x)Cγμck(x)u¯m(x)γμCd¯nT(x),Ju¯s¯;0(x)=εijkεimncjT(x)Cγμck(x)u¯m(x)γμCs¯nT(x),Js¯s¯;0(x)=εijkεimncjT(x)Cγμck(x)s¯m(x)γμCs¯nT(x),Jμν;u¯d¯;1/2(x)=εijkεimn[cjT(x)Cγμck(x)u¯m(x)γνCd¯nT(x)cjT(x)Cγνck(x)u¯m(x)γμCd¯nT(x)],Jμν;u¯s¯;1/2(x)=εijkεimn[cjT(x)Cγμck(x)u¯m(x)γνCs¯nT(x)cjT(x)Cγνck(x)u¯m(x)γμCs¯nT(x)],Jμν;s¯s¯;1/2(x)=εijkεimn[cjT(x)Cγμck(x)s¯m(x)γνCs¯nT(x)cjT(x)Cγνck(x)s¯m(x)γμCs¯nT(x)],

to study the mass spectrum of the JP=0+, 1± and 2+ doubly charmed tetraquark states systematically, where the subscripts 0, 1 and 2 denote the spins.

At the phenomenological side, we obtain the hadronic representation and isolate the ground state contributions,

Π(p)=λZ2MZ2p2+=Π0(p2),

Πμναβ;1(p)=λ~Z2MZ2p2(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+λ~Y2MY2p2(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+=ΠZ(p2)(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+ΠY(p2)(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ),

Πμναβ;2(p)=λZ2MZ2p2(g~μαg~νβ+g~μβg~να2g~μνg~αβ3)+=Π2(p2)(g~μαg~νβ+g~μβg~να2g~μνg~αβ3),

where g~μν=gμνpμpνp2, the pole residues λZ and λY are defined analogous to Eq. (127) with the simple replacements ZY and Z+Z, and λY=λ~YMY, λZ=λ~ZMZ.

We perform the calculations routinely, and obtain the QCD sum rules for the masses and pole residues from the components Π0(p2), Π1;A(p2), Π1;V(p2) and Π2(p2), respectively, where Π1;A(p2)=p2ΠZ(p2) and Π1;V(p2)=p2ΠY(p2). Again, we adopt the modified energy scale formula in Eq. (197) to determine the best energy scales of the QCD spectral densities.

After trial and error, we obtain the Borel windows T2, continuum threshold parameters s0, optimal energy scales μ, pole contributions, see Tab.35. From the Table, we can see clearly that the pole dominance can be well satisfied. In calculations, we observe that for the JP=1 tetraquark states, the operator product expansion is well convergent, while in the case of the JP=0+, 1+ and 2+ tetraquark states, the contributions of the vacuum condensates of dimensions 6,8,10 have the hierarchy |D(6)||D(8)||D(10)|, the operator product expansion is also convergent. At last, we take account of all uncertainties of the relevant parameters, and obtain the values of the masses and pole residues, which are shown explicitly in Tab.35.

The centroids of the masses of the CγμγνC type tetraquark states are

MCγμγνC(ccu¯d¯)=Mccu¯d¯;0++3Mccu¯d¯;1++5Mccu¯d¯;2+9=3.92GeV,MCγμγνC(ccu¯s¯)=Mccu¯s¯;0++3Mccu¯s¯;1++5Mccu¯s¯;2+9=3.99GeV,MCγμγνC(ccs¯s¯)=Mccs¯s¯;0++3Mccs¯s¯;1++5Mccs¯s¯;2+9=4.04GeV,

which are slightly larger than the centroids of the masses of the corresponding Cγμγ5C type tetraquark states,

MCγμγ5C(ccu¯d¯)=3.90GeV,MCγμγ5C(ccu¯s¯)=3.95GeV,

the lowest states are the Cγμγ5C type tetraquark states, which is consistent with our naive expectation that the axialvector (anti)diquarks have larger masses than the corresponding scalar (anti)diquarks. The lowest centroids Mccu¯d¯;0+=3.87GeV and Mccu¯s¯;0+=3.94GeV originate from the spin splitting, in other words, the spin-spin interaction between the doubly heavy diquark and light antidiquark. In fact, the predicted masses have uncertainties, the centroids of the masses are not the super values, all values within uncertainties make sense.

The QCD sum rules indicate that the masses of the light axialvector diquark states lie about (150200)MeV above that of the light scalar diquark states [501-504], if they have the same valence quarks. Therefore, the centroids of the masses of the CγμγνC type tetraquark states should be larger than 4.0GeV, the present calculations maybe under-estimate the doubly-heavy tetraquark masses. If we take the simple replacement ms(μ)Ms in the modified energy scale formula in Eq. (197), the predictions should be improved, about +100MeV.

After observation of the Tcc(3875), several new works on the Tcc(3875) in the 3¯3 type tetraquark scenario appear [324, 472, 597-601]. Roughly speaking, the centroid of the CγαγβC type tetraquark states maybe lie about 100MeV above the corresponding Cγαγ5C type tetraquark states, and more works are still needed.

The doubly-charmed tetraquark states with the JP=0+, 1+ and 2+ lie near the corresponding charmed meson pair thresholds, the decays are Okubo−Zweig−Iizuka super-allowed,

Zccu¯d¯;0+D0D+,Zccu¯s¯;0+D0Ds+,Zccs¯s¯;0+Ds+Ds+,Zccu¯d¯;1+D0D+,D+D0,Zccu¯s¯;1+D0Ds+,Ds+D0,Zccs¯s¯;1+Ds+Ds+,Zccu¯d¯;2+D0D+,D0D+,Zccu¯s¯;2+D0Ds+,Zccs¯s¯;2+Ds+Ds+,

but the available phase spaces are very small, thus the decays are kinematically depressed, the doubly charmed tetraquark states with the JP=0+, 1+ and 2+ maybe have small widths. On the other hand, the doubly charmed tetraquark states with the JP=1 lie above the corresponding charmed meson pair thresholds, the decays are Okubo−Zweig−Iizuka super-allowed,

Yccu¯d¯;1D0D+,D0D+,D+D0,Yccu¯s¯;1D0Ds+,D0Ds+,Ds+D0,Yccs¯s¯;1Ds+Ds+,Ds+Ds+,

and the available phase spaces are large, thus the decays are kinematically facilitated, the doubly charmed tetraquark states with the JP=1 should have large widths.

3.3 Fully heavy tetraquark states

The exotic states Zc(3900), Zc(4020), Zc(4430), Tcc(3875), Pc(4312), Pc(4380), Pc(4440), Pc(4457), Zb(10610), Zb(10650), are excellent candidates for the multiquark states, which consist two heavy quarks and two or three light quarks, we have to deal with both the heavy and light degrees of freedom of the dynamics. If there are multiquark configurations consisting of fully heavy quarks, the dynamics is much simple at first glance, and the QQQ¯Q¯ tetraquark states have been studied extensively before the LHCb data [323, 333, 346, 349, 493, 569, 602-613].

The quarks have color SU(3) symmetry, we can construct the tetraquark states according to the routine quarkdiquarktetraquark,

(33)(3¯3¯)(3¯6)(36¯)(3¯3)(66¯)(18).

For the 3¯ diquarks, only the operators εijkQjTCγμQk and εijkQjTCσμνQk could exist due to Fermi−Dirac statistics, and we usually take the operators εijkQjTCγμQk to construct the four-quark currents J(x) and Jμν(x) [610, 611], where Jμν(x)=Jμν1(x), Jμν2(x), and

J(x)=εijkεimnQjT(x)CγμQk(x)Q¯m(x)γμCQ¯nT(x),Jμν1(x)=εijkεimn{QjT(x)CγμQk(x)Q¯m(x)γνCQ¯nT(x)QjT(x)CγνQk(x)Q¯m(x)γμCQ¯nT(x)},Jμν2(x)=εijkεimn2{QjT(x)CγμQk(x)Q¯m(x)γνCQ¯nT(x)+QjT(x)CγνQk(x)Q¯m(x)γμCQ¯nT(x)},

then resort to the correlation functions Π(p) and Πμναβ(p) shown in Eq. (125) to obtain the QCD sum rules.

In Ref. [493], Chen et al. constructed the currents ηi(x), ημk(x) and ημνj(x) with i=1,2,3,4,5, k=1,2 and j=1,2, to interpolate the QQQ¯Q¯ tetraquark states with the JPC=0++, 1+ and 2++, respectively,

ηi(x)=QaT(x)CΓiQb(x)Q¯a(x)ΓiCQ¯bT(x),ημ1(x)=QaT(x)Cγμγ5Qb(x)Q¯a(x)CQ¯bT(x)QaT(x)CQb(x)Q¯a(x)γμγ5CQ¯bT(x),ημ2(x)=QaT(x)Cσμνγ5Qb(x)Q¯a(x)γνCQ¯bT(x)QaT(x)CγνQb(x)Q¯a(x)σμνγ5CQ¯bT(x),ημνj(x)=QaT(x)CΓμjQb(x)Q¯a(x)ΓνjCQ¯bT(x)+QaT(x)CΓνjQb(x)Q¯a(x)ΓμjCQ¯bT(x),

where Γ1=γ5, Γ2=γμγ5, Γ3=σμν, Γ4=γμ, Γ5=1, Γμ1=γμ, Γμ2=γμγ5, the a and b are color indexes. The Cγ5, C, Cγμγ5 are antisymmetric, while the Cγμ, Cσμν, Cσμνγ5 are symmetric. The currents η1/2/5(x), ημ1(x) and ημν2(x) are in the color 66¯ representation, while the currents η3/4(x), ημ2(x) and ημν1(x) are in the color 3¯3 representation. For more currents and predictions in Scheme II, we can consult Ref. [493].

At the hadron side, we isolate the ground state contributions and obtain the results [610, 611],

Π(p)=λX2MX2p2+=ΠS(p2),

Πμναβ1(p)=λ~Y+2MY+2p2(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+λ~Y2MY2p2(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+=ΠA(p2)(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+ΠV(p2)(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ).

Πμναβ2(p)=λX2MX2p2(g~μαg~νβ+g~μβg~να2g~μνg~αβ3)+=ΠT(p2)(g~μαg~νβ+g~μβg~να2g~μνg~αβ3)+,

where g~μν=gμνpμpνp2, we add the superscripts 1 and 2 to denote the spins, and define the pole residues λX and λY with λY±=λ~Y±MY± analogous to Eq. (127) with the simple replacements Z+X, Y+ and ZY.

If we take into account the first radial excited states, we obtain

ΠS/T(p2)=λX2MX2p2+λX2MX2p2+,ΠA/V(p2)=λ~Y±2MY±2p2+λ~Y±2MY±2p2+.

Again we take the quark-hadron duality below the continuum thresholds s0 and s0, respectively, and perform the Borel transformation with respect to the variable P2=p2 to obtain the QCD sum rules:

λX/Y2exp(MX/Y2T2)=16mc2s0dszizfdztitfdtrirfdrρ(s,z,t,r)exp(sT2),

λX/Y2exp(MX/Y2T2)+λX/Y2exp(MX/Y2T2)=16mc2s0dszizfdztitfdtrirfdrρ(s,z,t,r)exp(sT2),

where the QCD spectral densities ρ(s,z,t,r)=ρS(s,z,t,r), ρA(s,z,t,r), ρV(s,z,t,r) and ρT(s,z,t,r) [610, 611, 614].

In Fig.21, we plot the masses of the ccc¯c¯ tetraquark states with variations of the energy scales and Borel parameters for the threshold parameters sS0=42GeV2 and sT0=44GeV2. The predicted masses decrease monotonously and slowly with increase of the energy scales, the QCD sum rules are stable with variations of the Borel parameters at the energy scales 1.2GeV<μ<2.2GeV. We take the largest energy scale μ=2.0GeV in Ref. [610]. In Refs. [610, 611], we obtain the predictions for the masses of the ground states, see Tab.36, where Mccc¯c¯(0++)<2MJ/ψ.

In 2020, the LHCb Collaboration observed a broad structure above the J/ψJ/ψ threshold ranging from 6.2 to 6.8 GeV and a narrow structure at about 6.9 GeV in the J/ψJ/ψ mass spectrum, and they also observed some vague structures around 7.2 GeV [142]. Accordingly, we obtain the masses of the 2S and 2P states from the QCDSR II in Eq. (217), and obtain the 3/4S and 3/4P masses by fitting the Regge trajectories,

Mn2=α(n1)+α0,

where the α and α0 are constants, see Tab.37 and Tab.38, which support assigning the broad structure from 6.2 to 6.8 GeV in the di-J/ψ mass spectrum as the 2S or 2P tetraquark state, and assigning the narrow structure at about 6.9 GeV in the di-J/ψ mass spectrum as the 3S tetraquark state [614].

In 2023, the ATLAS Collaboration observed a narrow resonance at about 6.9GeV and a broader structure at much lower mass in the J/ψJ/ψ channel, moreover, they observed a statistically significant excess at about 7.0GeV in the J/ψψ channel [143]. In 2024, the CMS Collaboration observed three resonant structures in the J/ψJ/ψ mass spectrum,

X(6600):M=6552±10±12MeV,Γ=12426+32±33MeV,X(6900):M=6927±9±4MeV,Γ=12221+24±18MeV,X(7300):M=728718+20±5MeV,Γ=9540+59±19MeV,

in the no-interference model [144].

In Ref. [615], we update the calculations by taking the energy scale μ=1.4GeV, the lower bound in Fig.21, and adopt the relation,

M1<s0<M2<s0<M3,

and obtain the predictions, see Tab.39 and Tab.40, and make the possible assignments, see Tab.41. From Tab.41, we can see explicitly that the lowest state lies at about 6.2 GeV, which is consistent with the recent coupled-channel analysis [616].

The thresholds of the J/ψJ/ψ, J/ψψ and J/ψψ(3770) are 6194MeV, 6783MeV and 6875MeV, respectively [97], we cannot obtain a simple molecule scenario to interpret those X states without introducing complex coupled-channel effects [617-620], or just assign them as the 3¯3 and 66¯ type tetraquark states [334, 337, 344, 352, 460, 595, 621-631], or with gluonic constituent [632].

We perform Fierz transformation for the currents J(x) and Jμν(x), and obtain particular superpositions of a series of color 11 type currents,

J=2Q¯QQ¯Q+2Q¯iγ5QQ¯iγ5Q+Q¯γαQQ¯γαQQ¯γαγ5QQ¯γαγ5Q,Jμν1=2iQ¯σμνQQ¯Q+2iεμναβQ¯γαγ5QQ¯γβQ2Q¯σμνγ5QQ¯iγ5Q,J~μν2=2Q¯γμγ5QQ¯γνγ5Q2Q¯γμQQ¯γνQ+gαβ(Q¯σμαQQ¯σνβQ+Q¯σναQQ¯σμβQ)+gμν(Q¯QQ¯Q+Q¯iγ5QQ¯iγ5Q+Q¯γαQQ¯γαQQ¯γαγ5QQ¯γαγ5Q12Q¯σαβQQ¯σαβQ),

with J~μν2=2Jμν2, the components Q¯ΓQQ¯ΓQ couple potentially to the molecular states, where the Γ and Γ stand for some Dirac γ-matrixes. Therefore the 3¯3 type tetraquark states have some important 11 type Fock components, which would decay to their constituents via the Okubo−Zweig−Iizuka super-allowed fall apart mechanism if they are kinematically permitted. There exists a term Q¯γμQQ¯γνQ, the decay to the di-J/ψ is super-allowed, which is consistent with the observations of the ATLAS, CMS and LHCb experiments.

If we insist on that the di-J/ψ system should have positive charge conjugation, we would like to construct a cousin of currents, J,μA~A(x) and J+,μA~A(x),

J,μA~A(x)=εijkεimn2[cjT(x)Cσμνγ5ck(x)c¯m(x)γνCc¯nT(x)cjT(x)Cγνck(x)c¯m(x)γ5σμνCc¯nT(x)],J+,μA~A(x)=εijkεimn2[cjT(x)Cσμνγ5ck(x)c¯m(x)γνCc¯nT(x)+cjT(x)Cγνck(x)c¯m(x)γ5σμνCc¯nT(x)],

which couple potentially to the fully-charm tetraquark states with the JPC=1+ and 1++, respectively.

In Ref. [633], we introduce a relative P-wave to construct the doubly-charmed vector diquark operator V^, then construct the scalar and tensor four-quark currents,

J(x)=εijkεimncjT(x)Cγ5μck(x)c¯m(x)νγ5Cc¯nT(x)gμν,Jμν1(x)=εijkεimn{cjT(x)Cγ5μck(x)c¯m(x)νγ5Cc¯nT(x)cjT(x)Cγ5νck(x)c¯m(x)μγ5Cc¯nT(x)},Jμν2(x)=εijkεimn{cjT(x)Cγ5μck(x)c¯m(x)νγ5Cc¯nT(x)+cjT(x)Cγ5νck(x)c¯m(x)μγ5Cc¯nT(x)},

to study the scalar, axialvector and tensor fully-charm tetraquark states with the QCD sum rules. And we observe that the ground state V^V^ type tetraquark states and the first radial excited states of the AA type tetraquark states have almost degenerated masses.

We can extend this subsection directly to study the 3¯3¯3¯ type fully heavy pentaquark states and 333 type fully heavy hexaquark states [634, 635].

4 11 type tetraquark states

The X, Y, Z, T and P states always lie near the two-particle thresholds, such as

DD¯/D¯D:X(3872),Zc(3885/3900),DD:Tcc(3875),DD¯:Zc(4020/4025),DD¯s/DD¯s:Zcs(3985/4000),DsD¯s:X(4140),DD¯1/D¯D1:Y(4260/4220),Zc(4250),DD¯0/D¯D0:Y(4360/4320),D¯Σc:Pc(4312),D¯Ξc:Pcs(4338),D¯Ξc/D¯Ξc:Pcs(4459),D¯Σc:Pc(4380),D¯Σc:Pc(4440/4457),Λc+Λc/f0(980)ψ:Y(4660),BB¯/B¯B:Zb(10610),BB¯:Zb(10650),

naively, we expect that they consist of two color-neutral clusters, and they are molecular states, more precisely, they are the 11 type hidden-charm or doubly-charmed tetraquark or pentaquark states [636]. The establishment of the JPC=1++ of the Y(4140) by the LHCb Collaboration [113, 114] excludes its assignment as the DsD¯s molecular state with the JPC=0++ [432, 435, 477, 478, 637, 638], however, which does not mean non-existence of the DsD¯s molecular state with the JPC=0++.

The 3¯3 type four-quark currents could be reformed in a series of 11 type four-quark currents through Fierz transformation [424, 639], some useful examples are given explicitly in the appendix, see Eqs. (482)−(489).

According to the quark-hadron duality, the 3¯3 and 11 type local currents couple potentially to the 3¯3 and 11 type tetraquark states, respectively. The 3¯3 type tetraquark states could be taken as a particular superposition of a series of the 11 type tetraquark states, while the 11 type tetraquark states could decay through the Okubo−Zweig−Iizuka super-allowed fall-apart mechanism. We usually use the identities in Eqs. (482)−(489) to analyze the strong decays [424, 639]. For example, the current in Eq. (482) couples potentially to the Zc(3900), in the nonrelativistic and heavy quark limit, the component c¯σμνγ5ud¯γνc can be reduced to the form,

c¯σ0jγ5ud¯γjcξcσjζuχdσkdσjξcξcσj2ζuχdσj2ξc=SD¯SD,c¯σijγ5ud¯γjcϵijkξcσkσkuζuχdσkdσjξcϵijkξcσk2ζuχdσj2ξc=SD×SD¯,

where the ξ, ζ and χ are two-component spinors of the quark fields, the k are the three-vectors of the quark fields, the σi are the pauli matrixes, and the S are the spin operators. It is obvious that the currents c¯σμνγ5ud¯γνc and c¯γνud¯σμνγ5c couple potentially to the JP=0+ and 1+ (DD¯)+ states. However, the strong decays Zc±(3900)(DD¯)± are kinematically forbidden.

4.1 Hidden heavy tetraquark states

Again, let us adopt the correlation functions Π(p), Πμν(p) and Πμναβ(p) defined in Eq. (125) and write down the currents

J(x)=JDD¯(x),JDD¯s(x),JDsD¯s(x),JDD¯(x),JDD¯s(x),JDsD¯s(x),Jμ(x)=JDD¯,±,μ(x),JDD¯s,±,μ(x),JDsD¯s,±,μ(x),Jμν(x)=J±,μν(x),J±,μν(x)=JDD¯,±,μν(x),JDD¯s,±,μν(x),JDsD¯s,±,μν(x),

and

JDD¯(x)=q¯(x)iγ5c(x)c¯(x)iγ5q(x),JDD¯s(x)=q¯(x)iγ5c(x)c¯(x)iγ5s(x),JDsD¯s(x)=s¯(x)iγ5c(x)c¯(x)iγ5s(x),JDD¯(x)=q¯(x)γμc(x)c¯(x)γμq(x),JDD¯s(x)=q¯(x)γμc(x)c¯(x)γμs(x),JDsD¯s(x)=s¯(x)γμc(x)c¯(x)γμs(x),

JDD¯,±,μ(x)=12[u¯(x)iγ5c(x)c¯(x)γμd(x)u¯(x)γμc(x)c¯(x)iγ5d(x)],JDD¯s,±,μ(x)=12[q¯(x)iγ5c(x)c¯(x)γμs(x)q¯(x)γμc(x)c¯(x)iγ5s(x)],JDsD¯s,±,μ(x)=12[s¯(x)iγ5c(x)c¯(x)γμs(x)s¯(x)γμc(x)c¯(x)iγ5s(x)],

JDD¯,±,μν(x)=12[u¯(x)γμc(x)c¯(x)γνd(x)±u¯(x)γνc(x)c¯(x)γμd(x)],JDD¯s,±,μν(x)=12[q¯(x)γμc(x)c¯(x)γνs(x)±q¯(x)γνc(x)c¯(x)γμs(x)],JDsD¯s,±,μν(x)=12[s¯(x)γμc(x)c¯(x)γνs(x)±s¯(x)γνc(x)c¯(x)γμs(x)],

and q=u, d. The subscripts DD¯, DD¯s, and DsD¯s stand for the two color-neutral clusters; especially, the subscripts DD¯,±, DD¯s,± and DsD¯s,± correspond to the two color-neutral clusters DD¯DD¯, DD¯sDD¯s and DsD¯sDsD¯s, respectively, etc.

Again, we take the isospin limit, the currents with the symbolic quark structures c¯cd¯u, c¯cu¯d, c¯cu¯ud¯d2, c¯cu¯u+d¯d2 couple potentially to the hidden-charm molecular states with degenerated masses, the currents with the isospin I=1 and 0 lead to the same QCD sum rules [81-83, 422].

Under parity transformation P^, the currents J(x), Jμ(x) and Jμν(x) have the properties,

P^J(x)P^1=+J(x~),P^Jμ(x)P^1=Jμ(x~),P^J±,μν(x)P^1=+J±,μν(x~),

where xμ=(t,x) and x~μ=(t,x). Under charge conjugation transformation C^, the currents J(x), Jμ(x) and Jμν(x) have the properties,

C^J(x)C^1=+J(x),C^J±,μ(x)C^1=±J±,μ(x),C^J±,μν(x)C^1=±J±,μν(x).

At the hadron side, we obtain the hadronic representation and isolate the ground state hidden-charm molecule contributions [82, 83],

Π(p)=λZ+2MZ+2p2+=Π+(p2),Πμν(p)=λZ+2MZ+2p2(gμν+pμpνp2)+=Π+(p2)(gμν+pμpνp2)+,

Π,μναβ(p)=λ~Z+2MZ+2p2(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+λ~Z2MZ2p2(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+=Π~+(p2)(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+Π~(p2)(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ),

Π+,μναβ(p)=λZ+2MZ+2p2(g~μαg~νβ+g~μβg~να2g~μνg~αβ3)+=Π+(p2)(g~μαg~νβ+g~μβg~να2g~μνg~αβ3)+,

where the Z represents the molecular states Zc, Xc, Zcs, etc. We add the subscripts ± in the hidden-charm molecular states Z± and the components Π±(p2) and Π~±(p2) to represent the positive and negative parity contributions, respectively, and define the pole residues λZ± analogous to Eq. (127) with λZ±=MZ±λ~Z±. We choose the components Π+(p2) and p2Π~+(p2) to study the scalar, axialvector and tensor hidden-charm molecular states. According to the discussions in Section 2.3, we have no reason to worry about the contaminations from the two-meson scattering states.

At the QCD side, we carry out the operator product expansion up to the vacuum condensates of dimension 10, and take account of the vacuum condensates q¯q, αsGGπ, q¯gsσGq, q¯q2, gs2q¯q2, q¯qαsGGπ, q¯qq¯gsσGq, q¯gsσGq2 and q¯q2αsGGπ, where q=u, d or s quark, just like in Section 3.1.1.

According to analogous routine of Section 3.1.1, we obtain the QCD sum rules:

λZ+2exp(MZ+2T2)=4mc2s0dsρQCD(s)exp(sT2),

and

MZ+2=4mc2s0dsddτρQCD(s)exp(τs)4mc2s0dsρQCD(s)exp(τs)τ=1T2.

In the heavy quark limit, we describe the QQ¯qq¯ systems by a double-well potential model, the heavy quark Q (Q¯) serves as a static well potential and attracts the light antiquark q¯ (q) to form a color-neutral cluster. We introduce the effective heavy quark mass MQ and divide the molecular states into both the heavy and light degrees of freedom, i.e., 2MQ and μ=MX/Y/Z2(2MQ)2, respectively.

Analysis of the J/ψ and Υ mass spectrum with the famous Cornell potential or Coulomb-potential-plus-linear-potential leads to the constituent quark masses mc=1.84GeV and mb=5.17GeV [640], we can set the effective c-quark mass as the constituent quark mass Mc=mc=1.84GeV. The old value Mc=1.84GeV and updated value Mc=1.85GeV, which are fitted for the hidden-charm molecular states, are all consistent with the constituent quark mass mc=1.84GeV [81, 422, 641]. We choose the value Mc=1.84GeV to determine the ideal energy scales of the QCD spectral densities, and add an uncertainty δμ=±0.1GeV to account for the difference between the values Mc=1.84GeV and 1.85GeV. Furthermore, we take the modified energy scale formula

μ=MX/Y/Z2(2Mc)2kMs,

with k=0, 1, 2 and Ms=0.2GeV to account for the light flavor SU(3) breaking effects.

After trial and error, we obtain the Borel windows, continuum threshold parameters, energy scales of the QCD spectral densities, pole contributions, and contributions of the vacuum condensates of dimension 10, which are shown explicitly in Tab.42. At the hadron side, the pole contributions are about (40−60)%, while the central values are larger than 50%, the pole dominance condition is well satisfied. At the QCD side, the contributions of the vacuum condensates of dimension 10 are |D(10)|1% or 1%, the convergent behaviors of the operator product expansion are very good.

We take account of all the uncertainties of the relevant parameters, and obtain the masses and pole residues of the molecular states without strange, with strange and with hidden-strange, which are shown explicitly in Tab.43. From Tab.42 and Tab.43, it is obvious that the modified energy scale formula in Eq. (235) is well satisfied [82, 83].

In Fig.22, we plot the masses of the axialvector molecular states DD¯+DD¯, DD¯DD¯, DD¯s+DD¯s and DD¯ with variations of the Borel parameters at much larger ranges than the Borel widows as an example. From the figure, we can see plainly that there appear very flat platforms in the Borel windows indeed, where the regions between the two short vertical lines are the Borel windows.

In Fig.22, we also present the experimental values of the masses of the Zc(3900), Xc(3872), Zcs(3985) and Zc(4020) [169, 642], the predicted masses are in excellent agreement with the experimental data. The calculations support assigning the Zc(3900), Xc(3872), Zcs(3985) and Zc(4020) to be the DD¯+DD¯, DD¯DD¯, DD¯s+DD¯s and DD¯ tetraquark molecular states with the quantum numbers JPC=1+, 1++, 1+ and 1+, respectively. In Tab.44, we present the possible assignments of the ground state hidden-charm molecular states. However, the lattice QCD calculations do not favor the existence of the Zc(3900) [357, 370, 643, 644].

We could reproduce the experimental masses of the Xc(3872), Zc(3900), Zcs(3985), Zcs(4123) and Zc(4020) both in the scenarios of tetraquark states and molecular states, see Tab.9-Tab.11 and Tab.44, the tetraquark scenario can accommodate much more exotic X and Z states than the molecule scenario. Even in the tetraquark scenario, there are no rooms to accommodate the X(3940), X(4160), Zc(4100) and Zc(4200) without resorting to fine tuning. The X(3940) and X(4160) might be the conventional ηc(3S) and ηc(4S) states with the JPC=0+, respectively. The Zc(4100) might be a mixing scalar tetraquark state with the JPC=0++, and the Zc(4200) might be an axialvector color 88 type tetraquark state with the JPC=1+. For detailed discussions about this subject, one can consult Ref. [61].

The Zc(3900) and Zc(3885) have almost degenerated masses but quite different decay widths [162, 163, 167], they are taken as the same particle by the Particle Data Group [97] (also Ref. [60]), however, it is difficult to explain the large ratio,

Rexp=Γ(Zc(3885)DD¯)Γ(Zc(3900)J/ψπ)=6.2±1.1±2.7,

from the BESIII Collaboration [167]. If we assign the Zc(3900) as the 3¯3 type tetraquark state, and assign the Zc(3885) as the DD¯+DD¯ molecular state, it is easy to explain the large ratio Rexp.

The Zcs(3985) observed by the BESIII Collaboration near the DsD0 and DsD0 thresholds in the K+ recoil-mass spectrum in the processes e+eK+(DsD0+DsD0), its Breit-Wigner mass and width are 3985.22.0+2.1±1.7MeV and 13.85.2+8.1±4.9MeV respectively with an assignment of the spin-parity JP=1+ [169]. The Zcs(4000) observed in the J/ψK+ mass spectrum by the LHCb Collaboration has a mass of 4003±614+4MeV, a width of 131±15±26MeV, and the spin-parity JP=1+ [115]. The Zcs(3885) and Zcs(4000) are two quite different particles, although they have almost degenerated masses.

In the J/ψK+ mass spectrum from the LHCb Collaboration, there is a hint of a dip at the energy about 4.1GeV [115], which maybe due to the Zcs(4123) observed by the BESIII Collaboration with a mass about (4123.5±0.7±4.7)MeV [170]. More experimental data are still needed to obtain a precise resolution.

After Ref. [82] was published, the Belle Collaboration observed weak evidences for two structures in the γψ(2S) invariant mass spectrum in the two-photon process γγγψ(2S), one at 3922.4±6.5±2.0MeV with a width 22±17±4MeV, and another at 4014.3±4.0±1.5MeV with a width 4±11±6MeV [645]. The first structure is consistent with the X(3915) or χc2(3930), the second one might be an exotic charmonium-like state. We present its possible assignment in Tab.44, see X2(4014), which cannot be accommodated in the scenario of tetraquark state.

In Tab.43, the central values of the predicted molecule masses lie at the thresholds of the corresponding two-meson scattering states, where we have taken the currents having two color-neutral clusters, in each cluster, the constituents q and Q¯ (or Q and q¯) are in relative S-wave, see Eq. (226).

Now let us see the possible outcomes, if one of the color-neutral clusters has a relative P-wave between the constituents q and Q¯ (or Q and q¯), as one of the possible assignments of the Y(4260) is the DD¯1 molecular state with the JPC=1 [227, 229, 280, 555, 556].

For example, in the isospin limit, we write the valence quarks of the DD¯1(2420) and DD¯0(2400) molecular states symbolically as

u¯dc¯c,u¯ud¯d2c¯c,d¯uc¯c,u¯u+d¯d2c¯c,

the isospin triplet and singlet have degenerate masses. We take the isospin limit to study those molecular states [641].

Again we resort to the correlation functions Πμν(p) in Eq. (125) with the currents Jμ(x)=Jμ1(x), Jμ2(x), Jμ3(x) and Jμ4(x) to study the DD¯1(2420) and DD¯0(2400) molecular states, where

Jμ1/2(x)=12{u¯(x)iγ5c(x)c¯(x)γμγ5d(x)u¯(x)γμγ5c(x)c¯(x)iγ5d(x)},Jμ3/4(x)=12{u¯(x)c(x)c¯(x)γμd(x)±u¯(x)γμc(x)c¯(x)d(x)}.

Under charge conjugation transformation C^, the currents Jμ(x) have the properties,

C^Jμ1/3(x)C^1=Jμ1/3(x)|ud,C^Jμ2/4(x)C^1=+Jμ2/4(x)|ud.

According to the quark-hadron duality, we isolate the ground state contributions of the vector molecular states,

Πμν(p)=λY2MY2p2(gμν+pμpνp2)+,

where the λY are the pole residues.

We carry out the operator product expansion up to the vacuum condensates of dimension 10 in a consistent way routinely [82, 83, 641], and obtain the QCD sum rules for the masses and pole residues.

We adopt the energy scale formula in Eq. (235) to choose the best energy scales of the QCD spectral densities, and take the updated value Mc=1.85GeV [641].

After trial and error, we obtain the Borel parameters, continuum threshold parameters, pole contributions and energy scales, see Tab.45, where the central values of the pole contributions are larger than 50%, the pole dominance is well satisfied. In the Borel windows, the contributions D101%, the operator product expansion is well convergent.

We take account of all the uncertainties of the relevant parameters, and obtain the values of the masses and pole residues, which are also shown in Tab.45 [641].

The prediction MDD¯1(1)=4.36±0.08GeV is consistent with the experimental data MY(4390)=4391.6±6.3±1.0MeV from the BESIII Collaboration within uncertainties [153], while the predictions MDD¯1(1+)=4.60±0.08GeV, MDD¯0(1)=4.78±0.07GeV and MDD¯0(1+)=4.73±0.07GeV are much larger than upper bound of the experimental data MY(4220)=4218.4±4.0±0.9MeV and MY(4390)=4391.6±6.3±1.0MeV [153], moreover, they are much larger than the near thresholds MD+D1(2420)=4293MeV, MD0D1(2420)0=4285MeV, MD+D0(2400)=4361MeV, MD0D0(2400)0=4325MeV [646]. The present predictions only support assigning the Y(4390) to be the DD¯1(1) molecular state. Or the Y(4260) has sizable non DD¯1 component [647].

In Ref. [434], Zhang and Huang studied the Qq¯Q¯q type scalar, vector and axialvector molecular states with the QCD sum rules systematically by calculating the operator product expansion up to the vacuum condensates of dimension 6. The predicted molecule masses MDD¯0=4.26±0.07GeV and MDD¯1=4.34±0.07GeV are consistent with the Y(4220) and Y(4390), respectively. However, they do not distinguish the charge conjugation of the molecular states and neglect the higher dimensional vacuum condensates.

In Ref. [430], Lee, Morita and Nielsen distinguished the charge conjugation, and calculated the operator product expansion up to the vacuum condensates of dimension 6 including dimension 8 partly. They obtained the mass of the DD¯1(2420) molecular state with the JPC=1+, MDD¯1=4.19±0.22GeV, which differs from the prediction MDD¯1=4.34±0.07GeV significantly [434].

In Refs. [430, 434], Scheme II is chosen, some higher dimensional vacuum condensates are neglected, which are associated with 1T2, 1T4, 1T6 in the QCD spectral densities and manifest themselves at small values of the T2, thus we have to choose large values of T2 to warrant convergence of the operator product expansion. The higher dimensional vacuum condensates, see Fig.23, play an important role in determining the Borel windows therefore the ground state masses and pole residues, we should take them into account consistently. All in all, we observe that the QCD sum rules favor much larger masses than the two-meson thresholds if there exists a P-wave in one constituent, and we should bear in mind that the continuum threshold parameters should not be large enough to include contaminations from the higher resonances [648], i.e., if a bound state really exists, the continuum threshold s0 should be less than Mground+0.7GeV in the case of the hidden-charm four-quark systems.

Besides the Y(4260), it is also difficult to reproduce the mass of the Y(4660) with a cq¯qc¯ type current, however, we can reproduce its mass with a cc¯qq¯ type current [482, 485], i.e., it might be a ψf0(980) bound state [482, 485, 649, 650].

With the simple replacement cb, we obtain the corresponding QCD sum rules for the hidden-bottom tetraquark molecular states from Eq. (234). We can reproduce the experimental masses of the Zb(10610) and Zb(10650) as the BB¯BB¯ and BB¯ molecular states respectively with the JPC=1+ via the QCD sum rules [81, 422, 443, 445, 451], although those QCD sum rules suffer from shortcomings in one way or the other. It is more easy to reproduce a weak bound state if there exist weak attractive interactions between the two constituents [76, 211, 271, 274, 282, 651-655].

In Ref. [557], Qiao assigned the Y(4260) as the Λ¯cΛc baryonium state with the JPC=1. In Ref. [558], Qiao included the Σc0 constituent, and suggests a triplet and a singlet baryonium states,

B1+|Λc+Σ¯c0,Triplet:B1012(|Λc+Λ¯c|Σc0Σ¯c0),B1|ΛcΣc0,

and

Singlet:B0012(|Λc+Λ¯c+|Σc0Σ¯c0),

then he assigns the Zc±(4430) as the 2S B1± state with the JPC=1+, and the Y(4360) (Y(4660)) as the 2S Λ¯cΛc (Σ¯cΣc) baryonium state with the JPC=1.

We study the Λ¯cΛc and Σ¯cΣc type hidden-charm baryonium states via the QCD sum rules consistently by carrying out the operator product expansion up to the vacuum condensates of dimension 16 according to the counting roles in Sections 2.2 and 3.1.1, and observe that the Λ¯cΛc state with the JP=1 and Σ¯cΣc states with the JP=0, 1 lie at the corresponding baryon-antibaryon thresholds, respectively, while the Λ¯cΛc states with the JP=0, 1+ and Σ¯cΣc states with the JP=0+, 1+ lie above the corresponding baryon-antibaryon thresholds, respectively, whose masses are all much larger than the Y(4260) [656]. In Ref. [657], Wan, Tang and Qiao study the Λ¯QΛQ states with the JPC=0++, 0+, 1++ and 1 via the QCD sum rules by taking account of the vacuum condensates up to dimension 12, and observe that only the baryonium states with the JPC=0++ and 1 could exist, they also obtain masses much larger than the corresponding Λ¯QΛQ thresholds, respectively.

In Ref. [658], we construct the color singlet-singlet type six-quark current,

J(x)=J¯cc(x)iγ5Jcc(x),Jcc(x)=εijkciT(x)Cγαcj(x)γαγ5qk(x),

with q=u or d, to study the Ξ¯ccΞcc hexaquark molecular state by calculating the vacuum condensates up to dimension 14, the predicted mass MX7.2GeV supports assigning the X(7200) to be the Ξ¯ccΞcc hexaquark molecular state with the JPC=0+. However, the pole contribution is not larger than 40%, which would weaken the predictive ability.

In Ref. [659], we construct the color singlet-singlet-singlet type six-quark current with the I(JP)=32(1) to study the DD¯K system via the QCD sum rules by considering the contributions of the vacuum condensates up to dimension-16, and observe that there indeed exists a resonance state which lies above the DD¯K threshold, and suggest to search for it in the J/ψπK mass spectrum.

Such subjects need further studies to obtain definite conclusion, at the present time, they are open problems.

4.2 Doubly heavy tetraquark states

Before and after the observation of the doubly-charmed tetraquark candidate Tcc+(3875), especially after the observation, there have been many works on the doubly-charmed tetraquark (molecular) states with different theoretical approaches [371-373, 375, 471, 472, 513, 660-669].

If we perform Fierz rearrangements for the four-quark axialvector current Jμ(x) [576], see Eq. (194), we obtain a special superposition of the color singlet-singlet type currents,

Jμ(x)=εijkεimnQjT(x)CγμQk(x)u¯m(x)γ5Cd¯nT(x)=i2[u¯iγ5Qd¯γμQd¯iγ5Qu¯γμQ]+12[u¯Qd¯γμγ5Qd¯Qu¯γμγ5Q]i2[u¯σμνγ5Qd¯γνQd¯σμνγ5Qu¯γνQ]+i2[u¯σμνQd¯γνγ5Qd¯σμνQu¯γνγ5Q]=i2Jμ1(x)+12Jμ2(x)i2Jμ3(x)+i2Jμ4(x).

The currents Jμ1(x), Jμ2(x), Jμ3(x) and Jμ4(x) couple potentially to the color singlet-singlet type tetraquark states or two-meson scattering states. In fact, there exist spatial separations between the diquark and antidiquark pair, the currents Jμ(x) should be modified to Jμ(x,ϵ),

Jμ(x,ϵ)=εijkεimnQjT(x)CγμQk(x)u¯m(x+ϵ)γ5Cd¯nT(x+ϵ),

where the four-vector ϵα=(0,ϵ). The spatial distance between the diquark and antidiquark pair maybe frustrate the Fierz rearrangements or recombination, although we usually take the local limit ϵ0, we should not take it for granted that the Fierz rearrangements are feasible [401], we cannot obtain the conclusion that the Tcc+ has the DDDD Fock component according to the component q¯iγ5cq¯γμc in Eq. (244). According to the predictions in Tab.34, we can only obtain the conclusion tentatively that the Tcc+ has a diquark−antidiquark type tetraquark Fock component with the spin-parity JP=1+ and isospin I=0 [576]. It is interesting to explore whether or not there exists a color 11 type Fock component indeed.

Again, we resort to the correlation functions Π(p), Πμν(p) and Πμναβ(p) defined in Eq. (125), and construct the currents J(x), Jμ(x) and Jμν(x),

J(x)=JDD(x),JDDs(x),JDsDs(x),JDD(x),JDDs(x),JDsDs(x),

Jμ(x)=JDD,L,μ(x),JDD,H,μ(x),JDDs,L,μ(x),JDDs,H,μ(x),JDsDs,μ(x),JD1D0,L,μ(x),JD1D0,H,μ(x),JDs1D0,L,μ(x),JDs1D0,H,μ(x),JDs1Ds0,μ(x),

Jμν(x)=JDD,L,μν(x),JDD,H,μν(x),JDDs,L,μν(x),JDDs,H,μν(x),JDsDs,L,μν(x),JDsDs,H,μν(x),

JDD(x)=u¯(x)iγ5c(x)d¯(x)iγ5c(x),JDDs(x)=q¯(x)iγ5c(x)s¯(x)iγ5c(x),JDsDs(x)=s¯(x)iγ5c(x)s¯(x)iγ5c(x),JDD(x)=u¯(x)γμc(x)d¯(x)γμc(x),JDDs(x)=q¯(x)γμc(x)s¯(x)γμc(x),JDsDs(x)=s¯(x)γμc(x)s¯(x)γμc(x),

JDD,L/H,μ(x)=12[u¯(x)iγ5c(x)d¯(x)γμc(x)u¯(x)γμc(x)d¯(x)iγ5c(x)],JDDs,L/H,μ(x)=12[q¯(x)iγ5c(x)s¯(x)γμc(x)q¯(x)γμc(x)s¯(x)iγ5c(x)],JDsDs,μ(x)=s¯(x)iγ5c(x)s¯(x)γμc(x),

JD1D0,L/H,μ(x)=12[u¯(x)c(x)d¯(x)γμγ5c(x)±u¯(x)γμγ5c(x)d¯(x)c(x)],JDs1D0,L/H,μ(x)=12[q¯(x)c(x)s¯(x)γμγ5c(x)±q¯(x)γμγ5c(x)s¯(x)c(x)],JDs1Ds0,μ(x)=s¯(x)c(x)s¯(x)γμγ5c(x),

JDD,L/H,μν(x)=12[u¯(x)γμc(x)d¯(x)γνc(x)u¯(x)γνc(x)d¯(x)γμc(x)],JDDs,L/H,μν(x)=12[q¯(x)γμc(x)s¯(x)γνc(x)q¯(x)γνc(x)s¯(x)γμc(x)],JDsDs,L/H,μν(x)=12[s¯(x)γμc(x)s¯(x)γνc(x)s¯(x)γνc(x)s¯(x)γμc(x)],

and q=u, d, the subscripts DD, DD, and DsDs stand for the two color-neutral clusters, we add the subscripts L and H to distinguish the lighter and heavier states in the same doublet due to the mixing effects, as direct calculations indicate that there exists such a tendency.

Under parity transformation P^, the currents have the properties,

P^J(x)P^1=+J(x~),P^Jμ(x)P^1=Jμ(x~),P^Jμν(x)P^1=+Jμν(x~),

where xμ=(t,x) and x~μ=(t,x). We rewrite Eq. (253) in more explicit form,

P^Ji(x)P^1=+Ji(x~),P^Jij(x)P^1=+Jij(x~),

P^J0(x)P^1=J0(x~),P^J0i(x)P^1=J0i(x~),

where the space indexes i, j=1, 2, 3. There are both positive and negative components, and they couple potentially to the axialvector/tensor and pseudoscalar/vector molecular states, respectively. We will introduce an superscript to denote the negative parity.

According to the quark−hadron duality, we obtain the hadronic representation and isolate the ground state contributions of the scalar, axialvector and tensor molecular states [83, 513],

Π(p)=λT2MT2p2+=ΠT0(p2)+,

Πμν(p)=λT2MT2p2(gμν+pμpνp2)+=ΠT1(p2)(gμν+pμpνp2)+,

ΠL,μναβ(p)=λ~T2MT2p2(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+λ~T2MT2p2(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+=Π~T1(p2)(p2gμαgνβp2gμβgναgμαpνpβgνβpμpα+gμβpνpα+gναpμpβ)+Π~T1,(p2)(gμαpνpβgνβpμpα+gμβpνpα+gναpμpβ),

ΠH,μναβ(p)=λT2MT2p2(g~μαg~νβ+g~μβg~να2g~μνg~αβ3)+=ΠT2(p2)(g~μαg~νβ+g~μβg~να2g~μνg~αβ3)+,

where the pole residues λT and λT are defined analogous to Eq. (127) with the simple replacements Z+T and ZT, and λT=λ~TMT and λT=λ~TMT, we introduce the superscripts 0, 1 and 2 denote the spins of the molecular states. Thereafter, we choose the components ΠT0/2(p2) and p2Π~T1(p2).

We accomplish the operator product expansion up to the vacuum condensates of dimension 10 and take account of the vacuum condensates q¯q, αsGGπ, q¯gsσGq, q¯q2, q¯qαsGGπ, q¯qq¯gsσGq, q¯gsσGq2 and q¯q2αsGGπ with the vacuum saturation in a consistent way [670], where q=u, d or s [60, 61, 82, 83, 171, 172, 418, 425, 513-515]. Again, we neglect the u and d quark masses, and take account of the terms ms according to the light-flavor SU(3) breaking effects.

Then we match the hadron side with the QCD side below the continuum thresholds s0, perform the Borel transformation, and obtain the QCD sum rules for the masses and pole residues. With a simple replacement cb, we obtain the corresponding QCD sum rules for the doubly-bottom tetraquark states, the calculations are straightforward.

We take the modified energy scale formula shown in Eq. (235) and the effective c/s-quark masses Mc=1.82GeV and Ms=0.2GeV to determine the optimal energy scales of the QCD spectral densities. After trial and error, we obtain the Borel parameters, continuum threshold parameters, energy scales of the QCD spectral densities, pole contributions and contributions of the vacuum condensates of dimension 10, which are shown in Tab.46 [83, 513]. The pole contributions are about (40−60)% at the hadron side, the central values are larger than 50%, the pole dominance is satisfied very good. Moreover, the contributions of the vacuum condensates of dimension 10 are |D(10)|<1% or 1% at the QCD side, the operator product expansion converges very good [83, 513].

We take account of all the uncertainties of the relevant parameters, and obtain the masses and pole residues of the doubly-charmed molecular states without strange, with strange and with doubly-strange, which are presented explicitly in Tab.47 [83, 513].

In Fig.24, as an example, we plot the masses of the axialvector (DDDD)L and (DD+DD)H molecular states with variations of the Borel parameters at much larger ranges than the Borel widows. There appear very flat platforms in the Borel windows, the regions between the two short perpendicular lines.

There exist both a lighter and a heavier state for the ccu¯d¯ and ccq¯s¯ molecular states, the lighter state (DDDD)L with the isospin (I,I3)=(0,0) has a mass 3.88±0.11GeV, which is in excellent agreement with the mass of the doubly-charmed tetraquark candidate Tcc+ from the LHCb Collaboration [194, 195], and supports assigning the Tcc+ to be the (DDDD)L molecular state, as the Tcc+ has the isospin I=0. In other words, the exotic state Tcc+ maybe have a (DDDD)L Fock component. The heavier state (DD+DD)H with the isospin (I,I3)=(1,0) has a mass 3.90±0.11GeV, the central value lies slightly above the DD threshold, the strong decays to the final states DDπ are kinematically allowed but with small phase-space. If we choose the same input parameters, the DD molecular state with the isospin I=1 has slightly larger mass than the corresponding molecule with the isospin I=0, it is indeed that the isoscalar DD molecular state is lighter.

For the (D0D1+D1D0)L, (D0D1D1D0)H, (D0Ds1+Ds0D1)L and (D0Ds1Ds0D1)H molecular states, there exists a P-wave in the color-singlet constituents, the P-wave is embodied implicitly in the underlined γ5_ in the scalar currents q¯iγ5γ5_c, s¯iγ5γ5_c and axialvector currents q¯γμγ5_c, s¯γμγ5_c, as multiplying γ5 to the pseudoscalar currents q¯iγ5c, s¯iγ5c and vector currents q¯γμc, s¯γμc changes their parity. We should introduce the spin-orbit interactions to account for the large mass gaps between the lighter and heavier states (L,H), i.e., ((D0D1+D1D0)L,(D0D1D1D0)H), ((D0Ds1+Ds0D1)L,(D0Ds1Ds0D1)H) [83, 513].

From Tab.47, we can see explicitly that the D0D1D1D0, D0D1+D1D0, D0Ds1Ds0D1, D0Ds1+Ds0D1 and Ds1Ds0 molecular states have much larger masses than the corresponding two-meson thresholds, just like in the case of the hidden-charm molecular states studied in the previous sub-section, where there exist a relative P-wave in one of the color-neutral clusters.

Beyond those color 11 tetraquark states, there maybe also exist some corresponding hexaquark states. The QCD sum rules indicate that there exist the ΛcΛc, ΣcΣc and ΞccΣc dibaryon states [413, 656, 657], the 3¯3¯3¯ type triply-charmed hexaquark states [671], and the color 111 type triply-charmed hexaquark molecular states [426]. However, for the light dibaryon/baryonium states, we can only obtain very small pole contributions at the hadron side or bad convergent behaviors of the operator product expansion at the QCD side [205, 672, 673], which weakens the predictive ability. If we adopt the truncation rule in Sections 2.2 and 3.1.1, i.e., each heavy quark line emits a gluon and each light quark contributes a quark−antiquark pair, which leads to a quark−gluon operator to reach the highest dimensional vacuum condensates, the two basic criteria of the QCD sum rules are difficult to satisfy. Such subjects need further studies.

On the other hand, Lattice calculations indicate that there maybe exist the NΩ, ΩΩ, Ω(ccc)Ω(ccc), Ω(bbb)Ω(bbb), dibaryon states [674-679]. While the heavy-antiquark−diquark symmetry implies that there exists a model-independent relation between the spin-splitting in the masses of the hidden-charm pentaquark states and corresponding splitting for the triply-charmed dibaryon states [680].

5 Hidden heavy pentaquark states

If a baryon current J(x) has the spin-parity JP=12+, then the current iγ5J(x) would have the spin-parity JP=12, as multiplying iγ5 changes the parity of the current J(x) [681]. In 1993, Bagan et al. [682] took the infinite heavy quark limit to separate the contributions of the positive and negative parity heavy baryon states unambiguously. In 1996, Jido et al. [683] introduced a novel approach to separate the contributions of the negative-parity light-flavor baryons from the positive-parity ones.

At first, we write down the correlation functions Π±(p),

Π±(p)=id4xeipx0|T{J±(x)J¯±(0)}|0,

where we add the subscripts ± to denote the positive and negative parity, respectively, J=iγ5J+. We decompose the correlation functions Π±(p),

Π±(p)=pΠ1(p2)±Π0(p2),

according to Lorentz covariance, because

Π(p)=γ5Π+(p)γ5.

The currents J+ couple potentially to both the positive- and negative-parity baryons [681],

0|J+|B±B±|J¯+|0=γ50|J|B±B±|J¯|0γ5,

where the B± denote the positive and negative parity baryons, respectively.

According to the relation in Eq. (260), we obtain the hadronic representation [683],

Π+(p)=λ+2p+M+M+2p2+λ2pMM2p2+,

where the M± are the baryon masses, the pole residues are defined by 0|J±(0)|B±(p)=λ±U±, the U± are the Dirac spinors.

If we take p=0, then

limϵ0ImΠ+(p0+iϵ)π=λ+2γ0+12δ(p0M+)+λ2γ012δ(p0M)+=γ0A(p0)+B(p0)+,

where

A(p0)=12[λ+2δ(p0M+)+λ2δ(p0M)],B(p0)=12[λ+2δ(p0M+)λ2δ(p0M)],

the contribution A(p0)+B(p0) (A(p0)B(p0)) contains contributions from the positive parity (negative parity) states only.

We carry out the operator product expansion at large P2=p02 region, then use the dispersion relation to obtain the spectral densities ρA(p0) and ρB(p0) at the quark-gluon level. At last, we introduce the weight functions exp[p02T2], p02exp[p02T2], and obtain the QCD sum rules,

Δs0dp0[A(p0)+B(p0)]exp[p02T2]=Δs0dp0[ρA(p0)+ρB(p0)]exp[p02T2],

Δs0dp0[A(p0)+B(p0)]p02exp[p02T2]=Δs0dp0[ρA(p0)+ρB(p0)]p02exp[p02T2],

where the Δ and s0 are the threshold and continuum threshold respectively, the T2 is the Borel parameter [683]. Thereafter, such semi-analytical method was applied to study the heavy, doubly-heavy and triply-heavy baryons with the JP=12± and 32± [684-689].

As the procedure introduced in Ref. [683] is semi-analytical, in 2016, we suggested an analytical procedure to study the pentaquark states [423]. For a correlation function Π(p2)=Π(p), at the hadron side, we obtain the spectral densities through the dispersion relation,

ImΠ(s)π=p[λ2δ(sM2)+λ+2δ(sM+2)]+[Mλ2δ(sM2)M+λ+2δ(sM+2)]=pρH1(s)+ρH0(s),

where the subscript H denotes the hadron side, then we introduce the weight function exp(sT2) to obtain the QCD sum rules at the hadron side,

Δ2s0ds[sρH1(s)±ρH0(s)]exp(sT2)=2Mλ2exp(M2T2),

where the s0 are the continuum threshold parameters. We separate the contributions of the negative parity pentaquark states from that of the positive parity ones unambiguously [423]. The calculations at the QCD side are analytical as we do not set p=0.

For the early works on the pentaquark states, one can consult the QCD sum rules on the Θ(1540), where the currents JZ(x) [690], JO(x) [691] and JN(x) [692],

JZ(x)=12εijkuiT(x)Cγ5dj(x){um(x)s¯m(x)iγ5dk(x)(ud)},JO(x)=εijkεlmnεknbuiT(x)Cdj(x)ulT(x)Cγ5dm(x)Cs¯bT(x),JN(x)=cosθ2εijkuiT(x)Cγ5dj(x)ukT(x)Cγ5dm(x)Cs¯mT(x)(ud+sinθ2εijkuiT(x)Cdj(x)ukT(x)Cdm(x)Cs¯mT(x)(ud),

were constructed to interpolate it. The current JO(x) was studied in the semi-analytical method [691].

5.1 3¯3¯3¯ type pentaquark states

Now, let us turn to the pentaquark states completely and write down the correlation functions Π(p), Πμν(p) and Πμναβ(p),

Π(p)=id4xeipx0|T{J(x)J¯(0)}|0,Πμν(p)=id4xeipx0|T{Jμ(x)J¯ν(0)}|0,Πμναβ(p)=id4xeipx0|T{Jμν(x)J¯αβ(0)}|0,

where

J(x)=J1(x),J2(x),J3(x),J4(x),Jμ(x)=Jμ1(x),Jμ2(x),Jμ3(x),Jμ4(x),Jμν(x)=Jμν1(x),Jμν2(x),

J1(x)=εilaεijkεlmnujT(x)Cγ5dk(x)umT(x)Cγ5cn(x)Cc¯aT(x),J2(x)=εilaεijkεlmnujT(x)Cγ5dk(x)umT(x)Cγμcn(x)γ5γμCc¯aT(x),J3(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)dmT(x)Cγ5cn(x)+2ujT(x)Cγμdk(x)umT(x)Cγ5cn(x)]γ5γμCc¯aT(x),J4(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)dmT(x)Cγμcn(x)+2ujT(x)Cγμdk(x)umT(x)Cγμcn(x)]Cc¯aT(x),

Jμ1(x)=εilaεijkεlmnujT(x)Cγ5dk(x)umT(x)Cγμcn(x)Cc¯aT(x),Jμ2(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)dmT(x)Cγ5cn(x)+2ujT(x)Cγμdk(x)umT(x)Cγ5cn(x)]Cc¯aT(x),Jμ3(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)dmT(x)Cγαcn(x)+2ujT(x)Cγμdk(x)umT(x)Cγαcn(x)]γ5γαCc¯aT(x),

Jμ4(x)=εilaεijkεlmn3[ujT(x)Cγαuk(x)dmT(x)Cγμcn(x)+2ujT(x)Cγαdk(x)umT(x)Cγμcn(x)]γ5γαCc¯aT(x),

Jμν1(x)=εilaεijkεlmn6[ujT(x)Cγμuk(x)dmT(x)Cγνcn(x)+2ujT(x)Cγμdk(x)umT(x)Cγνcn(x)]Cc¯aT(x)+(μν),Jμν2(x)=12εilaεijkεlmnujT(x)Cγ5dk(x)[umT(x)Cγμcn(x)γ5γνCc¯aT(x)+umT(x)Cγνcn(x)γ5γμCc¯aT(x)],

we choose the Cγ5 and Cγμ diquarks in the color 3¯, the most stable diquark configurations, as the basic constituents to construct the diquark−diquark−antiquark type currents J(x), Jμ(x) and Jμν(x) with the spin-parity JP=12, 32 and 52, respectively, which are expected to couple potentially to the lowest pentaquark states [402, 423, 514].

In the currents J(x), Jμ(x) and Jμν(x), there are diquark constituents εijkujTCγ5dk, εijkujTCγμdk, εijkujTCγμuk, εijkqjTCγ5ck, εijkqjTCγμck with q=u, d. If we use the SL and SH to represent the spins of the light and heavy diquarks, respectively, the light diquarks εijkujTCγ5dk, εijkujTCγμdk and εijkujTCγμuk have the spins SL=0, 1 and 1, respectively, the heavy diquarks εijkqjTCγ5ck and εijkqjTCγμck have the spins SH=0 and 1, respectively. A light diquark and a heavy diquark form a charmed tetraquark in the color 3 with the angular momentum JLH=SL+SH, which has the values JLH=0, 1 or 2. The c¯-quark operator Cc¯aT has the spin-parity JP=12, the c¯-quark operator γ5γμCc¯aT has the spin-parity JP=32 due to the factor γ5γμ. The total angular momentums are J=JLH+Jc¯ with the values J=12, 32 or 52, which are shown explicitly in Tab.48. In Tab.48, we present the quark structures of the currents explicitly [402]. For example, in the current Jμν2(x), there are a scalar diquark εijkujT(x)Cγ5dk(x) with the spin-parity JP=0+, an axialvector diquark εlmnumT(x)Cγμcn(x) with the spin-parity JP=1+, and an antiquark γ5γνCc¯aT(x) with the spin-parity JP=32, the total angular momentum is J=52. For more intuitive and simple diquark models for the pentaquark states, one can consult Refs. [693-698].

Although the currents J(x), Jμ(x) and Jμν(x) have negative parity, but they couple potentially to the pentaquark states with positive parity, as multiplying iγ5 to the currents J(x), Jμ(x) and Jμν(x) changes their parity [681-689, 699-701].

Now we write down the current-pentaquark couplings explicitly,

0|J(0)|P12(p)=λ12U(p,s),0|J(0)|P12+(p)=λ12+iγ5U+(p,s),

0|Jμ(0)|P32(p)=λ32Uμ(p,s),0|Jμ(0)|P32+(p)=λ32+iγ5Uμ+(p,s),0|Jμ(0)|P12+(p)=f12+pμU+(p,s),0|Jμ(0)|P12(p)=f12pμiγ5U(p,s),

0|Jμν(0)|P52(p)=2λ52Uμν(p,s),0|Jμν(0)|P52+(p)=2λ52+iγ5Uμν+(p,s),0|Jμν(0)|P32+(p)=f32+[pμUν+(p,s)+pνUμ+(p,s)],0|Jμν(0)|P32(p)=f32iγ5[pμUν(p,s)+pνUμ(p,s)],0|Jμν(0)|P12(p)=g12pμpνU(p,s),0|Jμν(0)|P12+(p)=g12+pμpνiγ5U+(p,s),

where the superscripts ± denote the positive parity and negative parity, respectively, the subscripts 12, 32 and 52 denote the spins of the pentaquark states, the λ, f and g are the pole residues.

The spinors U±(p,s) satisfy the Dirac equations (M±)U±(p)=0, while the spinors Uμ±(p,s) and Uμν±(p,s) satisfy the Rarita−Schwinger equations (M±)Uμ±(p)=0 and (M±)Uμν±(p)=0, and relations γμUμ±(p,s)=0, pμUμ±(p,s)=0, γμUμν±(p,s)=0, pμUμν±(p,s)=0, Uμν±(p,s)=Uνμ±(p,s), respectively [423, 565].

At the hadron side, we insert a complete set of intermediate pentaquark states with the same quantum numbers as the currents J(x), iγ5J(x), Jμ(x), iγ5Jμ(x), Jμν(x) and iγ5Jμν(x) into the correlation functions Π(p), Πμν(p) and Πμναβ(p) to obtain the hadronic representation, and isolate the lowest states of the hidden-charm pentaquark states with negative and positive parity [402, 423, 514]:

Π(p)=λ122p+MM2p2+λ12+2pM+M+2p2+=Π121(p2)p+Π120(p2),

Πμν(p)=λ322p+MM2p2(gμν+γμγν3+2pμpν3p2pμγνpνγμ3p2)+λ32+2pM+M+2p2(gμν+γμγν3+2pμpν3p2pμγνpνγμ3p2)+f12+2p+M+M+2p2pμpν+f122pMM2p2pμpν+=[Π321(p2)p+Π320(p2)](gμν)+,

Πμναβ(p)=2λ522p+MM2p2[g~μαg~νβ+g~μβg~να2g~μνg~αβ5110(γμγα+γμpαγαpμp2pμpαp2)g~νβ110(γνγα+γνpαγαpνp2pνpαp2)g~μβ+]

+2λ52+2pM+M+2p2[g~μαg~νβ+g~μβg~να2g~μνg~αβ5110(γμγα+γμpαγαpμp2pμpαp2)g~νβ110(γνγα+γνpαγαpνp2pνpαp2)g~μβ+]+f32+2p+M+M+2p2[pμpα(gνβ+γνγβ3+2pνpβ3p2pνγβpβγν3p2)+]+f322pMM2p2[pμpα(gνβ+γνγβ3+2pνpβ3p2pνγβpβγν3p2)+]+g122p+MM2p2pμpνpαpβ+g12+2pM+M+2p2pμpνpαpβ+=[Π521(p2)p+Π520(p2)](gμαgνβ+gμβgνα)+,

where we have used the summations of the spinors [702],

sUU¯=(p+M±),sUμU¯ν=(p+M±)(gμν+γμγν3+2pμpν3p2pμγνpνγμ3p2),sUμνU¯αβ=(p+M±){g~μαg~νβ+g~μβg~να2g~μνg~αβ5110(γμγα+γμpαγαpμp2pμpαp2)g~νβ110(γνγα+γνpαγαpνp2pνpαp2)g~μβ110(γμγβ+γμpβγβpμp2pμpβp2)g~να110(γνγβ+γνpβγβpνp2pνpβp2)g~μα},

and p2=M±2 on mass-shell. We study the components Π121(p2), Π120(p2), Π321(p2), Π320(p2), Π521(p2) and Π520(p2) to avoid possible contaminations from other pentaquark states with different spins.

We obtain the spectral densities at the hadron side through dispersion relation,

ImΠj1(s)π=λj2δ(sM2)+λj+2δ(sM+2)=ρj,H1(s),

ImΠj0(s)π=Mλj2δ(sM2)M+λj+2δ(sM+2)=ρj,H0(s),

where j=12, 32, 52, the subscript H denotes the hadron side, then we introduce the weight functions sexp(sT2) and exp(sT2) to obtain the QCD sum rules at the hadron side,

4mc2s0ds[sρj,H1(s)±ρj,H0(s)]exp(sT2)=2Mλj2exp(M2T2),

where the s0 are the continuum threshold parameters, and the T2 are the Borel parameters. We distinguish the contributions of the negative and positive parity pentaquark states unambiguously, and there is no contamination.

Now we briefly outline the operator product expansion. Firstly, we contract the u, d and c quark fields in the correlation functions Π(p), Πμν(p) and Πμναβ(p) with Wick theorem, for example,

Π(p)=iεilaεijkεlmnεilaεijkεlmnd4xeipx{Tr[γ5Dkk(x)γ5CUjjT(x)C]Tr[γ5Cnn(x)γ5CUmmT(x)C]CCaaT(x)C+Tr[γ5Dkk(x)γ5CUmjT(x)Cγ5Cnn(x)γ5CUjmT(x)C]CCaaT(x)C},

for the current J(x)=J1(x), where the Uij(x), Dij(x) and Cij(x) are the full u, d and c quark propagators, respectively, see Eqs. (129) and (130). Then we compute all the integrals in the coordinate and momentum spaces sequentially to obtain the Π(p), Πμν(p) and Πμναβ(p) at the quark-gluon level, and finally we obtain the QCD spectral densities through dispersion relation,

ρj,QCD1(s)=ImΠj1(s)π,ρj,QCD0(s)=ImΠj0(s)π,

where j=12, 32, 52. In computing the integrals, we draw up all the Feynman diagrams from Eq. (282) and calculate them one by one. In Eq. (282), there are two c-quark propagators and three light quark propagators, if each c-quark line emits a gluon and each light quark line contributes a quark−antiquark pair, we obtain an operator GμνGαβu¯uu¯ud¯d according to the counting role in Eq. (38), which is of dimension 13, see Fig.25. We should take account of the vacuum condensates at least up to dimension 13 in stead of dimension 10, which is adopted in most literatures. The vacuum condensates q¯q2q¯gsσGq, q¯qq¯gsσGq2, q¯q3αsπGG are of dimension 11 and 13 respectively, and come from the Feynman diagrams shown in Fig.25 [402]. Those vacuum condensates are associated with the 1T2, 1T4 and 1T6, which manifest themselves at the small T2 and play an important role in choosing the Borel windows. We take the truncations n13 and k1 in a consistent way, the quark-gluon operators of the orders O(αsk) with k>1 and dimension n>13 are discarded.

In Refs. [423, 703-706], we take the truncations n10 and k1 and discard the quark-gluon operators of the orders O(αsk) with k>1 and dimension n>10. Sometimes we also neglected the vacuum condensates αsGGπ, q¯qαsGGπ, s¯sαsGGπ, q¯q2αsGGπ, q¯qs¯sαsGGπ, s¯s2αsGGπ, which are not associated with the 1T2, 1T4 and 1T6 to manifest themselves at the small T2. Such an approximation would impair the predictive ability.

Then let us match the hadron side with the QCD side of the correlation functions, take the quark-hadron duality below the continuum thresholds s0, and obtain the QCD sum rules:

2Mλj2exp(M2T2)=4mc2s0dsρQCD,j(s)exp(sT2),

where ρQCD,j(s)=sρQCD,j1(s)±ρQCD,j0(s).

We derive Eq. (284) with respect to 1T2, then eliminate the pole residues λj and obtain the QCD sum rules for the masses of the hidden-charm pentaquark states,

M2=4mc2s0dsdd(1/T2)ρQCD,j(s)exp(sT2)4mc2s0dsρQCD,j(s)exp(sT2).

With a simple replacement cb, we obtain the corresponding QCD sum rules for the hidden-bottom pentaquark states.

According to the discussions in Section 2.4, we take the energy scale formula,

μ=MX/Y/Z/P2(2MQ)2,

to determine the best energy scales of the QCD spectral densities [402, 423, 703-706], and choose the updated value of the effective c-quark mass Mc=1.82GeV [512].

After trial and error, we obtain the Borel windows T2, continuum threshold parameters s0, ideal energy scales of the QCD spectral densities, pole contributions of the ground states and contributions of the vacuum condensates of dimension 13, which are shown explicitly in Tab.49 [402].

In Fig.26, we plot the contributions of the vacuum condensates of dimensions 11 and 13 with variation of the Borel parameter T2 for the hidden-charm pentaquark state [ud][uc]c¯ (0, 0, 0, 12) with the central values of the parameters shown in Tab.49 as an example. The vacuum condensates of dimension 13 manifest themselves at the region T2<2GeV2, we should choose the value T2>2GeV2. While the vacuum condensates of dimension 11 manifest themselves at the region T2<2.6GeV2, which requires a larger Borel parameter T2>2.6GeV2 to warrant the convergence of the operator product expansion. The higher dimensional vacuum condensates play an important role in choosing the Borel windows, where they play an minor important role as the operator product expansion should be convergent, we should take them into account in a consistent way, for example, the |D(13)| is less than 1% [402].

In Fig.27, we plot the mass of the hidden-charm pentaquark state [ud][uc]c¯ (0, 0, 0, 12) with variation of the Borel parameter T2 for truncations of the operator product expansion up to the vacuum condensates of dimensions 10 and 13, respectively [402]. The vacuum condensates of dimensions 11 and 13 play an important role to obtain stable QCD sum rules, we should take them into account.

In Tab.49, the pole contributions are about (40−60)% and the contributions of the vacuum condensates of dimension 13 are 1% or 1%, the pole dominance and convergence of the operator product expansion are all satisfied, the two basic criteria of the QCD sum rules are satisfied in the case of the hidden-charm pentaquark states for the first time.

We take account of all uncertainties of the relevant parameters, and obtain the masses and pole residues of the hidden-charm pentaquark states with negative parity, which are shown explicitly in Tab.50.

The predicted masses MP=4.31±0.11GeV for the ground state [ud][uc]c¯ (0, 0, 0, 12) pentaquark state and MP=4.34±0.14GeV for the ground state [uu][dc]c¯+2[ud][uc]c¯ (1, 1, 0, 12) pentaquark state are both in excellent agreement with the experimental data MP(4312)=4311.9±0.70.6+6.8MeV from the LHCb Collaboration [197], and support assigning the Pc(4312) to be the hidden-charm pentaquark state with the JP=12.

After the work was published [402], the LHCb Collaboration observed an evidence for a structure Pc(4337) in the J/ψp and J/ψp¯ systems with a significance about 3.13.7σ depending on the JP hypothesis [199], the Breit−Wigner mass and width are 43374+72+2MeV and 2912+2614+14MeV respectively. The Pc(4337) can be assigned as the ground state [uu][dc]c¯+2[ud][uc]c¯ (1, 1, 0, 12) pentaquark state with the mass 4.34±0.14GeV.

The predicted masses MP=4.45±0.11GeV for the ground state [ud][uc]c¯ (0, 1, 1, 12) pentaquark state, MP=4.46±0.11GeV for the ground state [uu][dc]c¯+2[ud][uc]c¯ (1, 0, 1, 12) pentaquark state and MP=4.39±0.11 for the ground state [ud][uc]c¯ (0, 1, 1, 32), [uu][dc]c¯+2[ud][uc]c¯ (1, 1, 2, 52), [ud][uc]c¯ (0, 1, 1, 52) pentaquark states are in excellent agreement (or compatible with) the experimental data MP(4440)=4440.3±1.34.7+4.1MeV from the LHCb Collaboration [197], and support assigning the Pc(4440) to be the hidden-charm pentaquark state with JP=12, 32 or 52.

The predicted masses MP=4.45±0.11GeV for the ground state [ud][uc]c¯ (0, 1, 1, 12) pentaquark state, MP=4.46±0.11GeV for the ground state [uu][dc]c¯+2[ud][uc]c¯ (1, 0, 1, 12) pentaquark state and MP=4.47±0.11GeV for the ground state [uu][dc]c¯+2[ud][uc]c¯ (1, 0, 1, 32) pentaquark states are in excellent agreement the experimental data MP(4457)=4457.3±0.61.7+4.1MeV from the LHCb Collaboration [197], and support assigning the Pc(4457) to be the hidden-charm pentaquark state with JP=12 or 32.

In Tab.50, we present the possible assignments of the Pc(4312), Pc(4440) and Pc(4457) explicitly as a summary. In Tab.51, we compare the present predictions with our previous calculations [423, 703-705], where the vacuum condensates of dimension 11 and 13 were neglected, sometimes the vacuum condensates αsGGπ, q¯qαsGGπ and q¯q2αsGGπ were also neglected. From the Tab.51, we can see that in some cases the predicted masses change remarkably, while in other cases the predicted masses change slightly. In Ref. [423], we construct the current γ5Jμν2(x) to interpolate the hidden-charm tetraquark state with the JP=52+, which should be updated.

In Ref. [704], we construct the Jq1q2q3jLjH(x) with the spin-parity JP=12 to study the hidden-charm pentaquark states with the JP=12± according to the rules,

1q1q2+0q3c+12c¯=12q1q2q3cc¯_32q1q2q3cc¯,

1q1q2+1q3c+12c¯=[0q1q2q3c+1q1q2q3c+2q1q2q3c+]12c¯=12q1q2q3cc¯_[12q1q2q3cc¯32q1q2q3cc¯][32q1q2q3cc¯52q1q2q3cc¯],

1q1q2+0q3c+[112c¯]=1q1q2+0q3c+[12c¯+32c¯+]=[12q1q2q3cc¯+_32q1q2q3cc¯+][12q1q2q3cc¯+32q1q2q3cc¯+52q1q2q3cc¯+],

1q1q2+1q3c+[112c¯]=[0q1q2q3c+1q1q2q3c+2q1q2q3c+][12c¯+32c¯+]=12q1q2q3cc¯+_[12q1q2q3cc¯+32q1q2q3cc¯+][32q1q2q3cc¯+52q1q2q3cc¯+]32q1q2q3cc¯+[12q1q2q3cc¯+32q1q2q3cc¯+52q1q2q3cc¯+][12q1q2q3cc¯+32q1q2q3cc¯+52q1q2q3cc¯+72q1q2q3cc¯+],

where the 1 denotes the additional P-wave and is embodied by a γ5 in constructing the currents, the subscripts q1q2, q3c, denote the quark constituents. The quark and antiquark have opposite parity, we usually take it for granted that the quarks (antiquarks) have positive (negative) parity, and the c¯-quark has the JP=12.

We write down the currents Jq1q2q3jLjH(x) explicitly,

Juuu11(x)=εilaεijkεlmnujT(x)Cγμuk(x)umT(x)Cγμcn(x)Cc¯aT(x),Juud11(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)dmT(x)Cγμcn(x)+2ujT(x)Cγμdk(x)umT(x)Cγμcn(x)]Cc¯aT(x),Judd11(x)=εilaεijkεlmn3[djT(x)Cγμdk(x)umT(x)Cγμcn(x)+2djT(x)Cγμuk(x)dmT(x)Cγμcn(x)]Cc¯aT(x),Jddd11(x)=εilaεijkεlmndjT(x)Cγμdk(x)dmT(x)Cγμcn(x)Cc¯aT(x),

Juus11(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)smT(x)Cγμcn(x)+2ujT(x)Cγμsk(x)umT(x)Cγμcn(x)]Cc¯aT(x),Juds11(x)=εilaεijkεlmn3[ujT(x)Cγμdk(x)smT(x)Cγμcn(x)+ujT(x)Cγμsk(x)dmT(x)Cγμcn(x)+djT(x)Cγμsk(x)umT(x)Cγμcn(x)]Cc¯aT(x),Jdds11(x)=εilaεijkεlmn3[djT(x)Cγμdk(x)smT(x)Cγμcn(x)+2djT(x)Cγμsk(x)dmT(x)Cγμcn(x)]Cc¯aT(x),

Juss11(x)=εilaεijkεlmn3[sjT(x)Cγμsk(x)umT(x)Cγμcn(x)+2sjT(x)Cγμuk(x)smT(x)Cγμcn(x)]Cc¯aT(x),Jdss11(x)=εilaεijkεlmn3[sjT(x)Cγμsk(x)dmT(x)Cγμcn(x)+2sjT(x)Cγμdk(x)smT(x)Cγμcn(x)]Cc¯aT(x),

Jsss11(x)=εilaεijkεlmnsjT(x)Cγμsk(x)smT(x)Cγμcn(x)Cc¯aT(x),

Juuu10(x)=εilaεijkεlmnujT(x)Cγμuk(x)umT(x)Cγ5cn(x)γ5γμCc¯aT(x),Juud10(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)dmT(x)Cγ5cn(x)+2ujT(x)Cγμdk(x)umT(x)Cγ5cn(x)]γ5γμCc¯aT(x),Judd10(x)=εilaεijkεlmn3[djT(x)Cγμdk(x)umT(x)Cγ5cn(x)+2djT(x)Cγμuk(x)dmT(x)Cγ5cn(x)]γ5γμCc¯aT(x),Jddd10(x)=εilaεijkεlmndjT(x)Cγμdk(x)dmT(x)Cγ5cn(x)γ5γμCc¯aT(x),

Juus10(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)smT(x)Cγ5cn(x)+2ujT(x)Cγμsk(x)umT(x)Cγ5cn(x)]γ5γμCc¯aT(x),Juds10(x)=εilaεijkεlmn3[ujT(x)Cγμdk(x)smT(x)Cγ5cn(x)+ujT(x)Cγμsk(x)dmT(x)Cγ5cn(x)+djT(x)Cγμsk(x)umT(x)Cγ5cn(x)]γ5γμCc¯aT(x),Jdds10(x)=εilaεijkεlmn3[djT(x)Cγμdk(x)smT(x)Cγ5cn(x)+2djT(x)Cγμsk(x)dmT(x)Cγ5cn(x)]γ5γμCc¯aT(x),

Juss10(x)=εilaεijkεlmn3[sjT(x)Cγμsk(x)umT(x)Cγ5cn(x)+2sjT(x)Cγμuk(x)smT(x)Cγ5cn(x)]γ5γμCc¯aT(x),Jdss10(x)=εilaεijkεlmn3[sjT(x)Cγμsk(x)dmT(x)Cγ5cn(x)+2sjT(x)Cγμdk(x)smT(x)Cγ5cn(x)]γ5γμCc¯aT(x),

Jsss10(x)=εilaεijkεlmnsjT(x)Cγμsk(x)smT(x)Cγ5cn(x)γ5γμCc¯aT(x),

where the superscripts jL and jH are the spins of the light and heavy diquarks, respectively.

We take the isospin limit and classify the currents which couple potentially to the pentaquark states with degenerate masses into the following 8 types,

Juuu11(x),Juud11(x),Judd11(x),Jddd11(x);Juus11(x),Juds11(x),Jdds11(x);Juss11(x),Jdss11(x);Jsss11(x);Juuu10(x),Juud10(x),Judd10(x),Jddd10(x);Juus10(x),Juds10(x),Jdds10(x);Juss10(x),Jdss10(x);Jsss10(x),

and perform the operator product expansion up to the vacuum condensates of dimension 10 to obtain the QCD sum rules, the predictions for the hidden-charm pentaquark states Pq1q2q3jLjH12 with the spin-parity JP=12± are presented in Table 61 in the Appendix.

From Table 61, we can see that the two-body strong decays to the J/ψB10,

Pq1q2q31112(12),Pq1q2q31012(12)J/ψB10,

for example,

Puuu1112(12),Puuu1012(12)J/ψΔ++,Puus1112(12),Puus1012(12)J/ψΣ+,Puss1112(12),Puss1012(12)J/ψΞ0,Psss1112(12),Psss1012(12)J/ψΩ,

could take place, but the decay widths are rather small due to the small available phase-spaces; on the other hand, the two-body strong decays,

Pq1q2q31112(12),Pq1q2q31012(12)J/ψB8,

Pq1q2q31112(12+)J/ψB10,J/ψB8,

for example,

Puud1112(12±),Puud1012(12)J/ψp,

Puus1112(12±),Puus1012(12)J/ψΣ+,Puss1112(12±),Puss1012(12)J/ψΞ0,Puuu1112(12+)J/ψΔ++,Puus1112(12+)J/ψΣ+,Puss1112(12+)J/ψΞ0,Psss1112(12+)J/ψΩ,

can take place more easily, the decay widths are larger due to the larger available phase-spaces; furthermore, the two-body strong decays

Pq1q2q31012(12+)J/ψB8,J/ψB10,

for example,

Puud1012(12+)J/ψp,Puus1012(12+)J/ψΣ+,

Puss1012(12+)J/ψΞ0,Puuu1012(12+)J/ψΔ++,Puus1012(12+)J/ψΣ+,Puss1012(12+)J/ψΞ0,Psss1012(12+)J/ψΩ,

can take place fluently, the decay widths are rather large due to the large available phase-spaces. We can search for those pentaquark states in the J/ψB8 and J/ψB10 mass spectra in the decays of the bottom baryons to the final states J/ψB8 and J/ψB10 associated with the light vector mesons or pseudoscalar mesons [693, 707, 708], for example,

ΩbPuss1112(12±)KJ/ψΞ0K,ΩbPsss1112(12±)ϕJ/ψΩϕ.

In Ref. [705], we construct the currents Jq1q2q3,μjLjHj(x) to interpolate the JP=32± hidden-charm tetraquark states, where the superscripts jL and jH are the spins of the light and heavy diquarks, respectively, j=jH+jc¯, the jc¯ is the spin of the heavy antiquark, the subscripts q1, q2, q3 are the light quark constituents u, d or s. We write down the currents Jq1q2q3,μjLjHj(x) explicitly,

Juuu,μ1012(x)=εilaεijkεlmnujT(x)Cγμuk(x)umT(x)Cγ5cn(x)Cc¯aT(x),Juud,μ1012(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)dmT(x)Cγ5cn(x)+2ujT(x)Cγμdk(x)umT(x)Cγ5cn(x)]Cc¯aT(x),Judd,μ1012(x)=εilaεijkεlmn3[djT(x)Cγμdk(x)umT(x)Cγ5cn(x)+2djT(x)Cγμuk(x)dmT(x)Cγ5cn(x)]Cc¯aT(x),Jddd,μ1012(x)=εilaεijkεlmndjT(x)Cγμdk(x)dmT(x)Cγ5cn(x)Cc¯aT(x),

Juus,μ1012(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)smT(x)Cγ5cn(x)+2ujT(x)Cγμsk(x)umT(x)Cγ5cn(x)]Cc¯aT(x),Juds,μ1012(x)=εilaεijkεlmn3[ujT(x)Cγμdk(x)smT(x)Cγ5cn(x)+ujT(x)Cγμsk(x)dmT(x)Cγ5cn(x)+djT(x)Cγμsk(x)umT(x)Cγ5cn(x)]Cc¯aT(x),Jdds,μ1012(x)=εilaεijkεlmn3[djT(x)Cγμdk(x)smT(x)Cγ5cn(x)+2djT(x)Cγμsk(x)dmT(x)Cγ5cn(x)]Cc¯aT(x),

Juss,μ1012(x)=εilaεijkεlmn3[sjT(x)Cγμsk(x)umT(x)Cγ5cn(x)+2sjT(x)Cγμuk(x)smT(x)Cγ5cn(x)]Cc¯aT(x),Jdss,μ1012(x)=εilaεijkεlmn3[sjT(x)Cγμsk(x)dmT(x)Cγ5cn(x)+2sjT(x)Cγμdk(x)smT(x)Cγ5cn(x)]Cc¯aT(x),

Jsss,μ1012(x)=εilaεijkεlmnsjT(x)Cγμsk(x)smT(x)Cγ5cn(x)Cc¯aT(x),

Juuu,μ1112(x)=εilaεijkεlmnujT(x)Cγμuk(x)umT(x)Cγαcn(x)γ5γαCc¯aT(x),Juud,μ1112(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)dmT(x)Cγαcn(x)+2ujT(x)Cγμdk(x)umT(x)Cγαcn(x)]γ5γαCc¯aT(x),Judd,μ1112(x)=εilaεijkεlmn3[djT(x)Cγμdk(x)umT(x)Cγαcn(x)+2djT(x)Cγμuk(x)dmT(x)Cγαcn(x)]γ5γαCc¯aT(x),Jddd,μ1112(x)=εilaεijkεlmndjT(x)Cγμdk(x)dmT(x)Cγαcn(x)γ5γαCc¯aT(x),

Juus,μ1112(x)=εilaεijkεlmn3[ujT(x)Cγμuk(x)smT(x)Cγαcn(x)+2ujT(x)Cγμsk(x)umT(x)Cγαcn(x)]γ5γαCc¯aT(x),Juds,μ1112(x)=εilaεijkεlmn3[ujT(x)Cγμdk(x)smT(x)Cγαcn(x)+ujT(x)Cγμsk(x)dmT(x)Cγαcn(x)+djT(x)Cγμsk(x)umT(x)Cγαcn(x)]γ5γαCc¯aT(x),Jdds,μ1112(x)=εilaεijkεlmn3[djT(x)Cγμdk(x)smT(x)Cγαcn(x)+2djT(x)Cγμsk(x)dmT(x)Cγαcn(x)]γ5γαCc¯aT(x),

Juss,μ1112(x)=εilaεijkεlmn3[sjT(x)Cγμsk(x)umT(x)Cγαcn(x)+2sjT(x)Cγμuk(x)smT(x)Cγαcn(x)]γ5γαCc¯aT(x),Jdss,μ1112(x)=εilaεijkεlmn3[sjT(x)Cγμsk(x)dmT(x)Cγαcn(x)+2sjT(x)Cγμdk(x)smT(x)Cγαcn(x)]γ5γαCc¯aT(x),

Jsss,μ1112(x)=εilaεijkεlmnsjT(x)Cγμsk(x)smT(x)Cγαcn(x)γ5γαCc¯aT(x),

Juuu,μ1132(x)=εilaεijkεlmnujT(x)Cγαuk(x)umT(x)Cγμcn(x)γ5γαCc¯aT(x),Juud,μ1132(x)=εilaεijkεlmn3[ujT(x)Cγαuk(x)dmT(x)Cγμcn(x)+2ujT(x)Cγαdk(x)umT(x)Cγμcn(x)]γ5γαCc¯aT(x),Judd,μ1132(x)=εilaεijkεlmn3[djT(x)Cγαdk(x)umT(x)Cγμcn(x)+2djT(x)Cγαuk(x)dmT(x)Cγμcn(x)]γ5γαCc¯aT(x),Jddd,μ1132(x)=εilaεijkεlmndjT(x)Cγαdk(x)dmT(x)Cγμcn(x)γ5γαCc¯aT(x),

Juus,μ1132(x)=εilaεijkεlmn3[ujT(x)Cγαuk(x)smT(x)Cγμcn(x)+2ujT(x)Cγαsk(x)umT(x)Cγμcn(x)]γ5γαCc¯aT(x),Juds,μ1132(x)=εilaεijkεlmn3[ujT(x)Cγαdk(x)smT(x)Cγμcn(x)+ujT(x)Cγαsk(x)dmT(x)Cγμcn(x)+djT(x)Cγαsk(x)umT(x)Cγμcn(x)]γ5γαCc¯aT(x),Jdds,μ1132(x)=εilaεijkεlmn3[djT(x)Cγαdk(x)smT(x)Cγμcn(x)+2djT(x)Cγαsk(x)dmT(x)Cγμcn(x)]γ5γαCc¯aT(x),

Juss,μ1132(x)=εilaεijkεlmn3[sjT(x)Cγαsk(x)umT(x)Cγμcn(x)+2sjT(x)Cγαuk(x)smT(x)Cγμcn(x)]γ5γαCc¯aT(x),Jdss,μ1132(x)=εilaεijkεlmn3[sjT(x)Cγαsk(x)dmT(x)Cγμcn(x)+2sjT(x)Cγαdk(x)smT(x)Cγμcn(x)]γ5γαCc¯aT(x),

Jsss,μ1132(x)=εilaεijkεlmnsjT(x)Cγαsk(x)smT(x)Cγμcn(x)γ5γαCc¯aT(x).

We take the isospin limit and classify the thirty currents couple potentially to the hidden-charm pentaquark states with degenerate masses into the following twelve types,

Juuu,μ1012(x),Juud,μ1012(x),Judd,μ1012(x),Jddd,μ1012(x);Juus,μ1012(x),Juds,μ1012(x),Jdds,μ1012(x);Juss,μ1012(x),Jdss,μ1012(x);

Jsss,μ1012(x);Juuu,μ1112(x),Juud,μ1112(x),Judd,μ1112(x),Jddd,μ1112(x);Juus,μ1112(x),Juds,μ1112(x),Jdds,μ1112(x);Juss,μ1112(x),Jdss,μ1112(x);Jsss,μ1112(x);

Juuu,μ1132(x),Juud,μ1132(x),Judd,μ1132(x),Jddd,μ1132(x);Juus,μ1132(x),Juds,μ1132(x),Jdds,μ1132(x);Juss,μ1132(x),Jdss,μ1132(x);Jsss,μ1132(x),

and perform the operator product expansion up to the vacuum condensates of dimension 10 to obtain the QCD sum rules, the predictions are presented in Table 62 in the Appendix.

According to Fig.26 and Fig.27, the vacuum condensates of the dimensions 11 and 13 play an important role, we should update the old calculations, just like in Ref. [402].

In Ref. [514], we extend our previous works [402, 703] to explore the possible assignment of the Pcs(4459) as the [ud][sc]c¯ (0, 0, 0, 12) pentaquark state with the spin-parity JP=12. We choose the current J(x) and carry out the operator product expansion up to the vacuum condensates of dimension 13, where

J(x)=εilaεijkεlmnujT(x)Cγ5dk(x)smT(x)Cγ5cn(x)Cc¯aT(x).

In calculations, we take the modified energy scale formula μ=MP2(2Mc)2ms(μ)=2.4GeV to constraint the QCD spectral densities.

We obtain the Borel window T2=3.43.8GeV2 via trial and error, the pole contribution is about (4061)%, which is large enough to extract the pentaquark mass reliably. In Fig.28, we plot the contributions of the higher dimensional vacuum condensates with variation of the Borel parameter T2. The higher dimensional vacuum condensates manifest themselves at the region T2<2.5GeV2, we should choose the value T2>2.5GeV2. Their values decrease monotonously and quickly with the increase of the Borel parameter, in the Borel window T2=3.43.8GeV2, the contributions of the higher dimensional vacuum condensates are D(8)=(2126)%, D(9)=(68)%, D(10)=(23)%, D(11)=(12)%, D(13)1%, the convergent behavior is very good [514].

In Fig.29, we plot the predicted pentaquark mass with variation of the Borel parameter T2 with the truncations of the operator product expansion up to the vacuum condensates of dimensions n=10 and 13, respectively. From the figure, we can see explicitly that the vacuum condensates of dimensions 11 and 13 play an important role to obtain the flat platform, we should take them into account in a consistent way, just like in the case of Ref. [402], where no s-quark is present. All in all, the higher dimensional vacuum condensates play an important role in obtaining the flat platform, especially those associated with the inverse Borel parameters 1T2, 1T4, 1T6 and 1T8.

At last, we obtain the mass and pole residue [514],

MP=4.47±0.11GeV,λP=(1.86±0.31)×103GeV6.

The predicted mass MP=4.47±0.11GeV is in excellent agreement with the experimental data 4458.8±2.91.1+4.7MeV from the LHCb Collaboration [198], and supports assigning the Pcs(4459) as the [ud][sc]c¯ (0, 0, 0, 12) pentaquark state with the spin-parity JP=12. In Ref. [402], we observe that the Pc(4312) can be assigned to be the [ud][uc]c¯ (0, 0, 0, 12) pentaquark state with the spin-parity JP=12. The light-flavor SU(3) mass-breaking effect is about Δ=ms=147MeV, the estimations presented in Tab.52 are reasonable and reliable, where we take the experimental value of the mass of the Pcs(4459). In Refs. [704, 705], we study the JP=12± and 32± hidden-charm pentaquark states with the strangeness S=0, 1, 2 and 3 in a systematic way by carrying out the operator product expansion up to the vacuum condensates of dimension 10 and choosing the old value Mc=1.80GeV, and observe that the light-flavor SU(3) mass-breaking effects are Δ=(90130)MeV for the negative-parity pentaquark states. The new analysis supports a larger light-flavor SU(3) mass-breaking effect.

We can extend this section directly to study the 3¯3¯3¯ type triply-heavy pentaquark states [709].

5.2 11 type pentaquark states

In Ref. [710], Chen et al. constructed the color singlet-singlet type currents,

JμD¯Σc(x)=c¯(x)γμd(x)εijkuiT(x)Cγνuj(x)γνγ5ck(x),JμD¯Σc(x)=c¯(x)γ5d(x)εijkuiT(x)Cγμuj(x)ck(x),

JμνD¯Σc(x)=c¯(x)γμd(x)εijkuiT(x)Cγνuj(x)γ5ck(x)+(μν),JμνD¯Σc(x)=c¯(x)γμγ5d(x)εijkuiT(x)Cγνuj(x)ck(x)+(μν),JμνD¯Λc(x)=c¯(x)γμu(x)εijkuiT(x)Cγνγ5dj(x)c(x)+(μν),

for the first time, and take the currents JμD¯Σc(x) and JμνD¯Σc(x) to study the Pc(4380) and Pc(4450) by carrying out the operator product expansion up to the vacuum condensates of dimension 8. Thereafter, the 11 type currents were used to interpolate the pentaquark molecular states [418, 487, 711-719]. However, in those works, the isospins of the currents are not specified and should be improved, as the two-body strong decays to the final states J/ψp and J/ψΛ conserve isospins.

The u and d quarks have the isospin I=12, i.e., I^u=12u and I^d=12d, where the I^ is the isospin operator. Then the D¯0, D¯0, D¯, D¯, D¯s, D¯s, Σc+, Σc+, Σc++, Σc++, Ξc0, Ξc0, Ξc+, Ξc+, Ξc0, Ξc+ and Λc+ correspond to the eigenstates |12,12, |12,12, |12,12, |12,12, |0,0, |0,0, |1,0, |1,0, |1,1, |1,1, |12,12, |12,12, |12,12, |12,12, |12,12, |12,12 and |0,0 in the isospin space |I,I3, respectively. We construct the color-singlet currents to interpolate them,

JD¯0/(x)=c¯(x)iγ5u/d(x),JD¯s(x)=c¯(x)iγ5s(x),JμD¯0/(x)=c¯(x)γμu/d(x),JμD¯s(x)=c¯(x)γμs(x),

JΣc+(x)=εijkuiT(x)Cγμdj(x)γμγ5ck(x),JΣc++(x)=εijkuiT(x)Cγμuj(x)γμγ5ck(x),JμΣc+(x)=εijkuiT(x)Cγμdj(x)ck(x),JμΣc++(x)=εijkuiT(x)Cγμuj(x)ck(x),

JΞc0(x)=εijkdiT(x)Cγμsj(x)γμγ5ck(x),JμΞc0(x)=εijkdiT(x)Cγμsj(x)ck(x),JΞc+(x)=εijkuiT(x)Cγμsj(x)γμγ5ck(x),JμΞc+(x)=εijkuiT(x)Cγμsj(x)ck(x),

JΞc0(x)=εijkdiT(x)Cγ5sj(x)ck(x),JΞc+(x)=εijkuiT(x)Cγ5sj(x)ck(x),JΛc+(x)=εijkuiT(x)Cγ5dj(x)ck(x).

Accordingly, we construct the 11 type five-quark currents to interpolate the D¯()Σc(), D¯()Ξc(), D¯()Ξc and D¯()Λc type pentaquark sates, where the D¯(), Σc(), Ξc(), Ξc and Λc represent the color-singlet clusters having the same quantum numbers as the physical states D¯(), Σc(), Ξc(), Ξc and Λc respectively except for the masses,

J(x)=J12D¯Σc(x),J32D¯Σc(x),J0D¯Ξc(x),J1D¯Ξc(x),J0D¯Ξc(x),J1D¯Ξc(x),J12D¯Λc(x),J12D¯sΞc(x),J0D¯sΛc(x),Jμ(x)=J12;μD¯Σc(x),J32;μD¯Σc(x),J12;μD¯Σc(x),J32;μD¯Σc(x),J0;μD¯Ξc(x),J1;μD¯Ξc(x),J0;μD¯Ξc(x),J1;μD¯Ξc(x),J0;μD¯Ξc(x),J1;μD¯Ξc(x),J12;μD¯Λc(x),J12;μD¯sΞc(x),J0;μD¯sΛc(x),Jμν(x)=J12;μνD¯Σc(x),J32;μνD¯Σc(x),J0;μνD¯Ξc(x),J1;μνD¯Ξc(x),

J12D¯Σc(x)=13JD¯0(x)JΣc+(x)23JD¯(x)JΣc++(x),J32D¯Σc(x)=23JD¯0(x)JΣc+(x)+13JD¯(x)JΣc++(x),J12;μD¯Σc(x)=13JD¯0(x)JμΣc+(x)23JD¯(x)JμΣc++(x),J32;μD¯Σc(x)=23JD¯0(x)JμΣc+(x)+13JD¯(x)JμΣc++(x),

J12;μD¯Σc(x)=13JμD¯0(x)JΣc+(x)23JμD¯(x)JΣc++(x),J32;μD¯Σc(x)=23JμD¯0(x)JΣc+(x)+13JμD¯(x)JΣc++(x),J12;μνD¯Σc(x)=13JμD¯0(x)JνΣc+(x)23JμD¯(x)JνΣc++(x)+(μν),J32;μνD¯Σc(x)=23JμD¯0(x)JνΣc+(x)+13JμD¯(x)JνΣc++(x)+(μν),

J0D¯Ξc(x)=12JD¯0(x)JΞc0(x)12JD¯(x)JΞc+(x),J1D¯Ξc(x)=12JD¯0(x)JΞc0(x)+12JD¯(x)JΞc+(x),J0;μD¯Ξc(x)=12JD¯0(x)JμΞc0(x)12JD¯(x)JμΞc+(x),J1;μD¯Ξc(x)=12JD¯0(x)JμΞc0(x)+12JD¯(x)JμΞc+(x),

J0;μD¯Ξc(x)=12JμD¯0(x)JΞc0(x)12JμD¯(x)JΞc+(x),J1;μD¯Ξc(x)=12JμD¯0(x)JΞc0(x)+12JμD¯(x)JΞc+(x),J0;μνD¯Ξc(x)=12JμD¯0(x)JνΞc0(x)12JμD¯(x)JνΞc+(x)+(μν),J1;μνD¯Ξc(x)=12JμD¯0(x)JνΞc0(x)+12JμD¯(x)JνΞc+(x)+(μν),

J0D¯Ξc(x)=12JD¯0(x)JΞc0(x)12JD¯(x)JΞc+(x),J1D¯Ξc(x)=12JD¯0(x)JΞc0(x)+12JD¯(x)JΞc+(x),J12D¯Λc(x)=JD¯0(x)JΛc+(x),J12D¯sΞc(x)=JD¯s(x)JΞc+(x),J0D¯sΛc(x)=JD¯s(x)JΛc+(x),

J0;μD¯Ξc(x)=12JμD¯0(x)JΞc0(x)12JμD¯(x)JΞc+(x),J1;μD¯Ξc(x)=12JμD¯0(x)JΞc0(x)+12JμD¯(x)JΞc+(x),J12;μD¯Λc(x)=JμD¯0(x)JΛc+(x),J12;μD¯sΞc(x)=JμD¯s(x)JΞc+(x),J0;μD¯sΛc(x)=JμD¯s(x)JΛc+(x),

the subscripts 12, 32, 0 and 1 represent the isospins I [418, 720-722]. They are the isospin eigenstates |I,I3= |12,12, |32,12, |1,0 or |0,0.

For example, from Eq. (327), we obtain the relations,

JD¯0(x)JΣc+(x)=13J12D¯Σc(x)+23J32D¯Σc(x),JD¯(x)JΣc++(x)=13J32D¯Σc(x)23J12D¯Σc(x),

the currents JD¯0(x)JΣc+(x) and JD¯(x)JΣc++(x) have both the isospin (I,I3)=(12,12) and (32,12) components, and couple potentially to the pentaquark molecular states with the isospins (12,12) and (32,12), which decay to the final states J/ψp and J/ψΔ+, respectively. As the Pc(4312), Pc(4380), Pc(4440), Pc(4457) and Pc(4337) are observed in the J/ψp mass spectrum, it is better to choose the current J12D¯Σc(x) with the definite isospin, we prefer the color singlet−singlet type currents with the definite isospins [418, 720-722], and thereafter those currents are adopted in Refs. [723-725]. Although the mass splitting between the isospin cousins is of several 10-MeV in most cases, in some cases, the mass splitting can be as large as 150 MeV, see Tab.54. Phenomenologically, the molecule-type P states have been studied extensively with the help of heavy quark symmetry [214, 257, 268, 269, 277, 287-289, 726-738], it is more easy to apply isospin eigenstates in the effective field theory than in QCD.

We resort to the correlation functions Π(p), Πμν(p) and Πμναβ(p) in Eq. (269), and perform analogous analysis to obtain the QCD sum rules for the masses and pole residues like Eqs. (284)−(285),

2Mλj2exp(M2T2)=4mc2s0dsρQCD,j(s)exp(sT2),

M2=4mc2s0dsdd(1/T2)ρQCD,j(s)exp(sT2)4mc2s0dsρQCD,j(s)exp(sT2),

where j=12, 32 and 52, the pole residues are defined analogous to Eqs. (272)−(274).

As we study the 11 type pentaquark states, it is better to choose the updated values Mc=1.85±0.01GeV and Ms=0.2GeV to determine the optimal energy scales of the QCD spectral densities with the formula [641],

μ=MX/Y/Z/P2(2Mc)2kMs.

After trial and error, we obtain the Borel windows, continuum threshold parameters, energy scales and pole contributions, see Tab.53 [720-722], the pole contributions are about or slightly larger than (40−60)%, we obtain the largest pole contributions up to now. In the Borel windows, the highest dimensional condensate contributions |D(12)| and |D(13)| are approximately zero, the most important contributions are mainly from the lowest order contributions q¯q, q¯q2 and q¯gsσGqq¯q. The operator product expansion converges very well.

Then we calculate the uncertainties of the masses and pole residues, which are shown in Tab.54 [720-722]. The central value of the mass of the D¯Σc molecular state with the IJP=1212 is 4.31 GeV, it is only about 10 MeV below the D¯0Σc+ threshold, we assign it as the Pc(4312) tentatively. For the D¯Σc molecular state with the IJP=3212, the central value of the mass is 4.33 GeV, which is about 10 MeV above the D¯Σc++ threshold, we tentatively assign it as the D¯Σc resonance, the isospin cousin of the Pc(4312).

In a similar way, we assign the Pc(4380), Pc(4440) and Pc(4457) as the D¯Σc, D¯Σc and D¯Σc molecular states with the IJP=1232, 1232 and 1252, respectively. For the molecular states (resonances) D¯Σc with the IJP=3232, D¯Σc with the IJP=3232 and D¯Σc with the IJP=3252, the central values of the masses are about 20 MeV, 10 MeV and 90 MeV above the corresponding meson-baryon thresholds, respectively.

If we choose the same input parameters, the 11 type pentaquark states with the isospin I=32 (1) have slightly larger masses than the corresponding ones with the isospin I=12 (0).

The Pc(4312), Pc(4380), Pc(4440) and Pc(4457) can be assigned as the D¯Σc, D¯Σc, D¯Σc and D¯Σc molecular states with the isospin I=12 respectively, since the two-body strong decays PcJ/ψp conserve isospin. If the assignments are robust, there exist four slightly higher molecular states D¯Σc, D¯Σc, D¯Σc and D¯Σc with the isospin I=32, we can search for the four resonances in the J/ψΔ+ mass spectrum, as the two-body strong decays PcJ/ψΔ+ also conserve isospin, the J/ψ, p and Δ have the isospins I=0, 12 and 32, respectively.

The D¯Ξc and D¯Ξc molecular states lie about 0.1GeV and 0.2GeV above the Pcs(4459) respectively, the Pcs(4459) is unlikely to be the D¯Ξc or D¯Ξc molecular state. The mass of the D¯Ξ molecular state with the isospin I=1 is 4.450.08+0.07GeV, which is near the value 4459MeV, but lies slightly above the corresponding meson-baryon threshold, it is a resonance, from the decay channel Pcs(4459)J/ψΛ [198], the isospin of the Pcs(4459) is zero, which excludes assigning the Pcs(4459) as the D¯Ξc molecular state with the isospin I=1. The mass of the D¯Ξc molecular state with the isospin I=0 is 4.460.07+0.07GeV, which is in very good agreement with the Pcs(4459), it is very good to assign the Pcs(4459) as the D¯Ξc molecular state with the isospin I=0 and the spin-parity JP=32. The predications also favor assigning the Pcs(4459) as the D¯Ξc molecular state with the spin-parity JP=32 and isospin I=0.

The predictions support assigning the Pcs(4338) as the D¯Ξc molecular state with the spin-parity JP=12 and isospin I=0, the observation of its cousin with the isospin I=1 in the J/ψΣ0/ηcΣ0 mass spectrum would decipher the inner structure of the Pcs(4338). However, there exists no candidate for the Pc(4337) [199].

Beyond the 3¯3¯3¯ type currents, in Ref. [739], Pimikov, Lee and Zhang construct the color 88 type currents JA(Γ2,Γ3) and JμS(Γ2,Γ3) to interpolate the hidden-charm pentaquark states,

JA(Γ2,Γ3)=εf1f2f3εc1c3ctcc2m[q1TCq2Γ2q3q1TCγ5q2γ5Γ2q3][q¯5tmΓ3q4],JμS(Γ2,Γ3)=εf1f2f3εc1c3ctcc2mq1TCγμq2Γ2q3[q¯5tmΓ3q4],

where the quark fields have the flavors fi and color ci, the Γ2 and Γ3 are some Dirac matrixes. Then, they calculate the mass spectrum by taking account of the vacuum condensates q¯q, αsGGπ, q¯gsσGq, q¯q2, q¯qq¯gsσGq, q¯q3 and q¯gsσGq2.

6 Singly-heavy exotic states

6.1 Singly-heavy tetraquark states

The X0(2900) and X1(2900) observed in the DK+ mass spectrum are the first exotic structures with fully open flavor [101, 102], they have the valence quarks udc¯s¯. The Tcs¯0(2900) and Tcs¯++(2900) are observed in the Ds+π and Ds+π+ mass spectra, respectively [130, 131], they have the valence quarks cs¯ud¯ and cs¯u¯d, respectively.

Based on the predicted masses of the AA-type scalar tetraquark states from the QCD sum rules [61, 740],

Mqqq¯q¯=1.86±0.11GeV,Msss¯s¯=2.08±0.13GeV,Mcqc¯q¯=3.95±0.09GeV,

we estimate the mass of the AA-type csu¯d¯ tetraquark state crudely,

Mcsu¯d¯=Mqqq¯q¯+Msss¯s¯+2Mcqc¯q¯4=2.96±0.11GeV,

which is consistent with the mass of the X0(2900) within uncertainties [119].

In Ref. [119], we construct the AA and SS-type scalar four-quark currents,

JAA(x)=εijkεimnsjT(x)Cγαck(x)u¯m(x)γαCd¯nT(x),JSS(x)=εijkεimnsjT(x)Cγ5ck(x)u¯m(x)γ5Cd¯nT(x),

to study the csu¯d¯ tetraquark states with the correlation function Π(p), see Eq. (125). We carry out the operator product expansion up to the vacuum condensates of dimension-11 and assume vacuum saturation for the higher dimensional vacuum condensates according to the routine in Section 2.2 and Section 3.1.1. As there are three q-quark lines and one Q-quark line, if each Q-quark line emits a gluon and each q-quark line contributes a quark-antiquark pair, we obtain a quark-gluon operator gsGμνq¯qq¯qs¯s, which is of dimension 11, and leads to the vacuum condensates q¯q2s¯gsσGs and q¯qs¯sq¯gsσGq.

We obtain the QCD sum rules routinely, at the QCD side, we choose the flavor numbers nf=4 and the typical energy scale μ=1GeV. After trial and error, we obtain the Borel windows and continuum threshold parameters, therefore the pole contributions of the ground states and convergent behaviors of the operator product expansion, see Tab.55. In the Borel windows, the pole contributions are about (3867)%, the central values exceed 52%. The contributions of the vacuum condensates |D(11)| are about (24)% and (01)% for the AA and SS-type tetraquark states, respectively.

At last, we obtain the values of the masses and pole residues, which are also shown in Tab.55. In Fig.30, we plot the masses of the AA and SS-type scalar csu¯d¯ tetraquark states with variations of the Borel parameters T2 in much larger ranges than the Borel windows, there appear platforms in the Borel windows indeed.

The predicted mass MAA=2.91±0.12GeV is consistent with the experimental value 2866±7MeV from the LHCb Collaboration [101, 102], and supports assigning the X0(2900) to be the AA-type csu¯d¯ tetraquark state with the spin-parity JP=0+. While the predicted mass MSS=3.05±0.10GeV lies above the experimental value 2866±7MeV [101, 102].

In Ref. [741], we construct the AA¯-type currents to study the ground state mass spectrum of the tetraquark states with strange and doubly-strange via the QCD sum rules to verify the inner structures of the Tcs¯(2900), where

J(x)=Js0(x),Jss0(x),Jμν(x)=Js,μν1(x),Jss,μν1(x),Js,μν2(x),Jss,μν2(x),

Js0(x)=εijkεimnujT(x)Cγμck(x)d¯m(x)γμCs¯nT(x),Jss0(x)=εijkεimnujT(x)Cγμck(x)s¯m(x)γμCs¯nT(x),Js,μν1(x)=εijkεimn[ujT(x)Cγμck(x)d¯m(x)γνCs¯nT(x)ujT(x)Cγνck(x)d¯m(x)γμCs¯nT(x)],Jss,μν1(x)=εijkεimn[ujT(x)Cγμck(x)s¯m(x)γνCs¯nT(x)ujT(x)Cγνck(x)s¯m(x)γμCs¯nT(x)],Js,μν2(x)=εijkεimn[ujT(x)Cγμck(x)d¯m(x)γνCs¯nT(x)+ujT(x)Cγνck(x)d¯m(x)γμCs¯nT(x)],Jss,μν2(x)=εijkεimn[ujT(x)Cγμck(x)s¯m(x)γνCs¯nT(x)+ujT(x)Cγνck(x)s¯m(x)γμCs¯nT(x)],

the superscripts 0, 1 and 2 denote the spins. With a simple replacement ud, we obtain the corresponding currents in the same isospin multiplets. In the isospin limit, the tetraquark states in the same multiplets have the same masses.

Again, we resort to the correlation functions Π(p) and Πμναβ(p), see Eq. (125), and obtain the ground state contributions according to the hadron representations in Eqs. (202)−(204), and obtain the QCD sum rules routinely, again we take the flavor numbers nf=4 and the typical energy scale μ=1.0GeV.

After trial and error, we obtain the Borel windows, continuum threshold parameters and pole contributions, see Tab.56, where the pole contributions are 38%66% and the central values for the six states are larger than 50%, furthermore, the contributions of the vacuum condensates show a descending trend |D(6)|>|D(8)|>|D(9)|>|D(10)||D(11)|0. At last, we obtain the masses and pole residues, which are also shown in Tab.56 [741].

In Tab.56, the predicted mass of the JP=0+ state cud¯s¯, M=2.92±0.12GeV, is in very good agreement with the experimental values M=2.892±0.014±0.015GeV and 2.921±0.017±0.020GeV from the LHCb Collaboration [130, 131], and supports assigning the Tcs¯(2900) to be the AA¯-type cs¯qq¯ tetraquark states with the spin-parity JP=0+.

Those typical singly-charmed tetraquark candidates, which lie at 2.9GeV, have attracted many theoretical works, and are assigned as the 3¯3 type tetraquark states [118-121, 742], or their radial/orbital excitations [122], or non-tetraquark states [123], DK¯ molecular states [121, 124-128, 743, 744, 745], triangle singularities [129, 746, 747], etc. We could only obtain a mass about 2.3GeV for the singly-charmed tetraquark states at the cost of sacrificing the pole dominance [748, 749, 750].

At the bottom sector, the singly-bottom tetraquark candidate X(5568) is not confirmed [751, 752]. The lowest mass of the ground state bs¯u¯d might have a mass MT/X+mb(mb)mc(mc)5.81GeV, which lies above the X(5568).

6.2 Singly-heavy pentaquark states

The experimental candidates for the singly-charmed pentaquark states are not as robust as that of the singly-charmed tetraquark states, the assignments in the scenario of pentaquark states are only conjectures.

In 2017, the LHCb Collaboration observed five narrow structures Ωc(3000), Ωc(3050), Ωc(3066), Ωc(3090) and Ωc(3119) [753]. Also in 2017, the Belle Collaboration confirmed the Ωc(3000), Ωc(3050), Ωc(3066) and Ωc(3090) in the Ξc+K decay mode [754]. In 2023, the LHCb Collaboration observed the Ωc(3185) and Ωc(3327) in the Ξc+K mass spectrum [755], which lie near the DΞ and DΞ thresholds, respectively, the measured Breit−Wigner masses and decay widths are

Ωc(3185):M=3185.1±1.70.9+7.4±0.2MeV,Γ=50±720+10MeV,Ωc(3327):M=3327.1±1.21.3+0.1±0.2MeV,Γ=20±51+13MeV.

As early as 2005, the Belle Collaboration tentatively assigned the Σc0(2800), Σc+(2800) and Σc++(2800) in the Λc+π/0/+ mass spectra as the isospin triplet states with the spin-parity JP=32 [756], the measured masses and decay widths are

Σc0(2800):M=MΛ++515.43.1+3.26.0+2.1MeV,Γ=6113+1813+22MeV,

Σc+(2800):M=MΛ++505.44.6+5.82.0+12.4MeV,Γ=6223+3738+52MeV,Σc++(2800):M=MΛ++514.53.1+3.44.9+2.8MeV,Γ=7513+1811+12MeV.

In 2008, the BaBar Collaboration observed the Σc0(2800) in the Λc+π mass spectrum with the possible spin-parity JP=12 [757], the mass and decay width are

Σc0(2800):M=2846±8±10MeV,Γ=8622+33±7MeV.

In 2007, the BaBar Collaboration observed the Λc(2940) in the D0p invariant mass spectrum [758]. Subsequently, the Belle Collaboration verified the Λc(2940) in the decay mode Λc(2940)Σc(2455)π [759]. In 2017, the LHCb Collaboration determined the spin-parity of the Λ(2940)+ to be JP=32 by analyzing the process Λb0D0pπ [760]. The measured masses and decay widths are

Λc(2940):M=2939.8±1.3±1.0MeV,Γ=17.5±5.0±5.9MeV(BaBar),Λc(2940):M=2938.0±1.34.0+2.0MeV,Γ=135+87+27MeV(Belle),Λc(2940):M=2944.82.5+3.5±0.44.6+0.1MeV,Γ=27.76.0+8.2±0.910.4+5.2MeV(LHCb).

In 2023, the Belle Collaboration studied the decays B¯0Σc(2455)0,++π±p¯ and found a new structure Λc+(2910) in the Σc(2455)0,++π± mass spectrum [761], the mass and decay width are

Λc(2910):M=2913.8±5.6±3.8MeV,Γ=51.8±20.0±18.8MeV.

The Σc(2800), Λc(2940) and Λc(2910) lie near the D()N thresholds, and they might be the D()N molecular states. For example, we can assign the Σc(2800) as the DN molecular (bound) state [762-765], and assign the Λc(2940) as the DN molecular state [763, 765-770]. We would like to study the 11 type charmed pentaquark states and explore the possible assignments in the scenario of molecular states.

Again, we resort to the correlation functions Π(p), Πμν(p) and Πμναβ(p) defined in Eq. (269), and write down the currents,

J(x)=J1,0DN(x),J0,0DN(x),J1,0DΞ(x),J0,0DΞ(x),J12,±12DsΞ(x),Jμ(x)=J1,0DN(x),J0,0DN(x),J1,0DΞ(x),J0,0DΞ(x),J12,±12DsΞ(x),J1,0DΞ(x),J0,0DΞ(x),J12,±12DsΞ(x),Jμν(x)=J1,0DΞ(x),J0,0DΞ(x),J12,±12DsΞ(x),

J1,0DN(x)=12J12,12D0(x)J12,12N+(x)+12J12,12D+(x)J12,12N0(x),J0,0DN(x)=12J12,12D0(x)J12,12N+(x)12J12,12D+(x)J12,12N0(x),J1,0DΞ(x)=12J12,12D0(x)J12,12Ξ0(x)+12J12,12D+(x)J12,12Ξ(x),J0,0DΞ(x)=12J12,12D0(x)J12,12Ξ0(x)12J12,12D+(x)J12,12Ξ(x),J12,±12DsΞ(x)=J0,0Ds+(x)J12,±12Ξ0/(x),

J1,0DN(x)=12J12,12D0(x)J12,12N+(x)+12J12,12D+(x)J12,12N0(x),J0,0DN(x)=12J12,12D0(x)J12,12N+(x)12J12,12D+(x)J12,12N0(x),J1,0DΞ(x)=12J12,12D0(x)J12,12Ξ0(x)+12J12,12D+(x)J12,12Ξ(x),J0,0DΞ(x)=12J12,12D0(x)J12,12Ξ0(x)12J12,12D+(x)J12,12Ξ(x),J12,±12DsΞ(x)=J0,0Ds+(x)J12,±12Ξ0/(x),J1,0DΞ(x)=12J12,12D0(x)J12,12Ξ0(x)+12J12,12D+(x)J12,12Ξ(x),J0,0DΞ(x)=12J12,12D0(x)J12,12Ξ0(x)12J12,12D+(x)J12,12Ξ(x),J12,±12DsΞ(x)=J0,0Ds+(x)J12,±12Ξ0/(x),

J1,0DΞ(x)=12J12,12D0(x)J12,12Ξ0(x)+12J12,12D+(x)J12,12Ξ(x),J0,0DΞ(x)=12J12,12D0(x)J12,12Ξ0(x)12J12,12D+(x)J12,12Ξ(x),J12,±12DsΞ(x)=J0,0Ds+(x)J12,±12Ξ0/(x),

and

J12,12D0(x)=u¯(x)iγ5c(x),J12,12D+(x)=d¯(x)iγ5c(x),J0,0Ds+(x)=s¯(x)iγ5c(x),J12,12D0(x)=u¯(x)γμc(x),J12,12D+(x)=d¯(x)γμc(x),J0,0Ds+(x)=s¯(x)γμc(x),

J12,12N+(x)=εijkuiT(x)Cγμuj(x)γμγ5dk(x),J12,12N0(x)=εijkdiT(x)Cγμdj(x)γμγ5uk(x),J12,12Ξ0(x)=εijksiT(x)Cγμsj(x)γμγ5uk(x),J12,12Ξ(x)=εijksiT(x)Cγμsj(x)γμγ5dk(x),J12,12Ξ0(x)=εijksiT(x)Cγμsj(x)uk(x),J12,12Ξ(x)=εijksiT(x)Cγμsj(x)dk(x),

the subscripts |1,0, |0,0, |12,12 and |12,12 are isospin indexes |I,I3, we choose the convention |12,12=d¯ [771].

The currents J(0), Jμ(0) and Jμν(0) couple potentially to the JP=12, 32 and 52 singly-charmed molecular states P12, P32, and P52, respectively, for more details, see Section 5.1. At the hadron side, we isolate the ground state contributions,

Π(p)=λ122p+MM2p2+λ12+2pM+M+2p2+=Π121(p2)p+Π120(p2),

Πμν(p)=λ322p+MM2p2(gμν+γμγν3+2pμpν3p2pμγνpνγμ3p2)+λ32+2pM+M+2p2(gμν+γμγν3+2pμpν3p2pμγνpνγμ3p2)+=Π321(p2)pgμνΠ320(p2)gμν+,

Πμναβ(p)=λ522p+MM2p2[g~μαg~νβ+g~μβg~να2g~μνg~αβ5110(γαγμ+)g~νβ+]+λ52+2pM+M+2p2[g~μαg~νβ+g~μβg~να2g~μνg~αβ5+]+=Π521(p2)pgμαgνβ+gμβgνα2+Π520(p2)gμαgνβ+gμβgνα2+.

We choose the components Π121(p2), Π120(p2), Π321(p2), Π320(p2), Π521(p2) and Π520(p2) to explore the spin-parity JP=12, 32 and 52 molecular states, respectively. Again see Section 5.1 for technical details in tensor analysis.

Again we obtain the spectral densities through dispersion relation,

ImΠj1(s)π=λj2δ(sM2)+λj+2δ(sM+2)=ρj,H1(s),ImΠj0(s)π=Mλj2δ(sM2)M+λj+2δ(sM+2)=ρj,H0(s),

where j=12, 32, 52, the subscript H denotes the hadron side, then we introduce the weight functions sexp(sT2) and exp(sT2) to obtain the QCD sum rules at the hadron side,

mc2s0ds[sρj,H1(s)+ρj,H0(s)]exp(sT2)=2Mλj2exp(M2T2),

which are free from contaminations of the positive-parity molecular states.

At the QCD side, we accomplish the operator product expansion with the full light/heavy-quark propagators up to the vacuum condensates of dimension 13 in a consistent way, and obtain the QCD spectral densities through dispersion relation,

ImΠj1/0(s)π=ρj,QCD1/0(s),

where j=12, 32, 52. See Section 3.1.1 for technical details.

At last, we obtain the QCD sum rules,

2Mλj2exp(M2T2)=mc2s0ds[sρj,QCD1(s)+ρj,QCD0(s)]exp(sT2).

We differentiate Eq. (360) with respect to τ=1T2, then eliminate the pole residues to obtain the QCD sum rules for the masses,

M2=ddτmc2s0ds[sρQCD1(s)+ρQCD0(s)]exp(τs)mc2s0ds[sρQCD1(s)+ρQCD0(s)]exp(τs),

where the spectral densities ρQCD1(s)=ρj,QCD1(s) and ρQCD0(s)=ρj,QCD0(s).

We take the quark flavor number nf=4 and resort to the modified energy scale formula

μ=MP2Mc2kMs,

to choose the ideal energy scales of the QCD spectral densities, where MP=M, the effective c-quark mass Mc=1.82GeV and effective s-quark mass Ms=0.2GeV, the k is the number of the s-quark in the currents/states, see Section 4.1 for details.

Routinely, we obtain the Borel windows, continuum threshold parameters, energy scales of the QCD spectral densities and contributions of the D(13), see Tab.57, where the pole dominance is well satisfied. On the other hand, the highest dimensional vacuum condensate contributions have the relation |D(11)||D(12)|>|D(13)|, the operator product expansion is convergent very well.

At last, we obtain the masses and pole residues of the pentaquark molecular states, see Tab.58 [771]. From Tab.57 and Tab.58, it is obvious that the modified energy scale formula μ=MP2Mc2kMs is satisfied in most cases. The energy scales MP2Mc23Ms for the DsΞ, DsΞ and DsΞ states are smaller than the energy scales MP2Mc22Ms for the corresponding DΞ, DΞ and DΞ states, respectively, we choose the same energy scales MP2Mc22Ms for those cousins, as larger masses correspond to larger energy scales naively.

The predicted masses 2.810.16+0.13GeV and 3.190.13+0.11GeV for the DN and DΞ molecular states (respectively) with the JP=12 support assigning the Σc(2800) and Ωc(3185) as the DN and DΞ molecular states, respectively. The predicted mass 3.350.13+0.11GeV for the DΞ molecular state with the JP=32 supports assigning the Ωc(3327) as the DΞ molecular state. The predicted mass 2.960.13+0.14GeV for the DN molecular state is consistent with the Λc(2940)/Λc(2910) within uncertainties, which does not exclude the possibility that they are molecular states.

In Ref. [767], Zhang studied the Σc(2800) and Λc(2940) as the S-wave DN and DN molecule candidates respectively with the QCD sum rules, obtains the masses 3.64±0.33GeV and 3.73±0.35GeV for the S-wave DN state with the JP=12 and S-wave DN state with the JP=32, respectively, which are somewhat bigger than the experimental data for the Σc(2800) and Λc(2940), respectively, and differ from the present calculations significantly. We should bear in mind that Zhang chose Scheme II while we choose Scheme I.

Taking the DΞ (DΞ) pentaquark molecular states with the JP=12 (32) as an example, we obtain a symmetric isotriplet |1,1, |1,0, |1,1 and an antisymmetric isosinglet |0,0. The Ωc(3185) (Ωc(3327)) is a good candidate for the DΞ (DΞ) molecular state with the isospin |0,0. We expect to search for the molecular states with the isospin |1,1 and |1,1 in the Ξc+K¯0 and Ξc0K¯ mass spectra respectively to shed light on the nature of the Ωc(3185) (Ωc(3327)).

Other interpretations also exist, the Σc(2800) could be assigned as overlap of the Σc(2813) and Σc(2840) [772] or P-wave charmed baryon state with the JP=12/32/52 [773-776]. Cheng et al suggest that the Σc(2800) is not possible to be a JP=12 charmed baryon state [777]. The Λc+(2940) is most likely to be the JP=12 or 32 (2P) state [776-779]. And the Λc+(2910) is probably the JP=12 (2P) [780] or 52 (1P) state [781]. The Ωc(3185) lies in the region of the 2S state [782, 783], while the Ωc(3327) is a good candidate for the 2S Ωc state [783] or D-wave Ωc baryon state with the JP=12+/32+/52+ [782, 784, 785].

With a simple replacement cb, we obtain the corresponding QCD sum rules for the singly-bottom pentaquark molecular states, see Eqs. (360) and (361). In Ref. [786], we choose the current,

J(x)=u¯(x)iγ5s(x)εijkuiT(x)Cγαdj(x)γαγ5bk(x),

to interpolate the molecular states with the JP=12±, the prediction supports assigning the Ξb(6227) as the pentaquark molecular state with the JP=12.

In Ref. [787], we construct the 3¯3¯3¯ type five-quark currents,

JSS(x)=εilaεijkεlmnsjT(x)Cγ5uk(x)smT(x)Cγ5cn(x)Cu¯aT(x),JAA(x)=εilaεijkεlmnsjT(x)Cγμuk(x)smT(x)Cγμcn(x)Cu¯aT(x),

to interpolate the singly-charmed pentaquark states suscu¯ with the JP=12±. With a simple replacement,

suscu¯12(suscu¯+sdscd¯),

we obtain the isospin singlet current, the expressions of the QCD sum rules survive. The predictions support assigning the excited Ωc states from the LHCb Collaboration as the SS or AA-type pentaquark states with the JP=12 and the mass about 3.08GeV. We group the quark flavors as [su][sc]u¯ to construct the currents, if we group the quark flavors as [ss][cu]u¯, the SS-type current will vanish due to Fermi-Dirac statistics, but the AA-type current survives and leads to almost degenerated mass according to the light-flavor SU(3) symmetry.

In Ref. [788], we study the 3¯3¯3¯ type singly-charmed pentaquark states with the JP=32± via the QCD sum rules. We distinguish the isospins in constructing the interpolating currents via two clusters, a diquark qjTCγμqk (Di) plus a triquark qmTCγ5cnCq¯aT (Tmna), which have the properties,

I^2εijkqjTCγμqj=1(1+1)εijkqjTCγμqj,I^2εijkqjTCγμsj=12(12+1)εijkqjTCγμsj,I^2εijksjTCγμsj=0(0+1)εijksjTCγμsj,

I^2umTCγ5cnCd¯aT=1(1+1)umTCγ5cnCd¯aT,I^2dmTCγ5cnCu¯aT=1(1+1)dmTCγ5cnCu¯aT,I^2[umTCγ5cnCu¯aTdmTCγ5cnCd¯aT]=1(1+1)[umTCγ5cnCu¯aTdmTCγ5cnCd¯aT],I^2[umTCγ5cnCu¯aT+dmTCγ5cnCd¯aT]=0(0+1)[umTCγ5cnCu¯aT+dmTCγ5cnCd¯aT],I^2qmTCγ5cnCs¯aT=12(12+1)qmTCγ5cnCs¯aT,I^2smTCγ5cnCq¯aT=12(12+1)smTCγ5cnCq¯aT,I^2smTCγ5cnCs¯aT=0(0+1)smTCγ5cnCs¯aT,

with q, q=u, d, the I^2 is the isospin operator. The diquark clusters Di and triquark clusters Tmna have the isospins I=1, 12 or 0. We could obtain some mass relations based on the SU(3) breaking effects of the u, d, s quarks, the mass relation among the diquark clusters Di is mI=0mI=12=mI=12mI=1, while the mass relation among the triquark clusters Tmna is mI=0mI=12=mI=12mI=1, if the hidden-flavor (for u and d) isospin singlet umTCγ5cnCu¯aT+dmTCγ5cnCd¯aT is excluded. Moreover, the isospin triplet umTCγ5cnCu¯aTdmTCγ5cnCd¯aT and isospin singlet umTCγ5cnCu¯aT+dmTCγ5cnCd¯aT are expected to have degenerate masses, which can be inferred from the tiny mass difference between the vector mesons ρ0(770) and ω(780). In fact, if we choose the currents Jμ(x)=u¯(x)γμu(x)d¯(x)γμd(x) and u¯(x)γμu(x)+d¯(x)γμd(x) to interpolate the ρ0(770) and ω(780), respectively, we obtain the same QCD sum rules.

We write down the currents explicitly,

Jqqqq¯,μ(x)=εilaεijkεlmnqjT(x)Cγμqk(x)qmT(x)Cγ5cn(x)Cq¯aT(x),

and study the singly-charmed pentaquark states uuucu¯ and ssscs¯ with the QCD sum rules in details. Then we estimate the masses of the singly-charmed pentaquark states ssucu¯, suscu¯, ssdcd¯ and sdscd¯ with JP=32 to be 3.15±0.13GeV according to the light-flavor SU(3) breaking effects, which is compatible with the experimental values of the masses of the Ωc(3050), Ωc(3066), Ωc(3090), Ωc(3119).

7 Strong decays of exotic states

In Refs. [520, 789], we suggest rigorous quark-hadron duality to calculate the hadronic coupling constants in the two-body strong decays of the tetraquark states with the QCD sum rules. At first, we write down the three-point correlation functions Π(p,q),

Π(p,q)=i2d4xd4yeipxeiqy0|T{JB(x)JC(y)JA(0)}|0,

where the currents JA(0) interpolate the tetraquark states A, the currents JB(x) and JC(y) interpolate the conventional mesons B and C, respectively,

0|JA(0)|A(p)=λA,0|JB(0)|B(p)=λB,0|JC(0)|C(q)=λC,

the λA, λB and λC are the pole residues or decay constants.

At the hadron side, we insert a complete set of intermediate hadronic states with the same quantum numbers as the currents JA(0), JB(x), JC(y) into the three-point correlation functions Π(p,q) and isolate the ground state contributions to obtain the result [789],

Π(p,q)=λAλBλCGABC(mA2p2)(mB2p2)(mC2q2)+1(mA2p2)(mB2p2)sC0dtρAC(p2,p2,t)tq2+1(mA2p2)(mC2q2)sB0dtρAB(p2,t,q2)tp2+1(mB2p2)(mC2q2)sA0dtρAB(t,p2,q2)+ρAC(t,p2,q2)tp2+=Π(p2,p2,q2),

where p=p+q, the GABC are the hadronic coupling constants defined by

B(p)C(q)|A(p)=iGABC,

the four functions ρAC(p2,p2,t), ρAB(p2,t,q2), ρAB(t,p2,q2) and ρAC(t,p2,q2) have complex dependence on the transitions between the ground states and the higher resonances or continuum states.

We rewrite the correlation functions ΠH(p2,p2,q2) at the hadron side as

ΠH(p2,p2,q2)=Δ2sA0dsΔs2sB0dsΔu2uC0duρH(s,s,u)(sp2)(sp2)(uq2)+sA0dsΔs2sB0dsΔu2uC0duρH(s,s,u)(sp2)(sp2)(uq2)+,

through triple-dispersion relation, where the ρH(s,s,u) are the hadronic spectral densities,

ρH(s,s,u)=limϵ30limϵ20limϵ10ImsImsImuΠH(s+iϵ3,s+iϵ2,u+iϵ1)π3,

where the Δ2, Δs2 and Δu2 are the thresholds, the sA0, sB0, uC0 are the continuum thresholds.

Now we carry out the operator product expansion at the QCD side, and write the correlation functions ΠQCD(p2,p2,q2) as

ΠQCD(p2,p2,q2)=Δs2sB0dsΔu2uC0duρQCD(p2,s,u)(sp2)(uq2)+,

through double-dispersion relation, where the ρQCD(p2,s,u) are the QCD spectral densities,

ρQCD(p2,s,u)=limϵ20limϵ10ImsImuΠQCD(p2,s+iϵ2,u+iϵ1)π2.

As the QCD spectral densities ρQCD(s,s,u) do not exist,

ρQCD(s,s,u)=limϵ30limϵ20limϵ10ImsImsImuΠQCD(s+iϵ3,s+iϵ2,u+iϵ1)π3=0,

because

limϵ30ImsΠQCD(s+iϵ3,p2,q2)π=0.

Thereafter we will write the QCD spectral densities ρQCD(p2,s,u) as ρQCD(s,u) for simplicity.

We match the hadron side with the QCD side of the correlation functions, and accomplish the integral over ds firstly to obtain the rigorous quark-hadron duality [520],

Δs2sB0dsΔu2uC0duρQCD(s,u)(sp2)(uq2)=Δs2sB0dsΔu2uC0du1(sp2)(uq2)[Δ2dsρH(s,s,u)sp2],

the Δ2 denotes the thresholds (Δs+Δu)2. Now we write down the quark-hadron duality explicitly,

Δs2sB0dsΔu2uC0duρQCD(s,u)(sp2)(uq2)=Δs2sB0dsΔu2uC0duΔ2dsρH(s,s,u)(sp2)(sp2)(uq2)=λAλBλCGABC(mA2p2)(mB2p2)(mC2q2)+CAB+CAC(mB2p2)(mC2q2).

No approximation is needed, we do not need the continuum threshold parameter sA0 in the s channel, as we match the hadron side with the QCD side below the continuum thresholds s0 and u0 to obtain rigorous quark-hadron duality, and we take account of the continuum contributions in the s channel.

In Eq. (380), we introduce the parameters CAC, CAB, CAB and CAC to parameterize the net effects,

CAC=sC0dtρAC(p2,p2,t)tq2,CAB=sB0dtρAB(p2,t,q2)tp2,CAB=sA0dtρAB(t,p2,q2)tp2,CAC=sA0dtρAC(t,p2,q2)tp2.

In numerical calculations, we take the relevant functions CAB and CAC as free parameters, and choose suitable values to eliminate the contaminations from the higher resonances and continuum states to obtain the stable QCD sum rules with the variations of the Borel parameters T2.

According to the discussions in Section 2.3, the quantum field theory does not forbid the couplings between the four-quark currents JA(0) and two-meson scattering states BC, if they have the same quantum numbers. The local currents JA(0) have direct non-vanishing couplings to the two-meson scattering states BC, although the overlaps of the wave-functions are very small [417], which leads to a finite width to modify the dispersion relation, see Eqs. (78) and (79).

There exists another term ΠHD(p2,p2,q2) at the hadron side beyond that shown in Eq. (380),

ΠHD(p2,p2,q2)=λBλCλBC(mB2p2)(mC2q2)+,

where

0|JA(0)|B(p)C(q)=λBC.

Such terms ΠHD(p2,p2,q2) shown in Eq. (382) could be absorbed into the parameters CAB+CAC with the simple replacement,

CAB+CACCAB+CAC+λBλCλBC.

In Ref. [446, 790], Nielsen et al approximate the hadron side of the correlation functions as

ΠH(p2,p2,q2)=λAλBλCGABC(mA2p2)(mB2p2)(mC2q2)+BH(s0p2)(mC2q2),

then match them with the QCD side below the continuum threshold s0 by taking the chiral limit mC20 and q20 sequentially, where the BH stands for the pole-continuum transitions, and we have rewritten their notations into the present form for convenience. Although Nielsen et al take account of the continuum contributions by introducing a parameter s0 in the s channel phenomenologically, they neglect the continuum contributions in the u channel at the hadron side by hand. Such an approximation is also adopted in Refs. [47, 532, 791, 792].

There is another scheme to study the strong hadronic coupling constants with the correlation functions Π(p,p),

Π(p,p)=i2d4xd4yeipxeipy0|T{JB(y)JC(0)JA(x)}|0.

At the QCD side, a double-dispersion relation,

Π(p2,p2,q2)=Δs2s0dsΔs2s0dsρ(s,s,q2)(sp2)(sp2)+,

with q=pp is adopted [465, 472]. However, we cannot obtain the QCD spectral densities ρ(s,s,q2). Such an scheme is also adopted in the case of pentaquark states [793-796].

Let us turn to Eq. (380) again. If the B are charmonium or bottomonium states, we set p2=p2 and perform the double Borel transformation with respect to the variables P2=p2 and Q2=q2, respectively to obtain the QCD sum rules,

λAλBλCGABCmA2mB2[exp(mB2T12)exp(mA2T12)]exp(mC2T22)+(CAB+CAC)exp(mB2T12mC2T22)=Δs2sB0dsΔu2uC0duρQCD(s,u)exp(sT12uT22),

where the T12 and T22 are the Borel parameters. If the B are open-charm or open-bottom mesons, we set p2=4p2 and perform the double Borel transformation with respect to the variables P2=p2 and Q2=q2, respectively to obtain the QCD sum rules,

λAλBλCGABC4(m~A2mB2)[exp(mB2T12)exp(m~A2T12)]exp(mC2T22)+(CAB+CAC)exp(mB2T12mC2T22)=Δs2sB0dsΔu2uC0duρQCD(s,u)exp(sT12uT22),

where m~A2=mA24. Or set p2=p2, just like in the first case. The scheme based on the rigorous quark-hadron duality is adopted in Refs. [62, 171, 507, 523, 797-806].

7.1 Strong decays of the Y(4500) as an example

In this sub-section, we would like to use a typical example to illustrate the procedure in details.

After Ref. [543] was published, the Y(4500) was observed by the BESIII Collaboration [155, 156, 159]. At the energy about 4.5GeV, we obtain three hidden-charm tetraquark states with the JPC=1, the [uc]V~[uc¯]A+[dc]V~[dc¯]A[uc]A[uc¯]V~[dc]A[dc¯]V~, [uc]A~[uc¯]V+[dc]A~[dc¯]V+[uc]V[uc¯]A~+[dc]V[dc¯]A~ and [uc]S[uc¯]V~+[dc]S[dc¯]V~[uc]V~[uc¯]S[dc]V~[dc¯]S tetraquark states have the masses 4.53±0.07GeV, 4.48±0.08GeV and 4.50±0.09GeV, respectively [543], see Tab.22. In Ref. [546], we study their two-body strong decays systematically with the three-point correlation functions,

ΠμD¯DA~V(p,q)=i2d4xd4yeipxeiqy0|T{JD¯(x)JD(y)J,μA~V(0)}|0,

ΠαμD¯DA~V(p,q)=i2d4xd4yeipxeiqy0|T{JαD¯(x)JD(y)J,μA~V(0)}|0,

ΠαβμD¯DA~V(p,q)=i2d4xd4yeipxeiqy0|T{JαD¯(x)JβD(y)J,μA~V(0)}|0,

ΠαμD¯0DA~V(p,q)=i2d4xd4yeipxeiqy0|T{JD¯0(x)JαD(y)J,μA~V(0)}|0,

ΠαμD¯1DA~V(p,q)=i2d4xd4yeipxeiqy0|T{JαD¯1(x)JD(y)J,μA~V(0)}|0,

ΠαμηcωA~V(p,q)=i2d4xd4yeipxeiqy0|T{Jηc(x)Jαω(y)J,μA~V(0)}|0,

ΠαβμJ/ψωA~V(p,q)=i2d4xd4yeipxeiqy0|T{JαJ/ψ(x)Jβω(y)J,μA~V(0)}|0,

Παμχc0ωA~V(p,q)=i2d4xd4yeipxeiqy0|T{Jχc0(x)Jαω(y)J,μA~V(0)}|0,

Παβμχc1ωA~V(p,q)=i2d4xd4yeipxeiqy0|T{Jαχc1(x)Jβω(y)J,μA~V(0)}|0,

ΠαμJ/ψf0A~V(p,q)=i2d4xd4yeipxeiqy0|T{JαJ/ψ(x)Jf0(y)J,μA~V(0)}|0.

With the simple replacement A~VV~A, we obtain the corresponding correlation functions for the current J,μV~A. And with the simple replacements A~VSV~ and μμν, we obtain the corresponding correlation functions for the current J,μνSV~, where the currents

JD¯(x)=c¯(x)iγ5u(x),JD(y)=u¯(y)iγ5c(y),JαD¯(x)=c¯(x)γαu(x),JβD(y)=u¯(y)γβc(y),JD¯0(x)=c¯(x)u(x),JαD¯1(x)=c¯(x)γαγ5u(x),

Jηc(x)=c¯(x)iγ5c(x),JαJ/ψ(x)=c¯(x)γαc(x),Jχc0(x)=c¯(x)c(x),Jαχc1(x)=c¯(x)γαγ5c(x),

Jαω(y)=u¯(y)γαu(y)+d¯(y)γαd(y)2,Jf0(y)=u¯(y)u(y)+d¯(y)d(y)2,

J,μA~V(x)=εijkεimn2[ujT(x)Cσμνγ5ck(x)u¯m(x)γ5γνCc¯nT(x)+djT(x)Cσμνγ5ck(x)d¯m(x)γ5γνCc¯nT(x)+ujT(x)Cγνγ5ck(x)u¯m(x)γ5σμνCc¯nT(x)+djT(x)Cγνγ5ck(x)d¯m(x)γ5σμνCc¯nT(x)],

J,μV~A(x)=εijkεimn2[ujT(x)Cσμνck(x)u¯m(x)γνCc¯nT(x)+djT(x)Cσμνck(x)d¯m(x)γνCc¯nT(x)ujT(x)Cγνck(x)u¯m(x)σμνCc¯nT(x)djT(x)Cγνck(x)d¯m(x)σμνCc¯nT(x)],

J,μνSV~(x)=εijkεimn2[ujT(x)Cγ5ck(x)u¯m(x)σμνCc¯nT(x)+djT(x)Cγ5ck(x)d¯m(x)σμνCc¯nT(x)ujT(x)Cσμνck(x)u¯m(x)γ5Cc¯nT(x)djT(x)Cσμνck(x)d¯m(x)γ5Cc¯nT(x)].

According to quark-hadron duality, we obtain the hadron representation and isolate the ground state contributions explicitly,

ΠμD¯DA~V(p,q)=ΠD¯DA~V(p2,p2,q2)i(pq)μ+,

ΠαμD¯DA~V(p,q)=ΠD¯DA~V(p2,p2,q2)(iεαμλτpλqτ)+,

ΠαβμD¯DA~V(p,q)=ΠD¯DA~V(p2,p2,q2)(igαβpμ)+,

ΠαμD¯0DA~V(p,q)=ΠD¯0DA~V(p2,p2,q2)(igαμpq)+,

ΠαμD¯1DA~V(p,q)=ΠD¯1DA~V(p2,p2,q2)gαμ+,

ΠαμηcωA~V(p,q)=ΠηcωA~V(p2,p2,q2)(iεαμλτpλqτ)+,

ΠαβμJ/ψωA~V(p,q)=ΠJ/ψωA~V(p2,p2,q2)(igαβpμ)+,

Παμχc0ωA~V(p,q)=Πχc0ωA~V(p2,p2,q2)igαμ+,

Παβμχc1ωA~V(p,q)=Πχc1ωA~V(p2,p2,q2)(εαβμλpλpq)+,

ΠαμJ/ψf0A~V(p,q)=ΠJ/ψf0A~V(p2,p2,q2)(igαμ)+,

other ground state contributions are given in Ref. [546], where

ΠD¯DA~V(p2,p2,q2)=λD¯DA~V(mY2p2)(mD2p2)(mD2q2)+CD¯DA~V(mD2p2)(mD2q2)+,

ΠD¯DA~V(p2,p2,q2)=λD¯DA~V(mY2p2)(mD2p2)(mD2q2)+CD¯DA~V(mD2p2)(mD2q2)+,

ΠD¯DA~V(p2,p2,q2)=λD¯DA~V(mY2p2)(mD2p2)(mD2q2)+CD¯DA~V(mD2p2)(mD2q2)+,

ΠD¯0DA~V(p2,p2,q2)=λD¯0DA~V(mY2p2)(mD02p2)(mD2q2)+CD¯0DA~V(mD02p2)(mD2q2)+,

ΠD¯1DA~V(p2,p2,q2)=λD¯1DA~V(mY2p2)(mD12p2)(mD2q2)+CD¯1DA~V(mD12p2)(mD2q2)+,

ΠηcωA~V(p2,p2,q2)=ληcωA~V(mY2p2)(mηc2p2)(mω2q2)+CηcωA~V(mηc2p2)(mω2q2)+,

with the simple replacements ηcJ/ψ, χc0 and χc1, we obtain the hadronic representation for the J/ψωA~V, χc0ωA~V and χc1ωA~V channels, respectively,

ΠJ/ψf0A~V(p2,p2,q2)=λJ/ψf0A~V(mY2p2)(mJ/ψ2p2)(mf02q2)+CJ/ψf0A~V(mJ/ψ2p2)(mf02q2)+.

With the simple replacements A~VV~A and SV~, we obtain the corresponding components Π(p2,p2,q2) for the currents J,μV~A(0) and J,μνSV~(0), except for the component ΠηcωSV~(p2,p2,q2),

ΠηcωSV~(p2,p2,q2)=ληcωSV~(mY2p2)(mηc2p2)(mω2q2)+CηcωSV~(mηc2p2)(mω2q2)+λ¯ηcωSA~(mX2p2)(mηc2p2)(mω2q2)+.

We introduce the collective notations to simplify the expressions,

λD¯DA~V=fD2mD4mc2λA~VGD¯DA~V,λD¯DA~V=fDmDfDmD2mcλA~VGD¯DA~V,λD¯DA~V=fD2mD2λA~VGD¯DA~V,λD¯0DA~V=fD0mD0fDmDλA~VGD¯0DA~V,λD¯1DA~V=fD1mD1fDmD2mcλA~VGD¯1DA~V,

ληcωA~V=fηcmηc2fωmω2mcλA~VGηcωA~V,λJ/ψωA~V=fJ/ψmJ/ψfωmωλA~VGJ/ψωA~V(1+mω2mY2mJ/ψ2mY2),λχc0ωA~V=fχc0mχc0fωmωλA~VGχc0ωA~V,λχc1ωA~V=fχc1mχc1fωmωλA~VGχc1ωA~V,λJ/ψf0A~V=fJ/ψmJ/ψff0mf0λA~VGJ/ψf0A~V.

With the simple replacements A~VV~A and SV~, we obtain the collective notations λ for the currents J,μV~A(0) and J,μνSV~(0), except for the ληcωSV~, λJ/ψωSV~, λχc1ωSV~, where

ληcωSV~=fηcmηc2fωmω32mcλSV~GηcωSV~,λJ/ψωSV~=fJ/ψmJ/ψfωmωλSV~GJ/ψωSV~,λχc1ωSV~=fχc1mχc13fωmωλSV~Gχc1ωSV~,

and

λ¯ηcωSA~=fηcmηc2fωmω2mcλ¯SA~G¯ηcωSA~.

We adopt the standard definitions for the decay constants and pole residues [546], and we define the hadronic coupling constants,

D¯(p)D(q)|YA~V(p)=(pq)εGD¯DA~V,D¯(p)D(q)|YV~A(p)=(pq)εGD¯DV~A,D¯(p)D(q)|YSV~(p)=i(pq)εGD¯DSV~,

D¯(p)D(q)|YA~V(p)=ελτρσpλξτpρεσGD¯DA~V,D¯(p)D(q)|YV~A(p)=ελτρσpλξτpρεσGD¯DV~A,D¯(p)D(q)|YSV~(p)=iελτρσpλξτpρεσGD¯DSV~,

etc. [546].

In Eq. (423), there are contributions coming from the JPC=1+ and 1 tetraquark states, and we cannot choose the pertinent structures to exclude the contaminations from the JPC=1+ tetraquark state X, so we include it at the hadron side. The unknown parameters, CD¯DA~V, CD¯DA~V, CD¯DA~V, etc., parameterize the complex interactions among the excitations in the p2 channels and the ground state charmed meson pairs or charmonium plus ω/f0. It is difficult to choose the pertinent tensor structures in Eqs. (406)−(423) to obtain good QCD sum rules without contaminations, and we have to reach the satisfactory results via trial and error.

At the QCD side, we accomplish the operator product expansion up to the vacuum condensates of dimension 5 and take account of both the connected and disconnected Feynman diagrams in the color space, see Fig.31 (in Refs. [47, 446, 532, 790, 791], only the connected Feynman diagrams are taken into account, i.e., the D, E and I in Fig.31), and choose the components ΠH(p2,p2,q2) to study the hadronic coupling constants GH based on the rigorous quark-hadron duality [520, 789].

Then we set p2=p2 in the components Π(p2,p2,q2), and carry out the double Borel transformation with respect to the variables P2=p2 and Q2=q2 respectively, and set T12=T22=T2 to obtain thirty QCD sum rules,

λD¯DA~VmY2mD2[exp(mD2T2)exp(mY2T2)]exp(mD2T2)+CD¯DA~Vexp(mD2T2mD2T2)=ΠD¯DA~VQCD(T2),

λD¯DA~VmY2mD2[exp(mD2T2)exp(mY2T2)]exp(mD2T2)+CD¯DA~Vexp(mD2T2mD2T2)=ΠD¯DA~VQCD(T2),

λD¯DA~VmY2mD2[exp(mD2T2)exp(mY2T2)]exp(mD2T2)+CD¯DA~Vexp(mD2T2mD2T2)=ΠD¯DA~VQCD(T2),

λD¯0DA~VmY2mD02[exp(mD02T2)exp(mY2T2)]exp(mD2T2)+CD¯0DA~Vexp(mD02T2mD2T2)=ΠD¯0DA~VQCD(T2),

λD¯1DA~VmY2mD12[exp(mD12T2)exp(mY2T2)]exp(mD2T2)+CD¯1DA~Vexp(mD12T2mD2T2)=ΠD¯1DA~VQCD(T2),

ληcωA~VmY2mηc2[exp(mηc2T2)exp(mY2T2)]exp(mω2T2)+CηcωA~Vexp(mηc2T2mω2T2)=ΠηcωA~VQCD(T2),

with the simple replacements ηcJ/ψ, χc0 and χc1, we obtain the QCD sum rules for the J/ψωA~V, χc0ωA~V and χc1ωA~V channels, respectively,

λJ/ψf0A~VmY2mJ/ψ2[exp(mJ/ψ2T2)exp(mY2T2)]exp(mf02T2)+CJ/ψf0A~Vexp(mJ/ψ2T2mf02T2)=ΠJ/ψf0A~VQCD(T2).

With the simple replacements A~VV~A and SV~, we obtain the QCD sum rules for the J,μV~A(0) and J,μνSV~(0), except for the ηcω channel,

ληcωSV~mY2mηc2[exp(mηc2T2)exp(mY2T2)]exp(mω2T2)+CηcωSV~exp(mηc2T2mω2T2)+λ¯ηcωSA~mX2mηc2[exp(mηc2T2)exp(mX2T2)]exp(mω2T2)=ΠηcωSV~QCD(T2),

the explicit expressions of the QCD side are given in Ref. [546].

We take the unknown parameters CD¯DA~V, CD¯DA~V, CD¯DA~V, as free parameters, and adjust the suitable values to obtain flat Borel platforms for the hadronic coupling constants [546],

CD¯DA~V=0.00045GeV5×T2,CD¯DA~V=0.000003GeV4×T2,CD¯DA~V=0.0001GeV5×T2,CD¯0DA~V=0.000055GeV4×T2,CD¯1DA~V=0.0031GeV6×T2,CηcωA~V=0.000082GeV4×T2,CJ/ψωA~V=0.0,Cχc0ωA~V=0.00085GeV6×T2,Cχc1ωA~V=0.000019GeV3×T2,CJ/ψf0A~V=0.00085GeV6×T2,

CD¯DV~A=0.0000015GeV5×T2,CD¯DV~A=0.000038GeV4×T2,CD¯DV~A=0.0000054GeV5×T2,CD¯0DV~A=0.00264GeV6×T2,CD¯1DV~A=0.000055GeV4×T2,CηcωV~A=0.00006GeV4×T2,CJ/ωωV~A=0.0,Cχc0ωV~A=0.00087GeV6×T2,Cχc1ωV~A=0.000018GeV3×T2,CJ/ψf0V~A=0.00075GeV6×T2,

CD¯DSV~=0.00014GeV6+0.00002GeV4×T2,CD¯DSV~=0.0000006GeV3×T2,CD¯DSV~=0.000002GeV4×T2,

CD¯0DSV~=0.0012GeV7+0.000164GeV5×T2,CD¯1DSV~=0.0017GeV7+0.000205GeV5×T2,CηcωSV~=0.00014GeV7,λ¯ηcωSA~=0.0012GeV7×T2,CJ/ψωSV~=0.0005GeV60.0000216GeV4×T2,Cχc0ωSV~=0.0018GeV70.000156GeV5×T2,Cχc1ωSV~=0.0012GeV8+0.00015GeV6×T2,CJ/ψf0SV~=0.0005GeV7+0.00006GeV5×T2,

the Borel windows are shown explicitly in Tab.59. We obtain uniform flat platforms Tmax2Tmin2=1GeV2, where the max and min denote the maximum and minimum, respectively. In calculations, we choose quark flavor numbers nf=4, and evolve all the input parameters to the energy scale μ=1GeV. For detailed information about the parameters, one can consult Ref. [546].

In Fig.32, we plot the GD¯1DA~V with variation of the Borel parameter at a large interval as an example, in the Borel window, there appears very flat platform indeed.

We estimate the uncertainties of the hadronic coupling constants routinely. For an input parameter ξ, ξ=ξ¯+δξ, the left side can be written as λA~VfD¯fDGD¯DA~V=λ¯A~Vf¯D¯f¯DG¯D¯DA~V+δλA~VfD¯fDGD¯DA~V, CD¯DA~V=C¯D¯DA~V+δCD¯DA~V, , where

δλA~VfD¯fDGD¯DA~V=λ¯A~Vf¯D¯f¯DG¯D¯DA~V(δfD¯f¯D¯+δfDf¯D+δλA~Vλ¯A~V+δGD¯DA~VG¯D¯DA~V),

. We set δCD¯DA~V=0, δfD¯f¯D¯=δfDf¯D=δλA~Vλ¯A~V=δGD¯DA~VG¯D¯DA~V, approximately.

After taking into account the uncertainties, we obtain the values of the hadronic coupling constants, which are shown explicitly in Tab.59, then we obtain the partial decay widths directly, and show them explicitly in Tab.60.

At last, we saturate the total widths with the summary of partial decay widths,

Γ(YA~V)=241.6±9.0MeV,Γ(YV~A)=210.6±9.4MeV,Γ(YSV~)=229.4±39.9MeV.

The widths of the Y(4484), Y(4469) and Y(4544) are 111.1±30.1±15.2MeV, 246.3±36.7±9.4MeV and 116.1±33.5±1.7MeV, respectively, from the BESIII Collaboration [155, 156, 159], which are compatible with the theoretical predictions in magnitude.

From Tab.60, we obtain the typical decay modes. For the YA~V state, the decays,

YA~VD¯10D0D¯0D102,D¯1D+D¯D1+2,

have the largest partial decay width 59.7±5.5MeV; while the decay,

YA~VJ/ψω,

has zero partial decay width. For the YV~A state, the decays,

YV~AD¯00D0D¯0D002,D¯0D+D¯D0+2,

have the largest partial decay width 66.3±6.1MeV; while the decay,

YV~AJ/ψω,

has zero partial decay width. For the YSV~ state, the decay,

YSV~ηcω,

has the largest partial decay width 96.0±34.8MeV; while the decay,

YSV~J/ψω,

has the partial decay width 18.0±6.3MeV. We can search for the Y(4500) in those typical decays to diagnose its nature.

7.2 Light-cone QCD sum rules for the Y(4500) as an example

The light-cone QCD sum rules have been applied extensively to study the two-body strong decays of the tetraquark (molecular) states [127, 462, 467, 807-811], where only the ground state contributions are isolated, and an unknown parameter is (or not) introduced by hand to parameterize the higher resonance contributions, then the Ioffe-Smilga-type trick,

1T2ddT2,

is (or not) adopted to subtract this parameter [812, 813]. The Ioffe−Smilga-type trick was suggested to deal with the traditional hadrons, where there exists a triangle Feynman diagram. In the case of tetraquark (molecular) states, we deal with two disconnected loop diagrams approximately, see Fig.31, we would like not to resort to the Ioffe−Smilga trick, and write down the higher resonance contributions explicitly.

We would like to use an example to illustrate how to study the strong decays of the tetraquark states via the light-cone QCD sum rules, and write down the three-point correlation function Πμαβ(p),

Πμαβ(p,q)=i2d4xd4yeipxeiqy0|T{JμY(0)JαD+(x)JβD¯0(y)}|π(r),

where

JμY(0)=εijkεimn2[ujT(0)Cσμνγ5ck(0)u¯m(0)γ5γνCc¯nT(0)+ujT(0)Cγνγ5ck(0)u¯m(0)γ5σμνCc¯nT(0)+djT(x)Cσμνγ5ck(0)d¯m(0)γ5γνCc¯nT(0)+djT(0)Cγνγ5ck(0)d¯m(0)γ5σμνCc¯nT(0)],JαD+(y)=d¯(x)γαc(x),JβD¯0(x)=c¯(y)γβu(y),

interpolate the Y(4500), D¯ and D, respectively [547], the |π(r) is the external π state.

At the hadron side, we insert a complete set of intermediate hadronic states with the same quantum numbers as the interpolating currents into the three-point correlation function, and isolate the ground state contributions,

Πμαβ(p,q)=λYfD2MD2iGπrτ+iGYpτ(MY2p2)(MD¯2p2)(MD2q2)ερσλτ(gμρ+pμpρp2)(gασ+pαpσp2)(gλβ+qλqβq2)+,

where p=p+q+r, we adopt the standard definitions for the decay constants λY, fD, fD¯, and define the hadronic coupling constants Gπ and GY,

Yc(p)|D¯(p)D(q)π(r)=GπερσλτερξσζλrτGYερσλτερξσζλpτ,

the εμ, ξα and ζβ are polarization vectors of the Y(4500), D¯ and D, respectively. In the isospin limit, mu=md, fD=fD¯ and MD=MD¯.

We multiply Eq. (452) with the tensor εθωαβ and obtain

Π~μθω(p,q)=εθωαβΠμαβ(p,q)=λYfD2MD2iGπ(gμωrθgμθrω)iGY(gμωpθgμθpω)(MY2p2)(MD¯2p2)(MD2q2)+.

Again, we take the isospin limit, then Π~μθω(p,q)=Π~μθω(q,p), and we write down the relevant components explicitly,

Π~μθω(p,q)=[iΠπ(p2,p2,q2)iΠY(p2,p2,q2)](gμωrθgμθrω)+iΠY(p2,p2,q2)(gμωqθgμθqω)+,

where

Ππ(p2,p2,q2)=λYfD2MD2Gπ(MY2p2)(MD¯2p2)(MD2q2)+,ΠY(p2,p2,q2)=λYfD2MD2GY(MY2p2)(MD¯2p2)(MD2q2)+.

Then we choose the tensor structures gμωrθgμθrω and gμωqθgμθqω to study the hadronic coupling constants Gπ and GY, respectively. We obtain the hadronic spectral densities ρH(s,s,u) through triple dispersion relation,

ΠH(p2,p2,q2)=Δs2dsΔs2dsΔu2duρH(s,s,u)(sp2)(sp2)(uq2),

where the Δs2, Δs2 and Δu2 are the thresholds, and we add the subscript H to represent the hadron side.

We carry out the operator product expansion up to the vacuum condensates of dimension 5 and neglect the tiny gluon condensate contributions [520, 789],

Ππ(p2,q2,q2)=fπmc01duφπ(u)[01dxxx¯Γ(ϵ1)2π2(p2m~c2)ϵ12mcq¯q3(p2mc2)+mc3q¯gsσGq3(p2mc2)3]1(q+ur)2mc2+fπmπ2mu+md01duφ5(u)u¯[01dxxx¯Γ(ϵ1)2π2(p2m~c2)ϵ12mcq¯q3(p2mc2)+mc3q¯gsσGq3(p2mc2)3]1(q+ur)2mc2fπmc2q¯gsσGq3601duφπ(u)1(p2mc2)[(q+ur)2mc2]2+fπmπ2mcq¯gsσGq36(mu+md)01duφ5(u)u¯1(p2mc2)[(q+ur)2mc2]2,

ΠY(p2,q2,q2)=fπmπ2mu+md01duφ5(u)[01dxxx¯Γ(ϵ1)2π2(p2m~c2)ϵ12mcq¯q3(p2mc2)+mc3q¯gsσGq3(p2mc2)3]1(q+ur)2mc2+fπmπ2mcq¯gsσGq36(mu+md)01duφ5(u)1(p2mc2)[(q+ur)2mc2]2+f3πmπ201dxx¯[3Γ(ϵ)8π2(p2m~c2)ϵp22π2(p2m~c2)]1q2mc2f3πmπ201dxxx¯[Γ(ϵ1)2π2(p2m~c2)ϵ1+p2Γ(ϵ)2π2(p2m~c2)ϵ]1(q2mc2)2f3πmπ201dxx[3Γ(ϵ)8π2(p2m~c2)ϵ+p24π2(p2m~c2)]1q2mc2,

where q=q+r, u¯=1u, x¯=1x, m~c2=mc2x, (qur)2mc2=(1u)q2+u(q+r)2uu¯mπ2mc2. And we have used the π light-cone distribution functions [814],

0|d¯(0)γμγ5u(x)|π(r)=ifπrμ01dueiurxφπ(u)+,0|d¯(0)σμνγ5u(x)|π(r)=i6fπmπ2mu+md(rμxνrνxμ)01dueiurxφσ(u),0|d¯(0)iγ5u(x)|π(r)=fπmπ2mu+md01dueiurxφ5(u),

and the approximation,

0|d¯(x1)σμνγ5gsGαβ(x2)u(x3)|π(r)=if3π(rμrαgνβ+rνrβgμαrνrαgμβrμrβgνα),

for the twist-3 quark-gluon light-cone distribution functions with the value f3π=0.0035GeV2 at the energy scale μ=1GeV [814, 815]. Such terms proportional to mπ2 and their contributions are greatly suppressed, and we also neglect the twist-4 light-cone distribution functions due to their small contributions. According to the Gell−Mann−Oakes−Renner relation fπmπ2mu+md=2q¯qfπ, we take account of the Chiral enhanced contributions fully in Eqs. (458) and (459).

In Fig.33, we draw the lowest order Feynman diagrams as an example to illustrate the operator product expansion.

In the soft limit rμ0, (q+r)2=q2, we can set Ππ/Y(p2,q2,q2)=Ππ/Y(p2,q2), then we obtain the QCD spectral densities ρQCD(s,u) through double dispersion relation,

Ππ/YQCD(p2,q2)=Δs2dsΔu2duρQCD(s,u)(sp2)(uq2),

we add the superscript (subscript) QCD to stand for the QCD side.

We match the hadron side with the QCD side below the continuum thresholds s0 and u0 to acquire rigorous quark-hadron duality [520, 789],

Δs2s0dsΔu2u0duρQCD(s,u)(sp2)(uq2)=Δs2s0dsΔu2u0du[Δs2dsρH(s,s,u)(sp2)(sp2)(uq2)],

and we carry out the integral over ds firstly, then

ΠH(p2,p2,q2)=λYfD2MD2Gπ/Y(MY2p2)(MD¯2p2)(MD2q2)+s0dsρ~H(s,MD¯2,MD2)(sp2)(MD¯2p2)(MD2q2)+=λYfD2MD2Gπ/Y(MY2p2)(MD¯2p2)(MD2q2)+Cπ/Y(MD¯2p2)(MD2q2)+,

where ρH(s,s,u)=ρ~H(s,s,u)δ(sMD¯2)δ(uMD2), and we introduce the parameters Cπ/Y to parameterize the contributions concerning the higher resonances and continuum states in the s channel,

Cπ/Y=s0dsρ~H(s,MD¯2,MD2)sp2.

As the strong interactions among the ground states π, D, D¯ and excited Y states are complex, and we have no knowledge about the corresponding four-hadron contact vertex. In practical calculations, we can take the unknown functions Cπ/Y as free parameters and adjust the values to acquire flat platforms for the hadronic coupling constants Gπ/Y with variations of the Borel parameters. Such a method works well in the case of three-hadron contact vertexes [520, 789], and we expect it also works here.

In Eq. (452) and Eq. (454), there exist three poles in the limit p2MY2, p2MD¯2 and q2MD2. According to the relation MYMD¯+MD, we set p2=4q2 and perform double Borel transformation with respect to the variables P2=p2 and Q2=q2 respectively, then we set T12=T22=T2 to obtain two QCD sum rules,

λYDDGπ4(M~Y2MD2)[exp(MD2T2)exp(M~Y2T2)]exp(MD¯2T2)+Cπexp(MD2+MD¯2T2)

=fπmcmc2s0ds01duφπ(u)[12π2xi1dxxx¯(sm~c2)(2mcq¯q3mc3q¯gsσGq6T4)δ(smc2)]exp(s+mc2+uu¯mπ2T2)+fπmπ2mu+mdmc2s0ds01duφ5(u)u¯[12π2xi1dxxx¯(sm~c2)(2mcq¯q3mc3q¯gsσGq6T4)δ(smc2)]exp(s+mc2+uu¯mπ2T2)+fπmc2q¯gsσGq36T201duφπ(u)exp(2mc2+uu¯mπ2T2)fπmπ2mcq¯gsσGq36(mu+md)T201duφ5(u)u¯exp(2mc2+uu¯mπ2T2),

λYDDGY4(M~Y2MD2)[exp(MD2T2)exp(M~Y2T2)]exp(MD¯2T2)+CYexp(MD2+MD¯2T2)=fπmπ2mu+mdmc2s0ds01duφ5(u)[12π2xi1dxxx¯(sm~c2)(2mcq¯q3mc3q¯gsσGq6T4)δ(smc2)]exp(s+mc2+uu¯mπ2T2)fπmπ2mcq¯gsσGq36(mu+md)T201duφ5(u)exp(2mc2+uu¯mπ2T2)f3πmπ22π2mc2s0dsxi1dxx¯[34+sδ(sm~c2)]exp(s+mc2T2)f3πmπ22π2T2mc2s0dsxi1dxxx¯m~c2exp(s+mc2T2)+f3πmπ24π2mc2s0dsxi1dxx[32sδ(sm~c2)]exp(s+mc2T2),

where λYDD=λYfD2MD2, M~Y2=MY24 and xi=mc2s.

In calculations, we fit the free parameters as Cπ=0.00101(T23.6GeV2)GeV4 and CY=0.00089(T23.2GeV2)GeV4 to acquire uniform flat Borel platforms Tmax2Tmin2=1GeV2. The Borel windows are Tπ2=(4.65.6)GeV2 and TY2=(4.45.4)GeV2, where the subscripts π and Y represent the corresponding channels. In Fig.34, we plot the hadronic coupling constants Gπ and GY with variations of the Borel parameters. In the Borel windows, there appear very flat platforms indeed.

We obtain the hadronic coupling constants routinely,

Gπ=15.9±0.5GeV1,GY=10.4±0.6GeV1,

by setting

δλ¯Yf¯Df¯D¯G¯π/Y=λ¯Yf¯Df¯D¯G¯π/Y4δGπ/YG¯π/Y.

It is direct to obtain the partial decay width,

Γ(Y(4500)DD¯π+)=124πMYdk2(2π)4δ4(pkp)d3k(2π)32k0d3p(2π)32p0(2π)4δ4(kqr)d3q(2π)32q0d3r(2π)32r0Σ|T|2=6.430.76+0.80MeV,

where T=Yc(p)|D¯(p)D(q)π(r).

The partial decay width Γ(Y(4500)DD¯π+)=6.430.76+0.80MeV is much smaller than the total width Γ=246.3±36.7±9.4MeV from the BESIII Collaboration [156].

The three-body strong decays of the Y(4230) are also studied in this scheme [816], this scheme could be applied to the two-body strong decays of the tetraquark (molecular) states straightforwardly.

We write down the two-point correlation functions Π(p,r),

Π(p,r)=id4xeipx0|T{JA(0)JB(x)}|P(r),

where the currents JA(0) and JB(x) interpolate the tetraquark (molecular) states and traditional mesons, respectively, the P(r) are external states.

At the hadron side, we obtain

Π(p,r)=λAλBGABP(MA2(p+r)2)(MB2p2)+=ΠH(p2,p2),

where p=p+r, and we rewrite the correlation functions ΠH(p2,p2) as

ΠH(p2,p2)=Δ2sA0dsΔs2sB0dsρH(s,s)(sp2)(sp2)+sA0dsΔs2sB0dsρH(s,s)(sp2)(sp2)+,

through double-dispersion relation, where the ρH(s,s) are the hadronic spectral densities,

ρH(s,s)=limϵ20limϵ10ImsImsΠH(s+iϵ2,s+iϵ1)π2,

where the Δ2 and Δs2 are the thresholds, the sA0 and sB0 are the continuum thresholds.

Then we carry out the operator product expansion (not necessary on the light-cone) at the QCD side, and write the correlation functions ΠQCD(p2,p2) as

ΠQCD(p2,p2)=Δs2sB0dsρQCD(p2,s)sp2+,

through single-dispersion relation, where the ρQCD(p2,s) are the QCD spectral densities,

ρQCD(p2,s)=limϵ10ImsΠQCD(p2,s+iϵ1)π.

As the QCD spectral densities ρQCD(s,s) do not exist,

ρQCD(s,s)=limϵ20limϵ10ImsImsΠQCD(s+iϵ2,s+iϵ1)π2=0,

because

limϵ20ImsΠQCD(s+iϵ2,p2)π=0.

And we will write the QCD spectral densities ρQCD(p2,s) as ρQCD(s) for simplicity.

We match the hadron side with the QCD side of the correlation functions, and accomplish the integral over ds firstly to obtain the rigorous quark-hadron duality [520],

Δs2sB0dsρQCD(s)sp2=Δs2sB0ds1sp2[Δ2dsρH(s,s)sp2].

And we write down the quark-hadron duality explicitly,

Δs2sB0dsρQCD(s)sp2=λAλBGABP(mA2p2)(mB2p2)+CABmB2p2.

Again, we introduce the parameters CAB to parameterize the net effects. In numerical calculations, we take the CAB as free parameters, and choose suitable values to obtain the stable QCD sum rules with the variations of the Borel parameters T2.

8 Conclusion and perspective

At the present time, we can only say confidently that the tetraquark and pentaquark states are established in sense of that there are four and five valence quarks, respectively. The under-structures are still under hot debates, more experimental and theoretical works are still needed before reaching definite conclusion. The QCD sum rules method is a reliable and powerful theoretical tool in studying the multiquark states and has given many successful descriptions, however, the predictions have arbitrariness depending on the treating schemes, only comprehensive and systematic works would work.

References

[1]

M. Gell-Mann, A schematic model of baryons and mesons, Phys. Lett. 8, 214 (1964)

[2]

R. L. Jaffe, Multi-quark hadrons. 1. The phenomenology of Q2Q¯2 mesons, Phys. Rev. D 15, 267 (1977)

[3]

R. L. Jaffe, Multi-quark hadrons. 2. Methods, Phys. Rev. D 15, 281 (1977)

[4]

D. Strottman, Multi-quark baryons and the MIT Bag model, Phys. Rev. D 20, 748 (1979)

[5]

H. J. Lipkin, New possibilities for exotic hadrons: Anticharmed strange baryons, Phys. Lett. B 195, 484 (1987)

[6]

F. Dyson and N. H. Xuong, Y=2 States in Su(6) Theory, Phys. Rev. Lett. 13, 815 (1964)

[7]

H. Clement, On the history of dibaryons and their final observation, Prog. Part. Nucl. Phys. 93, 195 (2017)

[8]

T. Sakai, K. Shimizu, and K. Yazaki, H dibaryon, Prog. Theor. Phys. Suppl. 137, 121 (2000)

[9]

R. L. Jaffe and K. Johnson, Unconventional states of confined quarks and gluons, Phys. Lett. B 60, 201 (1976)

[10]

K. L. Au, D. Morgan, and M. R. Pennington, Meson dynamics beyond the quark model: A study of final state interactions, Phys. Rev. D 35, 1633 (1987)

[11]

T. Barnes, F. E. Close, and F. de Viron, QQ¯G hermaphrodite mesons in the MIT Bag model, Nucl. Phys. B 224, 241 (1983)

[12]

S. K. Choi, . Observation of a narrow charmonium-like state in exclusive B+→K+π+π−J/ψ decays, Phys. Rev. Lett. 91, 262001 (2003)

[13]

J. D. Weinstein and N. Isgur, Do multi-quark hadrons exist?, Phys. Rev. Lett. 48, 659 (1982)

[14]

J. D. Weinstein and N. Isgur, KK¯ molecules, Phys. Rev. D 41, 2236 (1990)

[15]

N. N. Achasov, V. V. Gubin, and V. I. Shevchenko, Production of scalar KK¯ molecules in ϕ radiative decays, Phys. Rev. D 56, 203 (1997)

[16]

M. Boglione and M. R. Pennington, Dynamical generation of scalar mesons, Phys. Rev. D 65, 114010 (2002)

[17]

L. Maiani, F. Piccinini, A. D. Polosa, and V. Riquer, A New look at scalar mesons, Phys. Rev. Lett. 93, 212002 (2004)

[18]

R. L. Jaffe and F. Wilczek, Diquarks and exotic spectroscopy, Phys. Rev. Lett. 91, 232003 (2003)

[19]

T. V. Brito, F. S. Navarra, M. Nielsen, and M. E. Bracco, QCD sum rule approach for the light scalar mesons as four-quark states, Phys. Lett. B 608, 69 (2005)

[20]

Z. G. Wang and W. M. Yang, Analysis the f0(980) and a0(980) mesons as four-quark states with the QCD sum rules, Eur. Phys. J. C 42, 89 (2005)

[21]

Z. G. Wang, W. M. Yang, and S. L. Wan, Analysis the 0++ nonet mesons as four-quark states with the QCD sum rules, J. Phys. G 31, 971 (2005)

[22]

H. J. Lee, A QCD sum rule study of the light scalar meson, Eur. Phys. J. A 30, 423 (2006)

[23]

H. J. Lee and N. I. Kochelev, Instanton interpolating current for sigma-tetraquark, Phys. Lett. B 642, 358 (2006)

[24]

S. Groote, J. G. Korner, and D. Niinepuu, Perturbative O(αs) corrections to the correlation functions of light tetraquark currents, Phys. Rev. D 90, 054028 (2014)

[25]

Z. G. Wang, Analysis of the scalar nonet mesons with QCD sum rules, Eur. Phys. J. C 76, 427 (2016)

[26]

S. S. Agaev, K. Azizi, and H. Sundu, The strong decays of the light scalar mesons f0(500) and f0(980), Phys. Lett. B 784, 266 (2018)

[27]

S. S. Agaev, K. Azizi, and H. Sundu, The nonet of the light scalar tetraquarks: The mesons a0(980) and K0∗(800), Phys. Lett. B 789, 405 (2019)

[28]

H. J. Lee, K. S. Kim, and H. Kim, Testing the tetraquark mixing framework from QCD sum rules for a0(980), Phys. Rev. D 100, 034021 (2019)

[29]

S. Weinberg, Tetraquark mesons in large-N quantum chromodynamics, Phys. Rev. Lett. 110, 261601 (2013)

[30]

C. Amsler and N. A. Tornqvist, Mesons beyond the naive quark model, Phys. Rep. 389, 61 (2004)

[31]

F. E. Close and N. A. Tornqvist, Scalar mesons above and below 1-GeV, J. Phys. G 28, R249 (2002)

[32]

E. Klempt and A. Zaitsev, Glueballs, Hybrids, Multiquarks. Experimental facts versus QCD inspired concepts, Phys. Rep. 454, 1 (2007)

[33]

A. Hosaka, T. Iijima, K. Miyabayashi, Y. Sakai, and S. Yasui, Exotic hadrons with heavy flavors: X, Y, Z, and related states, Prog. Theor. Exp. Phys. 2016, 062C01 (2016)

[34]

H. X. Chen, W. Chen, X. Liu, and S. L. Zhu, The hidden-charm pentaquark and tetraquark states, Phys. Rep. 639, 1 (2016)

[35]

R. F. Lebed, R. E. Mitchell, and E. S. Swanson, Heavy-quark QCD exotica, Prog. Part. Nucl. Phys. 93, 143 (2017)

[36]

A. Ali, J. S. Lange, and S. Stone, Exotics: Heavy pentaquarks and tetraquarks, Prog. Part. Nucl. Phys. 97, 123 (2017)

[37]

A. Esposito, A. Pilloni, and A. D. Polosa, Multiquark resonances, Phys. Rep. 668, 1 (2017)

[38]

F. K. Guo, C. Hanhart, U. G. Meissner, Q. Wang, Q. Zhao, and B. S. Zou, Hadronic molecules, Rev. Mod. Phys. 90, 015004 (2018)

[39]

S. L. Olsen, T. Skwarnicki, and D. Zieminska, Non-standard heavy mesons and baryons: Experimental evidence, Rev. Mod. Phys. 90, 015003 (2018)

[40]

M. Karliner, J. L. Rosner, and T. Skwarnicki, Multiquark States, Ann. Rev. Nucl. Part. Sci. 68, 17 (2018)

[41]

N. Brambilla, S. Eidelman, C. Hanhart, A. Nefediev, C. P. Shen, C. E. Thomas, A. Vairo, and C. Z. Yuan, The XYZ states: Experimental and theoretical status and perspectives, Phys. Rep. 873, 1 (2020)

[42]

Y. R. Liu, H. X. Chen, W. Chen, X. Liu, and S. L. Zhu, Pentaquark and Tetraquark states, Prog. Part. Nucl. Phys. 107, 237 (2019)

[43]

H. X. Chen, W. Chen, X. Liu, Y. R. Liu, and S. L. Zhu, An updated review of the new hadron states, Rept. Prog. Phys. 86, 026201 (2023)

[44]

L. Meng, B. Wang, G. J. Wang, and S. L. Zhu, Chiral perturbation theory for heavy hadrons and chiral effective field theory for heavy hadronic molecules, Phys. Rep. 1019, 1 (2023)

[45]

M. Z. Liu, Y. W. Pan, Z. W. Liu, T. W. Wu, J. X. Lu, and L. S. Geng, Three ways to decipher the nature of exotic hadrons: Multiplets, three-body hadronic molecules, and correlation functions, Phys. Rep. 1108, 1 (2025)

[46]

M. Nielsen, F. S. Navarra, and S. H. Lee, New charmonium states in QCD sum rules: A concise review, Phys. Rep. 497, 41 (2010)

[47]

R. M. Albuquerque, J. M. Dias, K. P. Khemchandani, A. M. Torres, F. S. Navarra, M. Nielsen, and C. M. Zanetti, QCD sum rules approach to the X, Y and Z states, J. Phys. G 46, 093002 (2019)

[48]

K. Abe et al., ., Evidence for X(3872)→γJ/ψ and the sub-threshold decay X(3872)→ωJ/ψ, arXiv:

[49]

B. Aubert, . Search for B+→X(3872)K+, X(3872)→J/ψγ, Phys. Rev. D 74, 071101 (2006)

[50]

B. Aubert, . Evidence for X(3872)→ψ(2S)γ in B±→X(3872)K± decays, and a study of B→cc¯γK, Phys. Rev. Lett. 102, 132001 (2009)

[51]

A. Abulencia, . Analysis of the quantum numbers JPC of the X(3872) particle, Phys. Rev. Lett. 98, 132002 (2007)

[52]

S. K. Choi, . Bounds on the width, mass difference and other properties of X(3872)→π+π−J/ψ decays, Phys. Rev. D 84, 052004 (2011)

[53]

R Aaij, . Determination of the X(3872) meson quantum numbers, Phys. Rev. Lett. 110, 222001 (2013)

[54]

R Aaij, . Quantum numbers of the X(3872) state and orbital angular momentum in its ρ0J/ψ decay, Phys. Rev. D 92, 011102 (2015)

[55]

L. Maiani, F. Piccinini, A. D. Polosa, and V. Riquer, Diquark-antidiquarks with hidden or open charm and the nature of X(3872), Phys. Rev. D 71, 014028 (2005)

[56]

L. Maiani, F. Piccinini, A. D. Polosa, and V. Riquer, The Z(4430) and a new paradigm for spin interactions in tetraquarks, Phys. Rev. D 89, 114010 (2014)

[57]

S. J. Brodsky, D. S. Hwang, and R. F. Lebed, Dynamical picture for the formation and decay of the exotic XYZ mesons, Phys. Rev. Lett. 113, 112001 (2014)

[58]

R. D. Matheus, S. Narison, M. Nielsen, and J. M. Richard, Can the X(3872) be a 1++ four-quark state?, Phys. Rev. D 75, 014005 (2007)

[59]

D. Ebert, R. N. Faustov, and V. O. Galkin, Masses of heavy tetraquarks in the relativistic quark model, Phys. Lett. B 634, 214 (2006)

[60]

Z. G. Wang and T. Huang, Analysis of the X(3872), Zc(3900) and Zc(3885) as axial-vector tetraquark states with QCD sum rules, Phys. Rev. D 89, 054019 (2014)

[61]

Z. G. Wang, Analysis of the hidden-charm tetraquark mass spectrum with the QCD sum rules, Phys. Rev. D 102, 014018 (2020)

[62]

Z. G. Wang, Decipher the width of the X(3872) via the QCD sum rules, Phys. Rev. D 109, 014017 (2024)

[63]

Y. R. Liu, X. Liu, W. Z. Deng, and S. L. Zhu, Is X(3872) really a molecular state?, Eur. Phys. J. C 56, 63 (2008)

[64]

N. A. Tornqvist, Isospin breaking of the narrow charmonium state of Belle at 3872 MeV as a deuson, Phys. Lett. B 590, 209 (2004)

[65]

E. S. Swanson, Short range structure in the X(3872), Phys. Lett. B 588, 189 (2004)

[66]

E. S. Swanson, Diagnostic decays of the X(3872), Phys. Lett. B 598, 197 (2004)

[67]

S. Fleming, M. Kusunoki, T. Mehen, and U. v. Kolck, Pion interactions in the X(3872), Phys. Rev. D 76, 034006 (2007)

[68]

C. Bignamini, B. Grinstein, F. Piccinini, A. D. Polosa, and C. Sabelli, Is the X(3872) production cross section at tevatron compatible with a hadron molecule interpretation?, Phys. Rev. Lett. 103, 162001 (2009)

[69]

F. E. Close and P. R. Page, The D∗0D¯0 threshold resonance, Phys. Lett. B 578, 119 (2004)

[70]

D. Gamermann, J. Nieves, E. Oset, and E. R. Arriola, Couplings in coupled channels versus wave functions: Application to the X(3872) resonance, Phys. Rev. D 81, 014029 (2010)

[71]

M. B. Voloshin, Interference and binding effects in decays of possible molecular component of X(3872), Phys. Lett. B 579, 316 (2004)

[72]

F. K. Guo, C. Hidalgo-Duque, J. Nieves, and M. P. Valderrama, Consequences of heavy quark symmetries for hadronic molecules, Phys. Rev. D 88, 054007 (2013)

[73]

F. K. Guo, C. Hanhart, U. G. Meissner, Q. Wang, and Q. Zhao, Production of the X(3872) in charmonia radiative decays, Phys. Lett. B 725, 127 (2013)

[74]

C. Y. Wong, Molecular states of heavy quark mesons, Phys. Rev. C 69, 055202 (2004)

[75]

M. T. AlFiky, F. Gabbiani, and A. A. Petrov, X(3872): Hadronic molecules in effective field theory, Phys. Lett. B 640, 238 (2006)

[76]

J. Nieves and M. P. Valderrama, The Heavy Quark Spin Symmetry Partners of the X(3872), Phys. Rev. D 86, 056004 (2012)

[77]

C. Hanhart, Y. S. Kalashnikova, A. E. Kudryavtsev, and A. V. Nefediev, Reconciling the X(3872) with the near-threshold enhancement in the D0D¯∗0 final state, Phys. Rev. D 76, 034007 (2007)

[78]

Y. S. Kalashnikova, Coupled-channel model for charmonium levels and an option for X(3872), Phys. Rev. D 72, 034010 (2005)

[79]

P. Artoisenet and E. Braaten, Production of the X(3872) at the Tevatron and the LHC, Phys. Rev. D 81, 114018 (2010)

[80]

Y. B. Dong, A. Faessler, T. Gutsche, and V. E. Lyubovitskij, Estimate for the X(3872)→γJ/ψ decay width, Phys. Rev. D 77, 094013 (2008)

[81]

Z. G. Wang and T. Huang, Possible assignments of the X(3872), Zc(3900) and Zb(10610) as axial-vector molecular states, Eur. Phys. J. C 74, 2891 (2014)

[82]

Z. G. Wang, Analysis of the Hidden-charm Tetraquark molecule mass spectrum with the QCD sum rules, Int. J. Mod. Phys. A 36, 2150107 (2021)

[83]

Q. Xin, Z. G. Wang, and X. S. Yang, Analysis of the X(3960) and related tetraquark molecular states via the QCD sum rules, AAPPS Bull. 32, 37 (2022)

[84]

H. Mutuk, Y. Sarac, H. Gumus, and A. Ozpineci, X(3872) and its heavy quark spin symmetry partners in QCD sum rules, Eur. Phys. J. C 78, 904 (2018)

[85]

E. J. Eichten, K. Lane, and C. Quigg, New states above charm threshold, Phys. Rev. D 73, 014014 (2006)

[86]

T. Barnes and S. Godfrey, Charmonium options for the X(3872), Phys. Rev. D 69, 054008 (2004)

[87]

M. Suzuki, The X(3872) boson: Molecule or charmonium, Phys. Rev. D 72, 114013 (2005)

[88]

B. Q. Li and K. T. Chao, Higher Charmonia and X, Y, Z states with screened potential, Phys. Rev. D 79, 094004 (2009)

[89]

D. V. Bugg, Reinterpreting several narrow resonances as threshold cusps, Phys. Lett. B 598, 8 (2004)

[90]

S. K. Choi, . Observation of a near-threshold ωJ/ψ mass enhancement in exclusive B→KωJ/ψ decays, Phys. Rev. Lett. 94, 182002 (2005)

[91]

B. Aubert, . Observation of Y(3940)→J/ψω in B→J/ψωK at BABAR, Phys. Rev. Lett. 101, 082001 (2008)

[92]

P. del Amo Sanchez, . Evidence for the decay X(3872)→J/ψω, Phys. Rev. D 82, 011101 (2010)

[93]

K. Abe, . Observation of a new charmonium state in double charmonium production in e+e− annihilation at s≈10.6GeV, Phys. Rev. Lett. 98, 082001 (2007)

[94]

K. Abe, . Search for new charmonium states in the processes e+e−→J/ψD(∗)D¯(∗) at s≈10.6GeV, Phys. Rev. Lett. 100, 202001 (2008)

[95]

S. Uehara, . Observation of a charmonium-like enhancement in the γγ→ωJ/ψ process, Phys. Rev. Lett. 104, 092001 (2010)

[96]

J. P. Lees, . Study of X(3915)→J/ψω in two-photon collisions, Phys. Rev. D 86, 072002 (2012)

[97]

S. Navas, . The review of particle physics, Phys. Rev. D 110, 030001 (2024)

[98]

S. Uehara, . Observation of a χc2′ candidate in γγ→DD¯ production at Belle, Phys. Rev. Lett. 96, 082003 (2006)

[99]

B. Aubert, . Observation of the χc2(2P) meson in the reaction γγ→DD¯ at BABAR, Phys. Rev. D 81, 092003 (2010)

[100]

K. Chilikin, . Observation of an alternative χc0(2P) candidate in e+e−→J/ψDD¯, Phys. Rev. D 95, 112003 (2017)

[101]

R. Aaij, . Model-independent study of structure in B+→D+D−K+ decays, Phys. Rev. Lett. 125, 242001 (2020)

[102]

R. Aaij, . Amplitude analysis of the B+→D+D−K+ decay, Phys. Rev. D 102, 112003 (2020)

[103]

R. Aaij, . Observation of a resonant structure near the Ds+Ds− threshold in the B+→Ds+Ds−K+ decay, Phys. Rev. Lett. 131, 071901 (2023)

[104]

F. K. Guo and U. G. Meissner, Where is the χc0(2P)?, Phys. Rev. D 86, 091501 (2012)

[105]

T. Ji, X. K. Dong, M. Albaladejo, M. L. Du, F. K. Guo, J. Nieves, and B. S. Zou, Understanding the 0++ and 2++ charmonium(-like) states near 3.9 GeV, Sci. Bull. 68, 688 (2023)

[106]

Z. G. Wang, Analysis of the mass and width of the X∗(3860) with QCD sum rules, Eur. Phys. J. A 53, 192 (2017)

[107]

Z. Y. Zhou, Z. Xiao, and H. Q. Zhou, Could the X(3915) and the X(3930) Be the Same Tensor State?, Phys. Rev. Lett. 115, 022001 (2015)

[108]

T. Barnes, S. Godfrey, and E. S. Swanson, Higher charmonia, Phys. Rev. D 72, 054026 (2005)

[109]

R. Aaij, . Observation of new charmonium(-like) states in B+→D∗±D∓K+, Phys. Rev. Lett. 133, 131902 (2024)

[110]

T. Aaltonen, . Evidence for a Narrow Near-Threshold Structure in the J/ψϕ Mass Spectrum in B+→J/ψϕK+ Decays, Phys. Rev. Lett. 102, 242002 (2009)

[111]

T. Aaltonen, . Observation of the Y(4140) structure in the J/ψϕ Mass Spectrum in B±→J/ψϕK± Decays, Mod. Phys. Lett. A 32, 1750139 (2017)

[112]

S. Chatrchyan, . Observation of a peaking structure in the J/ψϕ mass spectrum from B±→J/ψϕK± decays, Phys. Lett. B 734, 261 (2014)

[113]

R. Aaij, . Observation of J/ψϕ structures consistent with exotic states from amplitude analysis of B+→J/ψϕK+ decays, Phys. Rev. Lett. 118, 022003 (2017)

[114]

R. Aaij, . Amplitude analysis of B+→J/ψϕK+ decays, Phys. Rev. D 95, 012002 (2017)

[115]

R. Aaij, . Observation of new resonances decaying to J/ψK+ and J/ψϕ, Phys. Rev. Lett. 127, 082001 (2021)

[116]

R. Aaij, . Amplitude analysis of B+→ψ(2S)K+π+π− decays, JHEP 01, 054 (2025)

[117]

C. P. Shen, . Evidence for a new resonance and search for the Y(4140) in γγ→ϕJ/ψ, Phys. Rev. Lett. 104, 112004 (2010)

[118]

M. Karliner and J. L. Rosner, First exotic hadron with open heavy flavor: csu¯d¯ tetraquark, Phys. Rev. D 102, 094016 (2020)

[119]

Z. G. Wang, Analysis of the X0(2900) as the scalar tetraquark state via the QCD sum rules, Int. J. Mod. Phys. A 35, 2050187 (2020)

[120]

J. R. Zhang, Open-charm tetraquark candidate: Note on X0(2900), Phys. Rev. D 103, 054019 (2021)

[121]

H. X. Chen, W. Chen, R. R. Dong, and N. Su, X0(2900) and X1(2900): Hadronic molecules or compact tetraquarks, Chin. Phys. Lett. 37, 101201 (2020)

[122]

X. G. He, W. Wang, and R. L. Zhu, Open-charm tetraquark Xc and open-bottom tetraquark Xb, Eur. Phys. J. C 80, 1026 (2020)

[123]

Q. F. Lu, D. Y. Chen, and Y. B. Dong, Open charm and bottom tetraquarks in an extended relativized quark model, Phys. Rev. D 102, 074021 (2020)

[124]

M. Z. Liu, J. J. Xie, and L. S. Geng, X0(2866) as a D∗K¯∗ molecular state, Phys. Rev. D 102, 091502 (2020)

[125]

M. W. Hu, X. Y. Lao, P. Ling, and Q. Wang, X0(2900) and its heavy quark spin partners in molecular picture, Chin. Phys. C 45, 021003 (2021)

[126]

R. Molina and E. Oset, Molecular picture for the X0(2866) as a D∗K¯∗ JP=0+ state and related 1+,2+ states, Phys. Lett. B 811, 135870 (2020)

[127]

S. S. Agaev, K. Azizi, and H. Sundu, New scalar resonance X0(2900) as a molecule: mass and width, J. Phys. G 48, 085012 (2021)

[128]

R. M. Albuquerque, S. Narison, D. Rabetiarivony, and G. Randriamanatrika, X0,1(2900) and (D−K+) invariant mass from QCD Laplace sum rules at NLO, Nucl. Phys. A 1007, 122113 (2021)

[129]

X. H. Liu, M. J. Yan, H. W. Ke, G. Li, and J. J. Xie, Triangle singularity as the origin of X0(2900) and X1(2900) observed in B+→D+D−K+, Eur. Phys. J. C 80, 1178 (2020)

[130]

R. Aaij, . First observation of a doubly charged tetraquark and its neutral partner, Phys. Rev. Lett. 131, 041902 (2023)

[131]

R. Aaij, . Amplitude analysis of B0→D¯0Ds+π− and B+→D−Ds+π+ decays, Phys. Rev. D 108, 012017 (2023)

[132]

V. M. Abazov, . Evidence for a Bs0π± State, Phys. Rev. Lett. 117, 022003 (2016)

[133]

Z. G. Wang, Analysis of the X(5568) as scalar tetraquark state in the diquark-antidiquark model with QCD sum rules, Commun. Theor. Phys. 66, 335 (2016)

[134]

W. Chen, H. X. Chen, X. Liu, T. G. Steele, and S. L. Zhu, Decoding the X(5568) as a fully open-flavor sub¯d¯ tetraquark state, Phys. Rev. Lett. 117, 022002 (2016)

[135]

S. S. Agaev, K. Azizi, and H. Sundu, Mass and decay constant of the newly observed exotic X(5568) state, Phys. Rev. D 93, 074024 (2016)

[136]

W. Wang and R. L. Zhu, Can X(5568) be a tetraquark state?, Chin. Phys. C 40, 093101 (2016)

[137]

C. M. Zanetti, M. Nielsen, and K. P. Khemchandani, QCD sum rule study of a charged bottom-strange scalar meson, Phys. Rev. D 93, 096011 (2016)

[138]

R. Aaij, . Search for structure in the Bs0π± invariant mass spectrum, Phys. Rev. Lett. 117, 152003 (2016)

[139]

A. M. Sirunyan, . Search for the X(5568) state decaying into Bs0π± in proton−proton collisions at s=8TeV, Phys. Rev. Lett. 120, 202005 (2018)

[140]

M. Aaboud, . Search for a structure in the Bs0π± invariant mass spectrum with the ATLAS experiment, Phys. Rev. Lett. 120, 202007 (2018)

[141]

T. Aaltonen, . A search for the exotic meson X(5568) with the collider detector at Fermilab, Phys. Rev. Lett. 120, 202006 (2018)

[142]

R. Aaij, . Observation of structure in the J/ψ-pair mass spectrum, Sci. Bull. 65, 1983 (2020)

[143]

G. Aad, . Observation of an excess of dicharmonium events in the four-muon final state with the ATLAS detector, Phys. Rev. Lett. 131, 151902 (2023)

[144]

A. Hayrapetyan, . New structures in the J/ψJ/ψ mass spectrum in proton−proton collisions at s=13TeV, Phys. Rev. Lett. 132, 111901 (2024)

[145]

B. Aubert, . Observation of a broad structure in the π+π−J/ψ mass spectrum around 4.26GeV/c2, Phys. Rev. Lett. 95, 142001 (2005)

[146]

C. Z. Yuan, . Measurement of e+e−→π+π−J/ψ cross section via initial state radiation at Belle, Phys. Rev. Lett. 99, 182004 (2007)

[147]

Q. He, . Confirmation of the Y(4260) resonance production in ISR, Phys. Rev. D 74, 091104 (2006)

[148]

X. L. Wang, . Observation of two resonant structures in e+e−→π+π−ψ(2S) via initial state radiation at Belle, Phys. Rev. Lett. 99, 142002 (2007)

[149]

X. L. Wang, . Measurement of e+e−→π+π−ψ(2S) via initial state radiation at Belle, Phys. Rev. D 91, 112007 (2015)

[150]

J. P. Lees, . Study of the reaction e+e−→ψ(2S)π+π− via initial state radiation at BaBar, Phys. Rev. D 89, 111103 (2014)

[151]

G. Pakhlova, . Observation of a near-threshold enhancement in the e+e−→Λc+Λc− cross section using initial-state radiation, Phys. Rev. Lett. 101, 172001 (2008)

[152]

M. Ablikim, . Study of e+e−→ωχcJ at center-of-mass energies from 4.21 to 4.42 GeV, Phys. Rev. Lett. 114, 092003 (2015)

[153]

M. Ablikim, . Evidence of two resonant structures in e+e−→π+π−hc, Phys. Rev. Lett. 118, 092002 (2017)

[154]

M. Ablikim, . Precise measurement of the e+e−→π+π−J/ψ cross section at center-of-mass energies from 3.77 to 4.60 GeV, Phys. Rev. Lett. 118, 092001 (2017)

[155]

M. Ablikim, . Observation of the Y(4230) and a new structure in e+e−→K+K−J/ψ, Chin. Phys. C 46, 111002 (2022)

[156]

M. Ablikim, . Observation of three charmoniumlike states with JPC=1−− in e+e−→D∗0D∗−π+, Phys. Rev. Lett. 130, 121901 (2023)

[157]

M. Ablikim, . Precise measurement of the e+e−→Ds∗+Ds∗− cross sections at center-of-mass energies from threshold to 4.95 GeV, Phys. Rev. Lett. 131, 151903 (2023)

[158]

M. Ablikim, . Observation of a vector charmoniumlike state at 4.7GeV/c2 and search for Zcs in e+e−→K+K−J/ψ, Phys. Rev. Lett. 131, 211902 (2023)

[159]

M. Ablikim, . Observation of structures in the processes e+e−→ωχc1 and ωχc2, Phys. Rev. Lett. 132, 161901 (2024)

[160]

M. Ablikim, . Study of e+e−→ωX(3872) and γX(3872) from 4.66 to 4.95 GeV, Phys. Rev. D 110, 012006 (2024)

[161]

M. Ablikim, . Measurement of energy-dependent pair-production cross section and electromagnetic form factors of a charmed baryon, Phys. Rev. Lett. 131, 191901 (2023)

[162]

M. Ablikim, . Observation of a charged charmoniumlike structure in e+e−→π+π−J/ψ at s=4.26GeV, Phys. Rev. Lett. 110, 252001 (2013)

[163]

Z. Q. Liu, . Study of e+e−→π+π−J/ψ and observation of a charged charmonium-like state at Belle, Phys. Rev. Lett. 110, 252002 (2013)

[164]

T. Xiao, S. Dobbs, A. Tomaradze, and K. K. Seth, Observation of the charged hadron Zc±(3900) and evidence for the neutral Zc0(3900) in e+e−→π+π−J/ψ at s=4170MeV, Phys. Lett. B 727, 366 (2013)

[165]

M. Ablikim, . Observation of a charged charmoniumlike structure in e+e−→(D∗D¯∗)±π∓ at s=4.26GeV, Phys. Rev. Lett. 112, 132001 (2014)

[166]

M. Ablikim, . Observation of a charged charmoniumlike structure Zc(4020) and search for the Zc(3900) in e+e−→π+π−hc, Phys. Rev. Lett. 111, 242001 (2013)

[167]

M. Ablikim, . Observation of a charged (DD¯∗)± mass peak in e+e−→πDD¯∗ at s=4.26GeV, Phys. Rev. Lett. 112, 022001 (2014)

[168]

M. Ablikim, . Determination of spin and parity of the Zc(3900), Phys. Rev. Lett. 119, 072001 (2017)

[169]

M. Ablikim, . Observation of a near-threshold structure in the K+ recoil-mass spectra in e+e−→K+(Ds−D∗0+Ds∗−D0), Phys. Rev. Lett. 126, 102001 (2021)

[170]

M. Ablikim, . Search for hidden-charm tetraquark with strangeness in e+e−→K+Ds∗−D∗0+c.c., Chin. Phys. C 47, 033001 (2023)

[171]

Z. G. Wang, Strange cousin of Zc(4020/4025) as a tetraquark state, Chin. Phys. C 46, 123106 (2022)

[172]

Z. G. Wang, Analysis of Zcs(3985) as the axialvector tetraquark state, Chin. Phys. C 45, 073107 (2021)

[173]

R. Mizuk, . Observation of two resonance-like structures in the π+χc1 mass distribution in exclusive B¯0→K−π+χc1 decays, Phys. Rev. D 78, 072004 (2008)

[174]

J. P. Lees, . Search for the Z1(4050)+ and Z2(4250)+ states in B¯0→χc1K−π+ and B+→χc1KS0π+, Phys. Rev. D 85, 052003 (2012)

[175]

R. Aaij, . Evidence for an ηc(1S)π− resonance in B0→ηc(1S)K+π− decays, Eur. Phys. J. C 78, 1019 (2018)

[176]

S. K. Choi, . Observation of a resonance-like structure in the π±ψ′ mass distribution in exclusive B→Kπ±ψ′ decays, Phys. Rev. Lett. 100, 142001 (2008)

[177]

R. Mizuk, . Dalitz analysis of B→Kπ+ψ′ decays and the Z(4430)+, Phys. Rev. D 80, 031104 (2009)

[178]

K. Chilikin, . Experimental constraints on the spin and parity of the Z(4430)+, Phys. Rev. D 88, 074026 (2013)

[179]

R. Aaij, . Observation of the resonant character of the Z(4430)− state, Phys. Rev. Lett. 112, 222002 (2014)

[180]

S. H. Lee, A. Mihara, F. S. Navarra, and M. Nielsen, QCD sum rules study of the meson Z+(4430), Phys. Lett. B 661, 28 (2008)

[181]

M. Nielsen and F. S. Navarra, Charged Exotic Charmonium States, Mod. Phys. Lett. A 29, 1430005 (2014)

[182]

Z. G. Wang, Analysis of the Z(4430) as the first radial excitation of the Zc(3900), Commun. Theor. Phys. 63, 325 (2015)

[183]

R. Aaij, . Model-independent observation of exotic contributions to B0→J/ψK+π− decays, Phys. Rev. Lett. 122, 152002 (2019)

[184]

Z. G. Wang, Axialvector tetraquark candidates for the Zc(3900), Zc(4020), Zc(4430), Zc(4600), Chin. Phys. C 44, 063105 (2020)

[185]

H. X. Chen and W. Chen, Settling the Zc(4600) in the charged charmonium-like family, Phys. Rev. D 99, 074022 (2019)

[186]

K. Chilikin, . Observation of a new charged charmoniumlike state in B¯0→J/ψK−π+ decays, Phys. Rev. D 90, 112009 (2014)

[187]

I. Adachi et al, ., Observation of two charged bottomonium-like resonances, arXiv:

[188]

A. Bondar, . Observation of two charged bottomonium-like resonances in Υ(5S) decays, Phys. Rev. Lett. 108, 122001 (2012)

[189]

P. Krokovny, . First observation of the Zb0(10610) in a dalitz analysis of Υ(5S)→Υ(nS)π0π0, Phys. Rev. D 88, 052016 (2013)

[190]

A. Abdesselam, . Observation of a new structure near 10.75 GeV in the energy dependence of the e+e−→Υ(nS)π+π−(n=1,2,3) cross sections, JHEP 10, 220 (2019)

[191]

I. Adachi, . Observation of e+e−→ωχbJ(1P) and search for Xb→ωΥ(1S) at s near 10.75 GeV, Phys. Rev. Lett. 130, 091902 (2023)

[192]

I. Adachi, . Study of Y(10753) decays to π+π−Υ(nS) final states at Belle II, JHEP 07, 116 (2024)

[193]

I. Adachi, . Search for the e+e−→ηb(1S)ω and e+e−→χb0(1P)ω processes at s=10.745GeV, Phys. Rev. D 109, 072013 (2024)

[194]

R. Aaij, . Observation of an exotic narrow doubly charmed tetraquark, Nature Phys. 18, 751 (2022)

[195]

R. Aaij, . Study of the doubly charmed tetraquark Tcc+, Nature Commun. 13, 3351 (2022)

[196]

R. Aaij, . Observation of J/ψp resonances consistent with pentaquark states in Λb0→J/ψK−p decays, Phys. Rev. Lett. 115, 072001 (2015)

[197]

R. Aaij, . Observation of a narrow pentaquark state, Pc(4312)+, and of two-peak structure of the Pc(4450)+, Phys. Rev. Lett. 122, 222001 (2019)

[198]

R. Aaij, . Evidence of a J/ψΛ structure and observation of excited Ξ− states in the Ξb−→J/ψΛK− decay, Sci. Bull. 66, 1278 (2021)

[199]

R. Aaij, . Evidence for a new structure in the J/ψp and J/ψp¯ systems in B¯s0→J/ψpp¯ decays, Phys. Rev. Lett. 128, 062001 (2022)

[200]

R. Aaij, . Observation of a J/ψΛ resonance consistent with a strange pentaquark candidate in B−→J/ψΛp¯ decays, Phys. Rev. Lett. 131, 031901 (2023)

[201]

P. Adlarson, . Evidence for a new resonance from polarized neutron−proton scattering, Phys. Rev. Lett. 112, 202301 (2014)

[202]

P. Adlarson, . Neutron−proton scattering in the context of the d∗(2380) resonance, Phys. Rev. C 90, 035204 (2014)

[203]

F. Huang, Z. Y. Zhang, P. N. Shen, and W. L. Wang, Is d∗ a candidate for a hexaquark-dominated exotic state?, Chin. Phys. C 39, 071001 (2015)

[204]

W. Park, A. Park, and S. H. Lee, Dibaryons in a constituent quark model, Phys. Rev. D 92, 014037 (2015)

[205]

H. X. Chen, E. L. Cui, W. Chen, T. G. Steele, and S. L. Zhu, QCD sum rule study of the d∗(2380), Phys. Rev. C 91, 025204 (2015)

[206]

A. Gal, The d∗(2380) dibaryon resonance width and decay branching ratios, Phys. Lett. B 769, 436 (2017)

[207]

F. Huang, Progress on d∗(2380) in a chiral SU(3) quark model, Rev. Mex. Fis. Suppl. 3, 0308031 (2022)

[208]

N. A. Tornqvist, From the deuteron to deusons, an analysis of deuteron-like meson meson bound states, Z. Phys. C 61, 525 (1994)

[209]

M. Karliner and J. L. Rosner, New exotic meson and baryon resonances from doubly-heavy hadronic molecules, Phys. Rev. Lett. 115, 122001 (2015)

[210]

C. E. Thomas and F. E. Close, Is X(3872) a molecule?, Phys. Rev. D 78, 034007 (2008)

[211]

C. Hidalgo-Duque, J. Nieves, and M. P. Valderrama, Light flavor and heavy quark spin symmetry in heavy meson molecules, Phys. Rev. D 87, 076006 (2013)

[212]

Y. R. Liu and M. Oka, ΛcN bound states revisited, Phys. Rev. D 85, 014015 (2012)

[213]

M. Cleven, F. K. Guo, C. Hanhart, Q. Wang, and Q. Zhao, Employing spin symmetry to disentangle different models for the XYZ states, Phys. Rev. D 92, 014005 (2015)

[214]

Y. Yamaguchi, A. Giachino, A. Hosaka, E. Santopinto, S. Takeuchi, and M. Takizawa, Hidden-charm and bottom meson−baryon molecules coupled with five-quark states, Phys. Rev. D 96, 114031 (2017)

[215]

F. Close, C. Downum, and C. E. Thomas, Novel charmonium and bottomonium spectroscopies due to deeply bound Hadronic molecules from single pion exchange, Phys. Rev. D 81, 074033 (2010)

[216]

F. K. Guo, X. H. Liu, and S. Sakai, Threshold cusps and triangle singularities in hadronic reactions, Prog. Part. Nucl. Phys. 112, 103757 (2020)

[217]

D. V. Bugg, How resonances can synchronise with thresholds, J. Phys. G 35, 075005 (2008)

[218]

E. S. Swanson, Zb and Zc exotic states as coupled channel cusps, Phys. Rev. D 91, 034009 (2015)

[219]

E. S. Swanson, Cusps and exotic charmonia, Int. J. Mod. Phys. E 25, 1642010 (2016)

[220]

D. Y. Chen and X. Liu, Zb(10610) and Zb(10650) structures produced by the initial single pion emission in the Υ(5S) decays, Phys. Rev. D 84, 094003 (2011)

[221]

D. Y. Chen, X. Liu, and T. Matsuki, Reproducing the Zc(3900) structure through the initial-single-pion-emission mechanism, Phys. Rev. D 88, 036008 (2013)

[222]

D. Y. Chen, X. Liu, and T. Matsuki, Predictions of charged charmoniumlike structures with hidden-charm and Open-Strange channels, Phys. Rev. Lett. 110, 232001 (2013)

[223]

X. K. Dong, F. K. Guo, and B. S. Zou, Explaining the many threshold structures in the heavy-quark hadron spectrum, Phys. Rev. Lett. 126, 152001 (2021)

[224]

X. H. Liu and G. Li, Exploring the threshold behavior and implications on the nature of Y(4260) and Zc(3900), Phys. Rev. D 88, 014013 (2013)

[225]

X. H. Liu, Q. Wang, and Q. Zhao, Understanding the newly observed heavy pentaquark candidates, Phys. Lett. B 757, 231 (2016)

[226]

F. K. Guo, U. G. Meissner, J. Nieves, and Z. Yang, Remarks on the Pc structures and triangle singularities, Eur. Phys. J. A 52, 318 (2016)

[227]

Q. Wang, C. Hanhart, and Q. Zhao, Decoding the riddle of Y(4260) and Zc(3900), Phys. Rev. Lett. 111, 132003 (2013)

[228]

Q. Wang, C. Hanhart, and Q. Zhao, Systematic study of the singularity mechanism in heavy quarkonium decays, Phys. Lett. B 725, 106 (2013)

[229]

M. Cleven, Q. Wang, F. K. Guo, C. Hanhart, U. G. Meissner, and Q. Zhao, Y(4260) as the first S-wave open charm vector molecular state?, Phys. Rev. D 90, 074039 (2014)

[230]

M. L. Du, M. Albaladejo, F. K. Guo, and J. Nieves, A combined analysis of the Zc(3900) and the Zcs(3985) exotic states, Phys. Rev. D 105, 074018 (2022)

[231]

L. W. Yan, Z. H. Guo, F. K. Guo, D. L. Yao, and Z. Y. Zhou, Reconciling experimental and lattice data of Zc(3900) in a J/ψπ−DD¯∗ coupled-channel analysis, Phys. Rev. D 109, 014026 (2024)

[232]

Q. Wu, D. Y. Chen, and F. K. Guo, Production of the Zb(′) states from the Υ(5S,6S) decays, Phys. Rev. D 99, 034022 (2019)

[233]

Y. H. Chen, M. Cleven, J. T. Daub, F. K. Guo, C. Hanhart, B. Kubis, U. G. Meissner, and B. S. Zou, Effects of Zb states and bottom meson loops on Υ(4S)→Υ(1S,2S)π+π− transitions, Phys. Rev. D 95, 034022 (2017)

[234]

M. Albaladejo, F. K. Guo, C. Hidalgo-Duque, and J. Nieves, Zc(3900): What has been really seen?, Phys. Lett. B 755, 337 (2016)

[235]

F. K. Guo, C. Hanhart, Q. Wang, and Q. Zhao, Could the near-threshold XYZ states be simply kinematic effects?, Phys. Rev. D 91, 051504 (2015)

[236]

A. P. Szczepaniak, Triangle singularities and XYZ quarkonium peaks, Phys. Lett. B 747, 410 (2015)

[237]

Y. H. Chen, M. L. Du, and F. K. Guo, Precise determination of the pole position of the exotic Zc(3900), Sci. China Phys. Mech. Astron. 67, 291011 (2024)

[238]

B. A. Li, Is X(3872) a possible candidate of hybrid meson?, Phys. Lett. B 605, 306 (2005)

[239]

X. H. Liu, Q. Zhao, and F. E. Close, Search for tetraquark candidate Z(4430) in meson photoproduction, Phys. Rev. D 77, 094005 (2008)

[240]

X. Y. Wang, J. He, X. R. Chen, Q. J. Wang, and X. M. Zhu, Pion-induced production of hidden-charm pentaquarks Pc(4312), Pc(4440) and Pc(4457), Phys. Lett. B 797, 134862 (2019)

[241]

J. Chen, F. K. Guo, Y. G. Ma, C. P. Shen, Q. Shou, Q. Wang, J. J. Wu, and B. S. Zou, Production of exotic hadrons in pp and nuclear collisions, Nucl. Sci. Tech. 36, 55 (2025)

[242]

R. Casalbuoni, A. Deandrea, N. Di Bartolomeo, F. Feruglio, R. Gatto, and G. Nardulli, Phenomenology of heavy meson chiral Lagrangians, Phys. Rep. 281, 145 (1997)

[243]

Z. G. Wang, Strong decays of the bottom mesons B1(5721), B2(5747), Bs1(5830), Bs2(5840) and B(5970), Eur. Phys. J. Plus 129, 186 (2014)

[244]

S. Campanella, P. Colangelo, and F. De Fazio, Excited heavy meson decays to light vector mesons: Implications for spectroscopy, Phys. Rev. D 98, 114028 (2018)

[245]

J. A. Oller, E. Oset, and A. Ramos, Chiral unitary approach to meson−meson and meson−baryon interactions and nuclear applications, Prog. Part. Nucl. Phys. 45, 157 (2000)

[246]

T. Hyodo, D. Jido, and A. Hosaka, Compositeness of dynamically generated states in a chiral unitary approach, Phys. Rev. C 85, 015201 (2012)

[247]

D. Gamermann and E. Oset, Axial resonances in the open and hidden charm sectors, Eur. Phys. J. A 33, 119 (2007)

[248]

R. Molina and E. Oset, The Y(3940), Z(3930) and the X(4160) as dynamically generated resonances from the vector-vector interaction, Phys. Rev. D 80, 114013 (2009)

[249]

F. Aceti, M. Bayar, E. Oset, A. Martinez Torres, K. P. Khemchandani, J. M. Dias, F. S. Navarra, and M. Nielsen, Prediction of an I=1 DD¯∗ state and relationship to the claimed Zc(3900), Zc(3885), Phys. Rev. D 90, 016003 (2014)

[250]

T. Ji, X. K. Dong, M. Albaladejo, M. L. Du, F. K. Guo, and J. Nieves, Establishing the heavy quark spin and light flavor molecular multiplets of the X(3872), Zc(3900) and X(3960), Phys. Rev. D 106, 094002 (2022)

[251]

J. M. Dias, F. Aceti, and E. Oset, Study of BB¯∗ and B∗B¯∗ interactions in I=1 and relationship to the Zb(10610), Zb(10650) states, Phys. Rev. D 91, 076001 (2015)

[252]

M. Cleven, Q. Wang, F. K. Guo, C. Hanhart, U. G. Meissner, and Q. Zhao, Confirming the molecular nature of the Zb(10610) and the Zb(10650), Phys. Rev. D 87, 074006 (2013)

[253]

M. Cleven, F. K. Guo, C. Hanhart, and U. G. Meissner, Bound state nature of the exotic Zb states, Eur. Phys. J. A 47, 120 (2011)

[254]

F. K. Guo, P. N. Shen, H. C. Chiang, and R. G. Ping, Heavy quarkonium π+π− transitions and a possible bb¯qq¯ state, Nucl. Phys. A 761, 269 (2005)

[255]

J. J. Wu, R. Molina, E. Oset, and B. S. Zou, Dynamically generated N∗ and Λ∗ resonances in the hidden charm sector around 4.3 GeV, Phys. Rev. C 84, 015202 (2011)

[256]

C. W. Xiao, J. Nieves, and E. Oset, Combining heavy quark spin and local hidden gauge symmetries in the dynamical generation of hidden charm baryons, Phys. Rev. D 88, 056012 (2013)

[257]

L. Roca, J. Nieves, and E. Oset, The LHCb pentaquark as a D¯∗Σc−D¯∗Σc∗ molecular state, Phys. Rev. D 92, 094003 (2015)

[258]

T. Uchino, W. H. Liang, and E. Oset, Baryon states with hidden charm in the extended local hidden gauge approach, Eur. Phys. J. A 52, 43 (2016)

[259]

C. W. Xiao, J. Nieves, and E. Oset, Heavy quark spin symmetric molecular states from D¯(∗)Σc(∗) and other coupled channels in the light of the recent LHCb pentaquarks, Phys. Rev. D 100, 014021 (2019)

[260]

T. Hyodo, Structure of near-threshold s-wave resonances, Phys. Rev. Lett. 111, 132002 (2013)

[261]

Z. H. Guo and J. A. Oller, Probabilistic interpretation of compositeness relation for resonances, Phys. Rev. D 93, 096001 (2016)

[262]

Y. Kamiya and T. Hyodo, Structure of near-threshold quasibound states, Phys. Rev. C 93, 035203 (2016)

[263]

X. W. Kang, Z. H. Guo, and J. A. Oller, General considerations on the nature of Zb(10610) and Zb(10650) from their pole positions, Phys. Rev. D 94, 014012 (2016)

[264]

Y. Li, F. K. Guo, J. Y. Pang, and J. J. Wu, Generalization of Weinberg’s compositeness relations, Phys. Rev. D 105, L071502 (2022)

[265]

Z. H. Zhang, T. Ji, X. K. Dong, F. K. Guo, C. Hanhart, U. G. Meissner, and A. Rusetsky, Predicting isovector charmonium-like states from X(3872) properties, JHEP 08, 130 (2024)

[266]

T. Ji,X. K. Dong,F. K. Guo,C. Hanhart,U. G. Meissner, Precise determination of the properties of X(3872) and of its isovector partner Wc1, arXiv:

[267]

J. He, The Zc(3900) as a resonance from the DD¯∗ interaction, Phys. Rev. D 92, 034004 (2015)

[268]

J. He, D¯Σc∗ and D¯∗Σc interactions and the LHCb hidden-charmed pentaquarks, Phys. Lett. B 753, 547 (2016)

[269]

J. He, Study of Pc(4457), Pc(4440), and Pc(4312) in a quasipotential Bethe-Salpeter equation approach, Eur. Phys. J. C 79, 393 (2019)

[270]

P. G. Ortega, J. Segovia, D. R. Entem, and F. Fernandez, The Zc structures in a coupled-channels model, Eur. Phys. J. C 79, 78 (2019)

[271]

H. W. Ke, X. Q. Li, Y. L. Shi, G. L. Wang, and X. H. Yuan, Is Zb(10610) a molecular state?, JHEP 04, 056 (2012)

[272]

H. W. Ke, M. Li, X. H. Liu, and X. Q. Li, Study on possible molecular states composed of ΛcD¯(ΛbB) and ΣcD¯(ΣbB) within the Bethe-Salpeter framework, Phys. Rev. D 101, 014024 (2020)

[273]

M. J. Zhao, Z. Y. Wang, C. Wang, and X. H. Guo, Investigation of the possible DD¯∗/BB¯∗ and DD∗/B¯B¯∗ molecule states, Phys. Rev. D 105, 096016 (2022)

[274]

J. Nieves and M. P. Valderrama, Deriving the existence of BB¯∗ bound states from the X(3872) and heavy quark symmetry, Phys. Rev. D 84, 056015 (2011)

[275]

M. Z. Liu, T. W. Wu, J. J. Xie, M. P. Valderrama, and L. S. Geng, and D∗Ξ molecular states from one boson exchange, Phys. Rev. D 98, 014014 (2018)

[276]

M. Z. Liu, T. W. Wu, M. P. Valderrama, J. J. Xie, and L. S. Geng, Heavy-quark spin and flavor symmetry partners of the X(3872) revisited: What can we learn from the one boson exchange model?, Phys. Rev. D 99, 094018 (2019)

[277]

M. Z. Liu, Y. W. Pan, F. Z. Peng, M. S. Sanchez, L. S. Geng, A. Hosaka, and M. Pa. Valderrama, Emergence of a complete heavy-quark spin symmetry multiplet: seven molecular pentaquarks in light of the latest LHCb analysis, Phys. Rev. Lett. 122, 242001 (2019)

[278]

M. Z. Liu, Y. W. Pan, and L. S. Geng, Can discovery of hidden charm strange pentaquark states help determine the spins of Pc(4440) and Pc(4457)?, Phys. Rev. D 103, 034003 (2021)

[279]

X. Liu, Z. G. Luo, Y. R. Liu, and S. L. Zhu, X(3872) and other possible heavy molecular states, Eur. Phys. J. C 61, 411 (2009)

[280]

G. J. Ding, Are Y(4260) and Z2+(4250) D1D or D0D∗ hadronic molecules?, Phys. Rev. D 79, 014001 (2009)

[281]

I. W. Lee, A. Faessler, T. Gutsche, and V. E. Lyubovitskij, X(3872) as a molecular DD∗ state in a potential model, Phys. Rev. D 80, 094005 (2009)

[282]

Z. F. Sun, J. He, X. Liu, Z. G. Luo, and S. L. Zhu, Zb(10610)± and Zb(10650)± as the B∗B¯ and B∗B¯∗ molecular states, Phys. Rev. D 84, 054002 (2011)

[283]

Z. C. Yang, Z. F. Sun, J. He, X. Liu, and S. L. Zhu, The possible hidden-charm molecular baryons composed of anti-charmed meson and charmed baryon, Chin. Phys. C 36, 6 (2012)

[284]

S. Ohkoda, Y. Yamaguchi, S. Yasui, K. Sudoh, and A. Hosaka, Exotic mesons with double charm and bottom flavor, Phys. Rev. D 86, 034019 (2012)

[285]

N. Li, Z. F. Sun, X. Liu, and S. L. Zhu, Coupled-channel analysis of the possible D(∗)D(∗), B¯(∗)B¯(∗) and D(∗)B¯(∗) molecular states, Phys. Rev. D 88, 114008 (2013)

[286]

H. Xu, B. Wang, Z. W. Liu, and X. Liu, DD∗ potentials in chiral perturbation theory and possible molecular states, Phys. Rev. D 99, 014027 (2019)

[287]

L. Meng, B. Wang, G. J. Wang, and S. L. Zhu, The hidden charm pentaquark states and ΣcD¯(∗) interaction in chiral perturbation theory, Phys. Rev. D 100, 014031 (2019)

[288]

B. Wang, L. Meng, and S. L. Zhu, Hidden-charm and hidden-bottom molecular pentaquarks in chiral effective field theory, JHEP 11, 108 (2019)

[289]

M. Z. Liu, T. W. Wu, M. S. Sanchez, M. P. Valderrama, L. S. Geng, and J. J. Xie, Spin-parities of the Pc(4440) and Pc(4457) in the one-boson-exchange model, Phys. Rev. D 103, 054004 (2021)

[290]

S. Y. Yu and X. W. Kang, Nature of X(3872) from its radiative decay, Phys. Lett. B 848, 138404 (2024)

[291]

E. Braaten and M. Kusunoki, Low-energy universality and the new charmonium resonance at 3870-MeV, Phys. Rev. D 69, 074005 (2004)

[292]

P. G. Ortega, J. Segovia, D. R. Entem, and F. Fernandez, Coupled channel approach to the structure of the X(3872), Phys. Rev. D 81, 054023 (2010)

[293]

S. Coito, G. Rupp, and E. van Beveren, X(3872) is not a true molecule, Eur. Phys. J. C 73, 2351 (2013)

[294]

P. G. Ortega, J. Segovia, D. R. Entem, and F. Fernandez, Charmonium resonances in the 3.9 GeV/c2 energy region and the X(3915)/X(3930) puzzle, Phys. Lett. B 778, 5 (2018)

[295]

Z. Y. Zhou and Z. Xiao, Understanding X(3862), X(3872), and X(3930) in a Friedrichs-model-like scheme, Phys. Rev. D 96, 054031 (2017)

[296]

Z. Y. Zhou and Z. Xiao, Comprehending Isospin breaking effects of X(3872) in a Friedrichs-model-like scheme, Phys. Rev. D 97, 034011 (2018)

[297]

J. Ferretti and E. Santopinto, Threshold corrections of χc(2P) and χb(3P) states and J/ψρ and J/ψω transitions of the X(3872) in a coupled-channel model, Phys. Lett. B 789, 550 (2019)

[298]

Y. S. Kalashnikova and A. V. Nefediev, X(3872) in the molecular model, Phys. Usp. 62, 568 (2019)

[299]

Y. Yamaguchi, A. Hosaka, S. Takeuchi, and M. Takizawa, Heavy hadronic molecules with pion exchange and quark core couplings: A guide for practitioners, J. Phys. G 47, 053001 (2020)

[300]

Q. Deng, R. H. Ni, Q. Li, and X. H. Zhong, Charmonia in an unquenched quark model, Phys. Rev. D 110, 056034 (2024)

[301]

E. Braaten, How the Zc(3900) Reveals the spectra of quarkonium hybrid and tetraquark mesons, Phys. Rev. Lett. 111, 162003 (2013)

[302]

E. Braaten, C. Langmack, and D. H. Smith, Selection rules for hadronic transitions of XYZ mesons, Phys. Rev. Lett. 112, 222001 (2014)

[303]

E. Braaten, C. Langmack, and D. H. Smith, Born-oppenheimer approximation for the XYZ mesons, Phys. Rev. D 90, 014044 (2014)

[304]

M. Berwein, N. Brambilla, J. T. Castella, and A. Vairo, Quarkonium hybrids with nonrelativistic effective field theories, Phys. Rev. D 92, 114019 (2015)

[305]

R. Oncala and J. Soto, Heavy quarkonium hybrids: spectrum, decay and mixing, Phys. Rev. D 96, 014004 (2017)

[306]

N. Brambilla, W. K. Lai, J. Segovia, J. T. Castella, and A. Vairo, Spin structure of heavy-quark hybrids, Phys. Rev. D 99, 014017 (2019)

[307]

K. J. Juge, J. Kuti, and C. J. Morningstar, Ab initio study of hybrid b¯gb mesons, Phys. Rev. Lett. 82, 4400 (1999)

[308]

G. S. Bali, B. Bolder, N. Eicker, T. Lippert, B. Orth, P. Ueberholz, K. Schilling, and T. Struckmann, Static potentials and glueball masses from QCD simulations with Wilson sea quarks, Phys. Rev. D 62, 054503 (2000)

[309]

G. S. Bali, H. Neff, T. Duessel, T. Lippert, and K. Schilling, Observation of string breaking in QCD, Phys. Rev. D 71, 114513 (2005)

[310]

M. Foster and C. Michael, Hadrons with a heavy colour-adjoint particle, Phys. Rev. D 59, 094509 (1999)

[311]

R. L. Jaffe, Exotica, Phys. Rep. 409, 1 (2005)

[312]

R. Faccini, L. Maiani, F. Piccinini, A. Pilloni, A. D. Polosa, and V. Riquer, A JPG=1++ charged resonance in the Y(4260) to π+π−J/ψ decay?, Phys. Rev. D 87, 111102 (2013)

[313]

L. Maiani, V. Riquer, F. Piccinini, and A. D. Polosa, Four quark interpretation of Y(4260), Phys. Rev. D 72, 031502 (2005)

[314]

N. V. Drenska, R. Faccini, and A. D. Polosa, Exotic hadrons with hidden charm and strangeness, Phys. Rev. D 79, 077502 (2009)

[315]

R. F. Lebed and A. D. Polosa, χc0(3915) as the lightest cc¯ss¯ state, Phys. Rev. D 93, 094024 (2016)

[316]

L. Maiani, A. D. Polosa, and V. Riquer, Interpretation of axial resonances in J/ψϕ at LHCb, Phys. Rev. D 94, 054026 (2016)

[317]

L. Maiani, A. D. Polosa, and V. Riquer, A theory of X and Z multiquark resonances, Phys. Lett. B 778, 247 (2018)

[318]

A. Ali, L. Maiani, A. V. Borisov, I. Ahmed, M. Jamil Aslam, A. Y. Parkhomenko, A. D. Polosa, and A. Rehma, A new look at the Y tetraquarks and Ωc baryons in the diquark model, Eur. Phys. J. C 78, 29 (2018)

[319]

L. Maiani, A. D. Polosa, and V. Riquer, The new resonances Zcs(3985) and Zcs(4003) (almost) fill two tetraquark nonets of broken SU(3)f, Sci. Bull. 66, 1616 (2021)

[320]

H. Hogaasen, J. M. Richard, and P. Sorba, A Chromomagnetic mechanism for the X(3872) resonance, Phys. Rev. D 73, 054013 (2006)

[321]

F. Buccella, H. Hogaasen, J. M. Richard, and P. Sorba, Chromomagnetism, flavour symmetry breaking and S-wave tetraquarks, Eur. Phys. J. C 49, 743 (2007)

[322]

L. Zhao, W. Z. Deng, and S. L. Zhu, Hidden-charm tetraquarks and charged Zc states, Phys. Rev. D 90, 094031 (2014)

[323]

J. Wu, Y. R. Liu, K. Chen, X. Liu, and S. L. Zhu, Heavy-flavored tetraquark states with the QQQ¯Q¯ configuration, Phys. Rev. D 97, 094015 (2018)

[324]

T. Guo, J. Li, J. Zhao, and L. He, Mass spectra of doubly heavy tetraquarks in an improved chromomagnetic interaction model, Phys. Rev. D 105, 014021 (2022)

[325]

R. F. Lebed, Spectroscopy of exotic hadrons formed from dynamical diquarks, Phys. Rev. D 96, 116003 (2017)

[326]

J. F. Giron, R. F. Lebed, and C. T. Peterson, The dynamical diquark model: First numerical results, JHEP 05, 061 (2019)

[327]

J. F. Giron, R. F. Lebed, and C. T. Peterson, The dynamical diquark model: Fine structure and isospin, JHEP 01, 124 (2020)

[328]

J. F. Giron and R. F. Lebed, Spectrum of p-wave hidden-charm exotic mesons in the diquark model, Phys. Rev. D 101, 074032 (2020)

[329]

J. F. Giron and R. F. Lebed, Spectrum of the hidden-bottom and the hidden-charm-strange exotics in the dynamical diquark model, Phys. Rev. D 102, 014036 (2020)

[330]

J. F. Giron, R. F. Lebed, and S. R. Martinez, Spectrum of hidden-charm, open-strange exotics in the dynamical diquark model, Phys. Rev. D 104, 054001 (2021)

[331]

R. F. Lebed and S. R. Martinez, Tcc in the diabatic diquark model: Effects of D∗D isospin, Phys. Rev. D 110, 034033 (2024)

[332]

Q. F. Lu and Y. B. Dong, X(4140), X(4274), X(4500), and X(4700) in the relativized quark model, Phys. Rev. D 94, 074007 (2016)

[333]

M. N. Anwar, J. Ferretti, and E. Santopinto, Spectroscopy of the hidden-charm [qc][q¯c¯] and [sc][s¯c¯] tetraquarks in the relativized diquark model, Phys. Rev. D 98, 094015 (2018)

[334]

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)

[335]

J. Ferretti and E. Santopinto, Hidden-charm and bottom tetra- and pentaquarks with strangeness in the hadro-quarkonium and compact tetraquark models, JHEP 04, 119 (2020)

[336]

Q. F. Lu, D. Y. Chen, and Y. B. Dong, Masses of doubly heavy tetraquarks TQQ′ in a relativized quark model, Phys. Rev. D 102, 034012 (2020)

[337]

Q. F. Lu, D. Y. Chen, and Y. B. Dong, Masses of fully heavy tetraquarks QQQ¯Q¯ in an extended relativized quark model, Eur. Phys. J. C 80, 871 (2020)

[338]

G. L. Yu, Z. Y. Li, Z. G. Wang, J. Lu, and M. Yan, The S- and P-wave fully charmed tetraquark states and their radial excitations, Eur. Phys. J. C 83, 416 (2023)

[339]

W. C. Dong and Z. G. Wang, Going in quest of potential tetraquark interpretations for the newly observed Tψψ states in light of the diquark−antidiquark scenarios, Phys. Rev. D 107, 074010 (2023)

[340]

G. L. Yu, Z. Y. Li, Z. G. Wang, B. Wu, and Z. Zhou, The ground states of hidden-charm tetraquarks and their radial excitations, Eur. Phys. J. C 84, 1130 (2024)

[341]

W. C. Dong and Z. G. Wang, Hunting for the prospective Tcc family based on the diquark-antidiquark configuration, Nucl. Phys. B 1012, 116828 (2025)

[342]

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)

[343]

D. Ebert, R. N. Faustov, and V. O. Galkin, Excited heavy tetraquarks with hidden charm, Eur. Phys. J. C 58, 399 (2008)

[344]

R. N. Faustov, V. O. Galkin, and E. M. Savchenko, Masses of the QQQ¯Q¯ tetraquarks in the relativistic diquark-antidiquark picture, Phys. Rev. D 102, 114030 (2020)

[345]

J. Vijande, F. Fernandez, A. Valcarce, and B. Silvestre-Brac, Tetraquarks in a chiral constituent quark model, Eur. Phys. J. A 19, 383 (2004)

[346]

N. Barnea, J. Vijande, and A. Valcarce, Four-quark spectroscopy within the hyperspherical formalism, Phys. Rev. D 73, 054004 (2006)

[347]

J. Vijande, E. Weissma, N. Barnea, and A. Valcarce, Do cc¯nn¯ bound states exist?, Phys. Rev. D 76, 094022 (2007)

[348]

Y. R. Liu and Z. Y. Zhang, X(3872) and the bound state problem of D0D¯∗0 (D¯0D∗0) in a chiral quark model, Phys. Rev. C 79, 035206 (2009)

[349]

G. J. Wang, L. Meng, and S. L. Zhu, Spectrum of the fully-heavy tetraquark state QQQ¯′Q¯′, Phys. Rev. D 100, 096013 (2019)

[350]

G. Yang, J. L. Ping, and J. Segovia, Tetra- and penta-quark structures in the constituent quark model, Symmetry 12, 1869 (2020)

[351]

S. Noh, W. Park, and S. H. Lee, The doubly-heavy tetraquarks (qq′Q¯Q¯′) in a constituent quark model with a complete set of harmonic oscillator bases, Phys. Rev. D 103, 114009 (2021)

[352]

Z. Zhao, K. Xu, A. Kaewsnod, X. Liu, A. Limphirat, and Y. Yan, Study of charmoniumlike and fully-charm tetraquark spectroscopy, Phys. Rev. D 103, 116027 (2021)

[353]

C. R. Deng, J. L. Ping, H. X. Huang, and F. Wang, Hidden charmed states and multibody color flux-tube dynamics, Phys. Rev. D 98, 014026 (2018)

[354]

C. R. Deng, H. Chen, and J. L. Ping, Systematical investigation on the stability of doubly heavy tetraquark states, Eur. Phys. J. A 56, 9 (2020)

[355]

C. Michael, Adjoint sources in lattice gauge theory, Nucl. Phys. B 259, 58 (1985)

[356]

B. Blossier, M. D. Morte, G. v. Hippel, T. Mendes, and R. Sommer, On the generalized eigenvalue method for energies and matrix elements in lattice field theory, JHEP 04, 094 (2009)

[357]

S. Prelovsek, C. B. Lang, L. Leskovec, and D. Mohler, Study of the Zc+ channel using lattice QCD, Phys. Rev. D 91, 014504 (2015)

[358]

M. Padmanath, C. B. Lang, and S. Prelovsek, X(3872) and Y(4140) using diquark−antidiquark operators with lattice QCD, Phys. Rev. D 92, 034501 (2015)

[359]

A. Francis, R. J. Hudspith, R. Lewis, and K. Maltman, Lattice prediction for deeply bound doubly heavy tetraquarks, Phys. Rev. Lett. 118, 142001 (2017)

[360]

G. K. C. Cheung, C. E. Thomas, J. J. Dudek, and R. G. Edwards, Tetraquark operators in lattice QCD and exotic flavour states in the charm sector, JHEP 11, 033 (2017)

[361]

P. Junnarkar, N. Mathur, and M. Padmanath, Study of doubly heavy tetraquarks in Lattice QCD, Phys. Rev. D 99, 034507 (2019)

[362]

L. Leskovec, S. Meinel, M. Pflaumer, and M. Wagner, Lattice QCD investigation of a doubly-bottom b¯b¯ud tetraquark with quantum numbers I(JP)=0(1+), Phys. Rev. D 100, 014503 (2019)

[363]

A. Francis, R. J. Hudspith, R. Lewis, and K. Maltman, Evidence for charm-bottom tetraquarks and the mass dependence of heavy-light tetraquark states from lattice QCD, Phys. Rev. D 99, 054505 (2019)

[364]

A. Francis, Lattice perspectives on doubly heavy tetraquarks, Prog. Part. Nucl. Phys. 140, 104143 (2025)

[365]

P. Bicudo, Tetraquarks and pentaquarks in lattice QCD with light and heavy quarks, Phys. Rep. 1039, 1 (2023)

[366]

M. Luscher, Two particle states on a torus and their relation to the scattering matrix, Nucl. Phys. B 354, 531 (1991)

[367]

M. Luscher, Signatures of unstable particles in finite volume, Nucl. Phys. B 364, 237 (1991)

[368]

R. A. Briceno, J. J. Dudek, and R. D. Young, Scattering processes and resonances from lattice QCD, Rev. Mod. Phys. 90, 025001 (2018)

[369]

S. Prelovsek and L. Leskovec, Evidence for X(3872) from DD∗ scattering on the lattice, Phys. Rev. Lett. 111, 192001 (2013)

[370]

Y. Chen, . Low-energy scattering of (DD¯∗)± system and the resonance-like structure Zc(3900), Phys. Rev. D 89, 094506 (2014)

[371]

M. Padmanath and S. Prelovsek, Signature of a doubly charm tetraquark pole in DD∗ scattering on the lattice, Phys. Rev. Lett. 129, 032002 (2022)

[372]

S. Chen, C. Shi, Y. Chen, M. Gong, Z. Liu, W. Sun, and R. Zhang, Tcc+(3875) relevant DD∗ scattering from Nf=2 lattice QCD, Phys. Lett. B 833, 137391 (2022)

[373]

M. L. Du, A. Filin, V. Baru, X. K. Dong, E. Epelbaum, F. K. Guo, C. Hanhart, A. Nefediev, J. Nieves, and Q. Wang, Role of left-hand cut contributions on pole extractions from lattice data: Case study for Tcc(3875)+, Phys. Rev. Lett. 131, 131903 (2023)

[374]

S. Aoki, T. Doi, T. Hatsuda, Y. Ikeda, T. Inoue, N. Ishii, K. Murano, H. Nemura, and K. Sasaki, Lattice QCD approach to nuclear physics, PTEP 2012, 01A105 (2012)

[375]

Y. Lyu, S. Aoki, T. Doi, T. Hatsuda, Y. Ikeda, and J. Meng, Doubly charmed tetraquark Tcc+ from lattice QCD near physical point, Phys. Rev. Lett. 131, 161901 (2023)

[376]

M. A. Shifman, A. I. Vainshtein, and V. I. Zakharov, QCD and Resonance Physics. Theoretical Foundations, Nucl. Phys. B 147, 385 (1979)

[377]

M. A. Shifman, A. I. Vainshtein, and V. I. Zakharov, QCD and resonance physics: Applications, Nucl. Phys. B 147, 448 (1979)

[378]

B. L. Ioffe, Calculation of baryon masses in quantum Chromodynamics, Nucl. Phys. B 188, 317 (1981)

[379]

V. L. Chernyak and I. R. Zhitnitsky, Nucleon wave function and nucleon form-factors in QCD, Nucl. Phys. B 246, 52 (1984)

[380]

V. L. Chernyak and A. R. Zhitnitsky, Asymptotic Behavior of Exclusive Processes in QCD, Phys. Rep. 112, 173 (1984)

[381]

L. J. Reinders, H. Rubinstein, and S. Yazaki, Hadron properties from QCD sum rules, Phys. Rep. 127, 1 (1985)

[382]

I. I. Balitsky, V. M. Braun, and A. V. Kolesnichenko, Radiative Decay Σ+→pγ in Quantum Chromodynamics, Nucl. Phys. B 312, 509 (1989)

[383]

V. M. Braun and I. E. Filyanov, QCD sum rules in exclusive kinematics and pion wave function, Z. Phys. C 44, 157 (1989)

[384]

M. A. Shifman, Vacuum Structure and QCD Sum Rules, North-Holland, 1992

[385]

P. Colangelo,A. Khodjamirian, QCD sum rules, a modern perspective, arXiv:

[386]

S. Narison, QCD as a theory of hadrons: From partons to confinement, Camb. Monogr. Part. Phys. Nucl. Phys. Cosmol. 17, 1 (2007)

[387]

P. Gubler and D. Satow, Recent progress in QCD condensate evaluations and sum rules, Prog. Part. Nucl. Phys. 106, 1 (2019)

[388]

D. B. Leinweber, QCD sum rules for skeptics, Annals Phys. 254, 328 (1997)

[389]

I. I. Balitsky, D. Diakonov, and A. V. Yung, Exotic mesons with JPC=1−+ from QCD sum rules, Phys. Lett. B 112, 71 (1982)

[390]

J. Govaerts, F. de Viron, D. Gusbin, and J. Weyers, Exotic mesons from QCD sum rules, Phys. Lett. B 128, 262 (1983)

[391]

J. Govaerts, F. de Viron, D. Gusbin, and J. Weyers, QCD Sum Rules and Hybrid Mesons, Nucl. Phys. B 248, 1 (1984)

[392]

J. I. Latorre, S. Narison, P. Pascual, and R. Tarrach, Hermaphrodite Mesons and QCD Sum Rules, Phys. Lett. B 147, 169 (1984)

[393]

J. Govaerts, L. J. Reinders, and J. Weyers, Radial Excitations and Exotic Mesons via QCD Sum Rules, Nucl. Phys. B 262, 575 (1985)

[394]

S. A. Larin, V. A. Matveev, A. A. Ovchinnikov, and A. A. Pivovarov, Determination of the mass of the ΛΛ dibaryon by the method of QCD sum rules, Sov. J. Nucl. Phys. 44, 690 (1986)

[395]

I. I. Balitsky, D. Diakonov, and A. V. Yung, Exotic mesons with JPC=1−+ strange and non-strange, Z. Phys. C 33, 265 (1986)

[396]

V. M. Braun and A. V. Kolesnichenko, Exotic Scalar JPC=0+− Meson From QCD Sum Rules, Phys. Lett. B 175, 485 (1986)

[397]

J. I. Latorre, P. Pascual, and S. Narison, Spectra and Hadronic Couplings of Light Hermaphrodite Mesons, Z. Phys. C 34, 347 (1987)

[398]

J. Govaerts, L. J. Reinders, P. Francken, X. Gonze, and J. Weyers, Coupled QCD sum rules for hybrid mesons, Nucl. Phys. B 284, 674 (1987)

[399]

V. M. Braun and Y. M. Shabelski, Four quark exotic mesons u¯u¯sd from QCD sum rules, Sov. J. Nucl. Phys. 50, 306 (1989)

[400]

L. S. Kisslinger and Z. P. Li, Hybrid baryons via QCD sum rules, Phys. Rev. D 51, R5986 (1995)

[401]

Z. G. Wang, Two-particle contributions and nonlocal effects in the QCD sum rules for the axial vector tetraquark candidate Zc(3900), Int. J. Mod. Phys. A 35, 2050138 (2020)

[402]

Z. G. Wang, Analysis of the Pc(4312), Pc(4440), Pc(4457) and related hidden-charm pentaquark states with QCD sum rules, Int. J. Mod. Phys. A 35, 2050003 (2020)

[403]

Y. Kondo, O. Morimatsu, and T. Nishikawa, Two-hadron-irreducible QCD sum rule for pentaquark baryon, Phys. Lett. B 611, 93 (2005)

[404]

S. H. Lee, H. Kim, and Y. Kwon, Parity of Θ+(1540) from QCD sum rules, Phys. Lett. B 609, 252 (2005)

[405]

H. J. Lee and N. I. Kochelev, On the ππ contribution to the QCD sum rules for the light tetraquark, Phys. Rev. D 78, 076005 (2008)

[406]

H. X. Chen, A. Hosaka, H. Toki, and S. L. Zhu, Light scalar meson σ(600) in QCD sum rule with continuum, Phys. Rev. D 81, 114034 (2010)

[407]

W. Lucha, D. Melikhov, and H. Sazdjian, Tetraquark-adequate formulation of QCD sum rules, Phys. Rev. D 100, 014010 (2019)

[408]

W. Lucha, D. Melikhov, and H. Sazdjian, Tetraquark-adequate QCD sum rules for quark-exchange processes, Phys. Rev. D 100, 074029 (2019)

[409]

Z. G. Wang, Landau equation and QCD sum rules for the tetraquark molecular states, Phys. Rev. D 101, 074011 (2020)

[410]

L. D. Landau, On analytic properties of vertex parts in quantum field theory, Nucl. Phys. 13, 181 (1959)

[411]

M. Tanabashi, . The Review of Particle Physics, Phys. Rev. D 98, 030001 (2018)

[412]

N. Gray, D. J. Broadhurst, W. Grafe, and K. Schilcher, Three loop relation of quark (modified) MS and pole masses, Z. Phys. C 48, 673 (1990)

[413]

Z. G. Wang, Triply-charmed dibaryon states or two-baryon scattering states from the QCD sum rules?, Phys. Rev. D 102, 034008 (2020)

[414]

K. U. Can,G. Erkol,B. Isildak,M. Oka,T. Takahashi, Electromagnetic structure of charmed baryons in Lattice QCD, PoS LATTICE2014 (2015) 157

[415]

J. Y. Kim and H. C. Kim, Electromagnetic form factors of singly heavy baryons in the self-consistent SU(3) chiral quark-soliton model, Phys. Rev. D 97, 114009 (2018)

[416]

C. W. Hwang, Charge radii of light and heavy mesons, Eur. Phys. J. C 23, 585 (2002)

[417]

Z. G. Wang, Comment on “OPE and quark-hadron duality for two-point functions of tetraquark currents in 1/Nc expansion” arXiv:

[418]

Z. G. Wang and Q. Xin, Analysis of the hidden-charm pentaquark molecular states with and without strangeness via the QCD sum rules, Chin. Phys. C 45, 123105 (2021)

[419]

Z. G. Wang, Analysis of the masses and decay constants of the heavy-light mesons with QCD sum rules, Eur. Phys. J. C 75, 427 (2015)

[420]

K. G. Chetyrkin and M. Steinhauser, Three loop nondiagonal current correlators in QCD and NLO corrections to single top quark production, Phys. Lett. B 502, 104 (2001)

[421]

K. G. Chetyrkin and M. Steinhauser, Heavy-light current correlators at order αs2 in QCD and HQET, Eur. Phys. J. C 21, 319 (2001)

[422]

Z. G. Wang, Reanalysis of the Y(3940), Y(4140), Zc(4020), Zc(4025) and Zb(10650) as molecular states with QCD sum rules, Eur. Phys. J. C 74, 2963 (2014)

[423]

Z. G. Wang, Analysis of Pc(4380) and Pc(4450) as pentaquark states in the diquark model with QCD sum rules, Eur. Phys. J. C 76, 70 (2016)

[424]

Z. G. Wang, Analysis of the Zc(4020), Zc(4025), Y(4360) and Y(4660) as vector tetraquark states with QCD sum rules, Eur. Phys. J. C 74, 2874 (2014)

[425]

Z. G. Wang, Analysis of the hidden-charm-hidden-strange tetraquark mass spectrum via the QCD sum rules, Nucl. Phys. B 1007, 116661 (2024)

[426]

Z. G. Wang, Analysis of the tetraquark and hexaquark molecular states with the QCD sum rules, Commun. Theor. Phys. 73, 065201 (2021)

[427]

S. H. Lee, K. Morita, and M. Nielsen, Width of exotics from QCD sum rules: Tetraquarks or molecules?, Phys. Rev. D 78, 076001 (2008)

[428]

S. H. Lee, M. Nielsen, and U. Wiedner, DsD∗ molecule as an axial meson, J. Korean Phys. Soc. 55, 424 (2009)

[429]

M. E. Bracco, S. H. Lee, M. Nielsen, and R. Rodrigues da Silva, The meson Z+(4430) as a tetraquark state, Phys. Lett. B 671, 240 (2009)

[430]

S. H. Lee, K. Morita, and M. Nielsen, Can the π+χc1 resonance structures be D∗D¯∗ and D1D¯ molecules?, Nucl. Phys. A 815, 29 (2009)

[431]

R. M. Albuquerque,M. Nielsen, QCD sum rules study of the JPC=1−− charmonium Y mesons, Nucl. Phys. A 815 (2009) 53; Erratum-ibid. A 857 (2011) 48

[432]

R. M. Albuquerque, M. E. Bracco, and M. Nielsen, A QCD sum rule calculation for the Y(4140) narrow structure, Phys. Lett. B 678, 186 (2009)

[433]

R. D. Matheus, F. S. Navarra, M. Nielsen, and C. M. Zanetti, QCD sum rules for the X(3872) as a mixed molecule-charmoniun state, Phys. Rev. D 80, 056002 (2009)

[434]

J. R. Zhang and M. Q. Huang, {Qq¯}{Q¯(′)q} molecular states, Phys. Rev. D 80, 056004 (2009)

[435]

J. R. Zhang and M. Q. Huang, (Qs¯)(∗)(Q¯s)(∗) molecular states from QCD sum rules: A view on Y(4140), J. Phys. G 37, 025005 (2010)

[436]

J. R. Zhang and M. Q. Huang, {Qs¯}{Q¯(′)s} molecular states in QCD sum rules, Commun. Theor. Phys. 54, 1075 (2010)

[437]

W. Chen and S. L. Zhu, Possible JPC=0−− charmonium-like state, Phys. Rev. D 81, 105018 (2010)

[438]

M. Nielsen and C. M. Zanetti, Radiative decay of the X(3872) as a mixed molecule-charmonium state in QCD sum rules, Phys. Rev. D 82, 116002 (2010)

[439]

S. Narison, F. S. Navarra, and M. Nielsen, On the nature of the X(3872) from QCD, Phys. Rev. D 83, 016004 (2011)

[440]

R. M. Albuquerque, M. Nielsen, and R. Rodrigues da Silva, Exotic 1−− states in QCD sum rules, Phys. Rev. D 84, 116004 (2011)

[441]

J. M. Dias, S. Narison, F. S. Navarra, M. Nielsen, and J. M. Richard, Relation between Tcc,bb and Xc,b from QCD, Phys. Lett. B 703, 274 (2011)

[442]

S. I. Finazzo, X. Liu, and M. Nielsen, QCD sum rule calculation for the charmonium-like structures in the J/ψϕ and J/ψω invariant mass spectra, Phys. Lett. B 701, 101 (2011)

[443]

J. R. Zhang, M. Zhong, and M. Q. Huang, Could Zb(10610) be a B∗B¯ molecular state?, Phys. Lett. B 704, 312 (2011)

[444]

W. Chen and S. L. Zhu, The vector and axial-vector charmonium-like states, Phys. Rev. D 83, 034010 (2011)

[445]

C. Y. Cui, Y. L. Liu, and M. Q. Huang, Investigating different structures of the Zb(10610) and Zb(10650), Phys. Rev. D 85, 074014 (2012)

[446]

J. M. Dias, F. S. Navarra, M. Nielsen, and C. M. Zanetti, Zc+(3900) decay width in QCD sum rules, Phys. Rev. D 88, 016004 (2013)

[447]

W. Chen, H. Y. Jin, R. T. Kleiv, T. G. Steele, M. Wang, and Q. Xu, QCD sum-rule interpretation of X(3872) with JPC=1++ mixtures of hybrid charmonium and D¯D∗ molecular currents, Phys. Rev. D 88, 045027 (2013)

[448]

J. R. Zhang, Improved QCD sum rule study of Zc(3900) as a D¯D∗ molecular state, Phys. Rev. D 87, 116004 (2013)

[449]

C. F. Qiao and L. Tang, Estimating the mass of the hidden charm 1+(1+) tetraquark state via QCD sum rules, Eur. Phys. J. C 74, 3122 (2014)

[450]

C. F. Qiao and L. Tang, Interpretation of Zc(4025) as the hidden charm tetraquark states via QCD Sum Rules, Eur. Phys. J. C 74, 2810 (2014)

[451]

W. Chen, T. G. Steele, H. X. Chen, and S. L. Zhu, Mass spectra of Zc and Zb exotic states as hadron molecules, Phys. Rev. D 92, 054002 (2015)

[452]

R. Albuquerque, S. Narison, F. Fanomezana, A. Rabemananjara, D. Rabetiarivony, and G. Randriamanatrika, XYZ-like Spectra from Laplace Sum Rule at N2LO in the Chiral Limit, Int. J. Mod. Phys. A 31, 1650196 (2016)

[453]

L. Tang and C. F. Qiao, Tetraquark states with open flavors, Eur. Phys. J. C 76, 558 (2016)

[454]

R. Albuquerque, S. Narison, D. Rabetiarivony, and G. Randriamanatrika, XYZ-SU3 breakings from Laplace sum rules at higher orders, Int. J. Mod. Phys. A 33, 1850082 (2018)

[455]

L. Tang, B. D. Wan, K. Maltman, and C. F. Qiao, Doubly heavy tetraquarks in QCD sum rules, Phys. Rev. D 101, 094032 (2020)

[456]

R. M. Albuquerque, S. Narison, A. Rabemananjara, D. Rabetiarivony, and G. Randriamanatrika, Doubly-hidden scalar heavy molecules and tetraquarks states from QCD at NLO, Phys. Rev. D 102, 094001 (2020)

[457]

B. D. Wan and C. F. Qiao, About the exotic structure of Zcs, Nucl. Phys. B 968, 115450 (2021)

[458]

W. Chen, T. G. Steele, and S. L. Zhu, Exotic open-flavor bcq¯q¯, bcs¯s¯ and qcq¯b¯, scs¯b¯ tetraquark states, Phys. Rev. D 89, 054037 (2014)

[459]

B. C. Yang, L. Tang, and C. F. Qiao, Scalar fully-heavy tetraquark states QQ′Q¯Q′¯ in QCD sum rules, Eur. Phys. J. C 81, 324 (2021)

[460]

J. R. Zhang, 0+ fully-charmed tetraquark states, Phys. Rev. D 103, 014018 (2021)

[461]

R. M. Albuquerque, S. Narison, and D. Rabetiarivony, Zc-like spectra from QCD Laplace sum rules at NLO, Phys. Rev. D 103, 074015 (2021)

[462]

S. S. Agaev, K. Azizi, and H. Sundu, Strong Zc+(3900)→J/ψπ+; ηcρ+ decays in QCD, Phys. Rev. D 93, 074002 (2016)

[463]

U. Ozdem and K. Azizi, Magnetic and quadrupole moments of the Zc(3900), Phys. Rev. D 96, 074030 (2017)

[464]

S. S. Agaev, K. Azizi, and H. Sundu, Treating Zc(3900) and Z(4430) as the ground-state and first radially excited tetraquarks, Phys. Rev. D 96, 034026 (2017)

[465]

H. Sundu, S. S. Agaev, and K. Azizi, New charged resonance Zc(4100): The spectroscopic parameters and width, Eur. Phys. J. C 79, 215 (2019)

[466]

S. S. Agaev, K. Azizi, B. Barsbay, and H. Sundu, The doubly charmed pseudoscalar tetraquarks Tcc;s¯s¯++ and Tcc;d¯s¯++, Nucl. Phys. B 939, 130 (2019)

[467]

S. Agaev, K. Azizi, and H. Sundu, Four-quark exotic mesons, Turk. J. Phys. 44, 95 (2020)

[468]

U. Ozdem and K. Azizi, Magnetic dipole moment of the Zcs(3985) state: Diquark−antidiquark and molecular pictures, Eur. Phys. J. Plus 136, 968 (2021)

[469]

U. Ozdem and A. K. Yildirim, Magnetic dipole moments of the Zc(4020)+, Zc(4200)+, Zcs(4000)+ and Zcs(4220)+ states in light-cone QCD, Phys. Rev. D 104, 054017 (2021)

[470]

K. Azizi and N. Er, The newly observed Zcs(3985)− state: In vacuum and a dense medium, Eur. Phys. J. C 81, 61 (2021)

[471]

S. S. Agaev, K. Azizi, and H. Sundu, Hadronic molecule model for the doubly charmed state Tcc+, JHEP 06, 057 (2022)

[472]

S. S. Agaev, K. Azizi, and H. Sundu, Newly observed exotic doubly charmed meson Tcc+, Nucl. Phys. B 975, 115650 (2022)

[473]

H. Mutuk, Molecular interpretation of X(3960) as Ds+Ds− state, Eur. Phys. J. C 82, 1142 (2022)

[474]

H. Mutuk, Monte-Carlo based QCD sum rules analysis of X0(2900) and X1(2900), J. Phys. G 48, 055007 (2021)

[475]

Z. G. Wang, Mass spectrum of the vector hidden charmed and bottomed tetraquark states, J. Phys. G 36, 085002 (2009)

[476]

Z. G. Wang, Mass spectrum of the scalar hidden charmed and bottomed tetraquark states, Phys. Rev. D 79, 094027 (2009)

[477]

Z. G. Wang, Analysis of the Y(4140) with QCD sum rules, Eur. Phys. J. C 63, 115 (2009)

[478]

Z. G. Wang, Z. C. Liu, and X. H. Zhang, Analysis of the Y(4140) and related molecular states with QCD sum rules, Eur. Phys. J. C 64, 373 (2009)

[479]

Z. G. Wang, Possible tetraquark state in the π+χc1 invariant mass distribution, Eur. Phys. J. C 59, 675 (2009)

[480]

Z. G. Wang, Another tetraquark structure in the π+χc1 invariant mass distribution, Eur. Phys. J. C 62, 375 (2009)

[481]

Z. G. Wang, Reanalysis of the mass spectrum of the scalar hidden charm and hidden bottom tetraquark states, Eur. Phys. J. C 67, 411 (2010)

[482]

Z. G. Wang and X. H. Zhang, Analysis of the pseudoscalar partner of the Y(4660) and related bound states, Eur. Phys. J. C 66, 419 (2010)

[483]

Z. G. Wang, Mass spectrum of the axial-vector hidden charmed and hidden bottom tetraquark states, Eur. Phys. J. C 70, 139 (2010)

[484]

Z. G. Wang, Analysis of the X(4350) as a scalar c¯c and Ds∗D¯s∗ mixing state with QCD sum rules, Phys. Lett. B 690, 403 (2010)

[485]

Z. G. Wang and X. H. Zhang, Analysis of Y(4660) and related bound states with QCD sum rules, Commun. Theor. Phys. 54, 323 (2010)

[486]

Z. G. Wang, Analysis of the Y(4274) with QCD sum rules, Int. J. Mod. Phys. A 26, 4929 (2011)

[487]

Z. G. Wang, Analysis of the D¯Σc, D¯Σc∗, D¯∗Σc and D¯∗Σc∗ pentaquark molecular states with QCD sum rules, Int. J. Mod. Phys. A 34, 1950097 (2019)

[488]

H. X. Chen, A. Hosaka, and S. L. Zhu, Light scalar tetraquark mesons in the QCD sum rule, Phys. Rev. D 76, 094025 (2007)

[489]

H. X. Chen, A. Hosaka, and S. L. Zhu, The IGJPC=1−1−+ tetraquark states, Phys. Rev. D 78, 054017 (2008)

[490]

C. K. Jiao, W. Chen, H. X. Chen, and S. L. Zhu, The possible JPC=0−− exotic state, Phys. Rev. D 79, 114034 (2009)

[491]

Z. R. Huang, W. Chen, T. G. Steele, Z. F. Zhang, and H. Y. Jin, Investigation of the light four-quark states with exotic JPC=0−−, Phys. Rev. D 95, 076017 (2017)

[492]

W. Chen, H. X. Chen, X. Liu, T. G. Steele, and S. L. Zhu, Mass spectra for qcq¯c¯, scs¯c¯, qbq¯b¯, sbs¯b¯ tetraquark states with JPC=0++ and 2++, Phys. Rev. D 96, 114017 (2017)

[493]

W. Chen, H. X. Chen, X. Liu, T. G. Steele, and S. L. Zhu, Hunting for exotic doubly hidden-charm/bottom tetraquark states, Phys. Lett. B 773, 247 (2017)

[494]

Z. G. Wang, Analysis of the Y(2175) as a tetraquark state with QCD sum rules, Nucl. Phys. A 791, 106 (2007)

[495]

Z. G. Wang, Analysis of the mass and width of the Y(4274) as axialvector molecule-like state, Eur. Phys. J. C 77, 174 (2017)

[496]

Z. G. Wang, Analysis of the Zc(4200) as axial-vector molecule-like state, Int. J. Mod. Phys. A 30, 1550168 (2015)

[497]

C. M. Tang, C. G. Duan, and L. Tang, Fully charmed tetraquark states in 8[cc¯]⊗8[cc¯] color structure via QCD sum rules, Eur. Phys. J. C 84, 743 (2024)

[498]

Z. G. Wang, Analysis of the scalar and axial-vector heavy diquark states with QCD sum rules, Eur. Phys. J. C 71, 1524 (2011)

[499]

L. Tang and X. Q. Li, Discussions on stability of diquarks, Chin. Phys. C 36, 578 (2012)

[500]

R. T. Kleiv, T. G. Steele, A. L. Zhang, and I. Blokland, Heavy-light diquark masses from QCD sum rules and constituent diquark models of tetraquarks, Phys. Rev. D 87, 125018 (2013)

[501]

Z. G. Wang, Analysis of the light-flavor scalar and axial-vector diquark states with QCD sum rules, Commun. Theor. Phys. 59, 451 (2013)

[502]

H. G. Dosch, M. Jamin, and B. Stech, Diquarks, QCD sum rules and weak decays, Z. Phys. C 42, 167 (1989)

[503]

M. Jamin and M. Neubert, Diquark decay constants From QCD sum rules, Phys. Lett. B 238, 387 (1990)

[504]

A. L. Zhang, T. Huang, and T. G. Steele, Diquark and light four-quark states, Phys. Rev. D 76, 036004 (2007)

[505]

Z. G. Wang, Scalar or vector tetraquark state candidate: Zc(4100), Commun. Theor. Phys. 71, 1319 (2019)

[506]

Z. G. Wang and Y. F. Tian, Tetraquark state candidates: Y(4140), Y(4274) and X(4350), Int. J. Mod. Phys. A 30, 1550004 (2015)

[507]

Z. G. Wang and Z. Y. Di, Analysis of the mass and width of the X(4140) as axialvector tetraquark state, Eur. Phys. J. C 79, 72 (2019)

[508]

P. Pascual and R. Tarrach, QCD: renormalization for the practitioner, Lect. Notes Phys. 194, 1 (1984)

[509]

X. W. Wang, Z. G. Wang, and G. L. Yu, Study of ΛcΛc dibaryon and ΛcΛ¯c baryonium states via QCD sum rules, Eur. Phys. J. A 57, 275 (2021)

[510]

J. Beringer, . Review of particle physics, Phys. Rev. D 86, 010001 (2012)

[511]

S. Narison and R. Tarrach, Higher dimensional renormalization group invariant vacuum condensates in quantum chromodynamics, Phys. Lett. B 125, 217 (1983)

[512]

Z. G. Wang, Tetraquark state candidates: Y(4260), Y(4360), Y(4660) and Zc(4020/4025), Eur. Phys. J. C 76, 387 (2016)

[513]

Q. Xin and Z. G. Wang, Analysis of the doubly-charmed tetraquark molecular states with the QCD sum rules, Eur. Phys. J. A 58, 110 (2022)

[514]

Z. G. Wang, Analysis of the Pcs(4459) as the hidden-charm pentaquark state with QCD sum rules, Int. J. Mod. Phys. A 36, 2150071 (2021)

[515]

Z. G. Wang and Q. Xin, Analysis of the pseudoscalar hidden-charm tetraquark states with the QCD sum rules, Nucl. Phys. B 978, 115761 (2022)

[516]

R. Albuquerque, S. Narison, A. Rabemananjara, and D. Rabetiarivony, Nature of the X(5568): A critical Laplace sum rule analysis at N2LO, Int. J. Mod. Phys. A 31, 1650093 (2016)

[517]

R. M. Albuquerque, S. Narison, and D. Rabetiarivony, Tests of the Zc-like Laplace sum rule results using finite energy sum rule at NLO, Phys. Rev. D 105, 114035 (2022)

[518]

R. H. Wu, C. Y. Wang, C. Meng, Y. Q. Ma, and K. T. Chao, Zc and Zcs systems with operator mixing at NLO in QCD sum rules, JHEP 06, 216 (2024)

[519]

R. H. Wu, Y. S. Zuo, C. Y. Wang, C. Meng, Y. Q. Ma, and K. T. Chao, NLO results with operator mixing for fully heavy tetraquarks in QCD sum rules, JHEP 11, 023 (2022)

[520]

Z. G. Wang and J. X. Zhang, The decay width of the Zc(3900) as an axialvector tetraquark state in solid quark-hadron duality, Eur. Phys. J. C 78, 14 (2018)

[521]

A. Esposito, A. L. Guerrieri, and A. Pilloni, Probing the nature of Zc(′) states via the ηcρ decay, Phys. Lett. B 746, 194 (2015)

[522]

Z. G. Wang, Assignments of the X(4140), X(4500), X(4630) and X(4685) Based on the QCD Sum Rules, Adv. High Energy Phys. 2021, 4426163 (2021)

[523]

Z. G. Wang, Revisit the X(4274) as the axialvector tetraquark state, Acta Phys. Polon. B 51, 435 (2020)

[524]

F. Stancu, Can Y(4140) be a cc¯ss¯ tetraquark?, J. Phys. G 37, 075017 (2010)

[525]

R. L. Zhu, Hidden charm octet tetraquarks from a diquark-antidiquark model, Phys. Rev. D 94, 054009 (2016)

[526]

J. Wu, Y. R. Liu, K. Chen, X. Liu, and S. L. Zhu, X(4140), X(4270), X(4500) and X(4700) and their csc¯s¯ tetraquark partners, Phys. Rev. D 94, 094031 (2016)

[527]

H. X. Chen, E. L. Cui, W. Chen, X. Liu, and S. L. Zhu, Understanding the internal structures of the X(4140), X(4274), X(4500) and X(4700), Eur. Phys. J. C 77, 160 (2017)

[528]

N. Mahajan, Y(4140): Possible options, Phys. Lett. B 679, 228 (2009)

[529]

X. H. Liu, How to understand the underlying structures of X(4140), X(4274), X(4500) and X(4700), Phys. Lett. B 766, 117 (2017)

[530]

Z. G. Wang, Analysis of the X(4475), X(4500), Zc¯s¯(4600) and related tetraquark states with the QCD sum rules, arXiv:

[531]

Z. G. Wang, Analysis of the hidden-bottom tetraquark mass spectrum with the QCD sum rules, Eur. Phys. J. C 79, 489 (2019)

[532]

Z. G. Wang and T. Huang, The Zb(10610) and Zb(10650) as axial-vector tetraquark states in the QCD sum rules, Nucl. Phys. A 930, 63 (2014)

[533]

A. Ali, C. Hambrock, and W. Wang, Tetraquark interpretation of the charged bottomonium-like states Zb±(10610) and Zb±(10650) and implications, Phys. Rev. D 85, 054011 (2012)

[534]

A. Ali, L. Maiani, A. D. Polosa, and V. Riquer, Hidden-beauty charged tetraquarks and heavy quark spin conservation, Phys. Rev. D 91, 017502 (2015)

[535]

Z. G. Wang, Analysis of the axialvector Bc-like tetraquark states with the QCD sum rules, EPL 128, 11001 (2019)

[536]

S. S. Agaev, K. Azizi, and H. Sundu, Open charm-bottom scalar tetraquarks and their strong decays, Phys. Rev. D 95, 034008 (2017)

[537]

U. Ozdem, Electromagnetic form factors of the Bc-like tetraquarks: Molecular and diquark-antidiquark pictures, Phys. Lett. B 838, 137750 (2023)

[538]

M. S. Maior de Sousa and R. Rodrigues da Silva, The ρ(1S,2S), ψ(1S,2S), Υ(1S,2S) and ψt(1S,2S) mesons in a double pole QCD sum rule, Braz. J. Phys. 46, 730 (2016)

[539]

K. Azizi and N. Er, Modifications on parameters of Z(4430) in a dense medium, Phys. Lett. B 811, 135979 (2020)

[540]

Z. G. Wang, Scalar tetraquark state candidates: X(3915), X(4500) and X(4700), Eur. Phys. J. C 77, 78 (2017)

[541]

Z. G. Wang, Reanalysis of the X(3915), X(4500) and X(4700) with QCD sum rules, Eur. Phys. J. A 53, 19 (2017)

[542]

Z. G. Wang and H. J. Wang, Analysis of the 1S and 2S states of ΛQ and ΞQ with QCD sum rules, Chin. Phys. C 45, 013109 (2021)

[543]

Z. G. Wang, Analysis of the vector hidden-charm tetraquark states without explicit P-waves via the QCD sum rules, Nucl. Phys. B 973, 115592 (2021)

[544]

Z. G. Wang, Vector tetraquark state candidates: Y(4260/4220), Y(4360/4320), Y(4390) and Y(4660/4630), Eur. Phys. J. C 78, 518 (2018)

[545]

Z. G. Wang, Analysis of the vector hidden-charm-hidden-strange tetraquark states with implicit P-waves via the QCD sum rules, Nucl. Phys. B 1002, 116514 (2024)

[546]

Z. G. Wang, Strong decays of the vector tetraquark states with the masses about 4.5GeV via the QCD sum rules, Nucl. Phys. B 1005, 116580 (2024)

[547]

Z. G. Wang, Analysis of the decay Y(4500)→D∗D¯∗π with the light-cone QCD sum rules, Nucl. Phys. B 993, 116265 (2023)

[548]

J. R. Zhang and M. Q. Huang, The P-wave [cs][c¯s¯] tetraquark state: Y(4260) or Y(4660)?, Phys. Rev. D 83, 036005 (2011)

[549]

Z. G. Wang, Lowest vector tetraquark states: Y(4260/4220) or Zc(4100), Eur. Phys. J. C 78, 933 (2018)

[550]

Z. G. Wang, Analysis of the vector tetraquark states with P-waves between the diquarks and antidiquarks via the QCD sum rules, Eur. Phys. J. C 79, 29 (2019)

[551]

S. L. Zhu, The Possible interpretations of Y(4260), Phys. Lett. B 625, 212 (2005)

[552]

F. E. Close and P. R. Page, Gluonic charmonium resonances at BaBar and Belle?, Phys. Lett. B 628, 215 (2005)

[553]

E. Kou and O. Pene, Suppressed decay into open charm for the Y(4260) being an hybrid, Phys. Lett. B 631, 164 (2005)

[554]

L. Liu, . Excited and exotic charmonium spectroscopy from lattice QCD, JHEP 07, 126 (2012)

[555]

X. K. Dong, F. K. Guo, and B. S. Zou, A survey of heavy-antiheavy hadronic molecules, Progr. Phys. 41, 65 (2021)

[556]

T. W. Chiu and T. H. Hsieh, Y(4260) on the lattice, Phys. Rev. D 73, 094510 (2006)

[557]

C. F. Qiao, One explanation for the exotic state Y(4260), Phys. Lett. B 639, 263 (2006)

[558]

C. F. Qiao, A Uniform description of the states recently observed at B-factories, J. Phys. G 35, 075008 (2008)

[559]

X. Li and M. B. Voloshin, Y(4260) and Y(4360) as mixed hadrocharmonium, Mod. Phys. Lett. A 29, 1450060 (2014)

[560]

D. Y. Chen, J. He, and X. Liu, Nonresonant explanation for the Y(4260) structure observed in the e+e−→J/ψπ+π− process, Phys. Rev. D 83, 054021 (2011)

[561]

D. Y. Chen, X. Liu, X. Q. Li, and H. W. Ke, Unified Fano-like interference picture for charmoniumlike states Y(4008), Y(4260) and Y(4360), Phys. Rev. D 93, 014011 (2016)

[562]

Z. G. Wang, Ground states and first radial excitations of vector tetraquark states with explicit P-waves via QCD sum rules, Chin. Phys. C 48, 103103 (2024)

[563]

Z. G. Wang, Vector hidden-bottom tetraquark candidate: Y(10750), Chin. Phys. C 43, 123102 (2019)

[564]

R. Aaij, . Observation of the doubly charmed baryon Ξcc++, Phys. Rev. Lett. 119, 112001 (2017)

[565]

Z. G. Wang, Analysis of the doubly heavy baryon states and pentaquark states with QCD sum rules, Eur. Phys. J. C 78, 826 (2018)

[566]

M. Karliner and J. L. Rosner, Discovery of doubly-charmed Ξcc baryon implies a stable bbu¯d¯ tetraquark, Phys. Rev. Lett. 119, 202001 (2017)

[567]

E. J. Eichten and C. Quigg, Heavy-quark symmetry implies stable heavy tetraquark mesons QiQjq¯kq¯l, Phys. Rev. Lett. 119, 202002 (2017)

[568]

Q. Qin, Y. F. Shen, and F. S. Yu, Discovery potentials of double-charm tetraquarks, Chin. Phys. C 45, 103106 (2021)

[569]

J. P. Ader, J. M. Richard, and P. Taxil, Do narrow heavy multiquark states exist?, Phys. Rev. D 25, 2370 (1982)

[570]

L. Heller and J. A. Tjon, On the existence of stable dimesons, Phys. Rev. D 35, 969 (1987)

[571]

J. Carlson, L. Heller, and J. A. Tjon, Stability of Dimesons, Phys. Rev. D 37, 744 (1988)

[572]

S. Zouzou, B. Silvestre-Brac, C. Gignoux, and J. M. Richard, Four-quark bound states, Z. Phys. C 30, 457 (1986)

[573]

D. M. Brink and F. Stancu, Tetraquarks with heavy flavors, Phys. Rev. D 57, 6778 (1998)

[574]

A. Czarnecki, B. Leng, and M. B. Voloshin, Stability of tetrons, Phys. Lett. B 778, 233 (2018)

[575]

A. V. Manohar and M. B. Wise, Exotic QQq¯q¯ states in QCD, Nucl. Phys. B 399, 17 (1993)

[576]

Z. G. Wang, Analysis of the axialvector doubly heavy tetraquark states with QCD sum rules, Acta Phys. Polon. B 49, 1781 (2018)

[577]

M. L. Du, W. Chen, X. L. Chen, and S. L. Zhu, Exotic QQq¯q¯, QQq¯s¯ and QQs¯s¯ states, Phys. Rev. D 87, 014003 (2013)

[578]

Z. G. Wang and Z. H. Yan, Analysis of the scalar, axialvector, vector, tensor doubly charmed tetraquark states with QCD sum rules, Eur. Phys. J. C 78, 19 (2018)

[579]

M. A. Moinester, How to search for doubly charmed baryons and tetraquarks, Z. Phys. A 355, 349 (1996)

[580]

S. Pepin, F. Stancu, M. Genovese, and J. M. Richard, Tetraquarks with color blind forces in chiral quark models, Phys. Lett. B 393, 119 (1997)

[581]

B. A. Gelman and S. Nussinov, Does a narrow tetraquark ccu¯d¯ state exist?, Phys. Lett. B 551, 296 (2003)

[582]

D. Janc and M. Rosina, The Tcc=DD∗ molecular state, Few Body Syst. 35, 175 (2004)

[583]

F. S. Navarra, M. Nielsen, and S. H. Lee, QCD sum rules study of QQu¯d¯ mesons, Phys. Lett. B 649, 166 (2007)

[584]

Z. G. Wang, Y. M. Xu, and H. J. Wang, Analysis of the scalar doubly heavy tetraquark states with QCD sum rules, Commun. Theor. Phys. 55, 1049 (2011)

[585]

J. Vijande, E. Weissman, A. Valcarce, and N. Barnea, Are there compact heavy four-quark bound states?, Phys. Rev. D 76, 094027 (2007)

[586]

S. H. Lee and S. Yasui, Stable multiquark states with heavy quarks in a diquark model, Eur. Phys. J. C 64, 283 (2009)

[587]

Y. Yang, C. Deng, J. Ping, and T. Goldman, S-wave QQq¯q¯ state in the constituent quark model, Phys. Rev. D 80, 114023 (2009)

[588]

S. Q. Luo, K. Chen, X. Liu, Y. R. Liu, and S. L. Zhu, Exotic tetraquark states with the qqQ¯Q¯ configuration, Eur. Phys. J. C 77, 709 (2017)

[589]

W. Park, S. Noh, and S. H. Lee, Masses of the doubly heavy tetraquarks in a constituent quark model, Nucl. Phys. A 983, 1 (2019)

[590]

L. Maiani, A. D. Polosa, and V. Riquer, Hydrogen bond of QCD in doubly heavy baryons and tetraquarks, Phys. Rev. D 100, 074002 (2019)

[591]

G. Yang, J. Ping, and J. Segovia, Doubly-heavy tetraquarks, Phys. Rev. D 101, 014001 (2020)

[592]

E. Braaten, L. P. He, and A. Mohapatra, Masses of doubly heavy tetraquarks with error bars, Phys. Rev. D 103, 016001 (2021)

[593]

D. Gao, D. Jia, and Y. J. Sun, Masses of doubly heavy tetraquarks QQn¯q¯ with JP=1+, Mod. Phys. Lett. A 37, 2250223 (2022)

[594]

J. B. Cheng, S. Y. Li, Y. R. Liu, Z. G. Si, and T. Yao, Double-heavy tetraquark states with heavy diquark-antiquark symmetry, Chin. Phys. C 45, 043102 (2021)

[595]

R. N. Faustov, V. O. Galkin, and E. M. Savchenko, Heavy tetraquarks in the relativistic quark model, Universe 7, 94 (2021)

[596]

Y. Ikeda, . Charmed tetraquarks Tcc and Tcs from dynamical Lattice QCD simulations, Phys. Lett. B 729, 85 (2014)

[597]

X. Z. Weng, W. Z. Deng, and S. L. Zhu, Doubly heavy tetraquarks in an extended chromomagnetic model, Chin. Phys. C 46, 013102 (2022)

[598]

K. Azizi and U. Ozdem, Magnetic dipole moments of the Tcc+ and ZV++ tetraquark states, Phys. Rev. D 104, 114002 (2021)

[599]

Y. Kim, M. Oka, and K. Suzuki, Doubly heavy tetraquarks in a chiral-diquark picture, Phys. Rev. D 105, 074021 (2022)

[600]

T. W. Wu and Y. L. Ma, Doubly heavy tetraquark multiplets as heavy antiquark-diquark symmetry partners of heavy baryons, Phys. Rev. D 107, L071501 (2023)

[601]

M. Karliner and J. L. Rosner, Doubly charmed strange tetraquark, Phys. Rev. D 105, 034020 (2022)

[602]

Y. Iwasaki, Is a state cc¯cc¯ found at 6.0GeV?, Phys. Rev. Lett. 36, 1266 (1976)

[603]

K. T. Chao, The cc-c¯c¯ (diquark−anti-diquark) states in e+e− annihilation, Z. Phys. C 7, 317 (1981)

[604]

A. M. Badalyan, B. L. Ioffe, and A. V. Smilga, Four-quark states in heavy quark systems, Nucl. Phys. B 281, 85 (1987)

[605]

A. V. Berezhnoy, A. V. Luchinsky, and A. A. Novoselov, Tetraquarks composed of 4 heavy quarks, Phys. Rev. D 86, 034004 (2012)

[606]

M. Karliner, J. L. Rosner, and S. Nussinov, QQQ¯Q¯ states: Masses, production, and decays, Phys. Rev. D 95, 034011 (2017)

[607]

V. R. Debastiani and F. S. Navarra, A non-relativistic model for the ccc¯c¯ tetraquark, Chin. Phys. C 43, 013105 (2019)

[608]

M. S. Liu, Q. F. Lu, X. H. Zhong, and Q. Zhao, All-heavy tetraquarks, Phys. Rev. D 100, 016006 (2019)

[609]

R. J. Lloyd and J. P. Vary, All charm tetraquarks, Phys. Rev. D 70, 014009 (2004)

[610]

Z. G. Wang, Analysis of the QQQ¯Q¯ tetraquark states with QCD sum rules, Eur. Phys. J. C 77, 432 (2017)

[611]

Z. G. Wang and Z. Y. Di, Analysis of the vector and axialvector QQQ¯Q¯ tetraquark states with QCD sum rules, Acta Phys. Polon. B 50, 1335 (2019)

[612]

M. N. Anwar, J. Ferretti, F. K. Guo, E. Santopinto, and B. S. Zou, Spectroscopy and decays of the fully-heavy tetraquarks, Eur. Phys. J. C 78, 647 (2018)

[613]

C. Hughes, E. Eichten, and C. T. H. Davies, Searching for beauty-fully bound tetraquarks using lattice nonrelativistic QCD, Phys. Rev. D 97, 054505 (2018)

[614]

Z. G. Wang, Tetraquark candidates in the LHCb’s di-J/ψ mass spectrum, Chin. Phys. C 44, 113106 (2020)

[615]

Z. G. Wang, Analysis of the X(6600), X(6900), X(7300) and related tetraquark states with the QCD sum rules, Nucl. Phys. B 985, 115983 (2022)

[616]

Y. L. Song, Y. Zhang, V. Baru, F. K. Guo, C. Hanhart, and A. Nefediev, Towards a precision determination of the X(6200) parameters from data, Phys. Rev. D 111, 034038 (2025)

[617]

X. K. Dong, V. Baru, F. K. Guo, C. Hanhart, and A. Nefediev, Coupled-channel interpretation of the LHCb double-J/ψ spectrum and hints of a new state near the J/ψJ/ψ threshold, Phys. Rev. Lett. 126, 132001 (2021)

[618]

Z. H. Guo and J. A. Oller, Insights into the inner structures of the fully charmed tetraquark state X(6900), Phys. Rev. D 103, 034024 (2021)

[619]

Z. R. Liang, X. Y. Wu, and D. L. Yao, Hunting for states in the recent LHCb di-J/ψ invariant mass spectrum, Phys. Rev. D 104, 034034 (2021)

[620]

Q. F. Cao, H. Chen, H. R. Qi, and H. Q. Zheng, Some remarks on X(6900), Chin. Phys. C 45, 103102 (2021)

[621]

J. F. Giron and R. F. Lebed, The simple spectrum of cc¯cc¯ states in the dynamical diquark model, Phys. Rev. D 102, 074003 (2020)

[622]

M. Karliner and J. L. Rosner, Interpretation of structure in the di-J/ψ spectrum, Phys. Rev. D 102, 114039 (2020)

[623]

C. R. Deng, H. Chen, and J. L. Ping, Towards the understanding of fully-heavy tetraquark states from various models, Phys. Rev. D 103, 014001 (2021)

[624]

X. Z. Weng, X. L. Chen, W. Z. Deng, and S. L. Zhu, Systematics of fully heavy tetraquarks, Phys. Rev. D 103, 034001 (2021)

[625]

R. L. Zhu, Fully-heavy tetraquark spectra and production at hadron colliders, Nucl. Phys. B 966, 115393 (2021)

[626]

M. C. Gordillo, F. De Soto, and J. Segovia, Diffusion Monte Carlo calculations of fully-heavy multiquark bound states, Phys. Rev. D 102, 114007 (2020)

[627]

H. W. Ke, X. Han, X. H. Liu, and Y. L. Shi, Tetraquark state X(6900) and the interaction between diquark and antidiquark, Eur. Phys. J. C 81, 427 (2021)

[628]

H. Mutuk, Nonrelativistic treatment of fully-heavy tetraquarks as diquark−antidiquark states, Eur. Phys. J. C 81, 367 (2021)

[629]

K. T. Chao and S. L. Zhu, The possible tetraquark states ccc¯c¯ observed by the LHCb experiment, Sci. Bull. 65, 1952 (2020)

[630]

Q. Li, C. H. Chang, G. L. Wang, and T. Wang, Mass spectra and wave functions of TQQQ¯Q¯ tetraquarks, Phys. Rev. D 104, 014018 (2021)

[631]

F. X. Liu, M. S. Liu, X. H. Zhong, and Q. Zhao, Higher mass spectra of the fully-charmed and fully-bottom tetraquarks, Phys. Rev. D 104, 116029 (2021)

[632]

B. D. Wan and C. F. Qiao, Gluonic tetracharm configuration of X(6900), Phys. Lett. B 817, 136339 (2021)

[633]

Z. G. Wang, Revisit the tetraquark candidates in the J/ψJ/ψ mass spectrum, Int. J. Mod. Phys. A 36, 2150014 (2021)

[634]

Z. G. Wang, Analysis of the fully-heavy pentaquark states via the QCD sum rules, Nucl. Phys. B 973, 115579 (2021)

[635]

Z. G. Wang, Fully heavy hexaquark states via the QCD sum rules, Int. J. Mod. Phys. A 37, 2250166 (2022)

[636]

X. K. Dong, F. K. Guo, and B. S. Zou, A survey of heavy-heavy hadronic molecules, Commun. Theor. Phys. 73, 125201 (2021)

[637]

X. Liu and S. L. Zhu, Y(4143) is probably a molecular partner of Y(3930), Phys. Rev. D 80, 017502 (2009)

[638]

G. J. Ding, Possible Molecular states of Ds∗D¯s∗ system and Y(4140), Eur. Phys. J. C 64, 297 (2009)

[639]

Z. G. Wang, Reanalysis of the Zc(4020), Zc(4025), Z(4050) and Z(4250) as tetraquark states with QCD sum rules, Commun. Theor. Phys. 63, 466 (2015)

[640]

E. J. Eichten and C. Quigg, Mesons with beauty and charm: New horizons in spectroscopy, Phys. Rev. D 99, 054025 (2019)

[641]

Z. G. Wang, Analysis of the Y(4220) and Y(4390) as molecular states with QCD sum rules, Chin. Phys. C 41, 083103 (2017)

[642]

P. A. Zyla, . Review of particle physics, PTEP 2020, 083C01 (2020)

[643]

S. Prelovsek and L. Leskovec, Search for Zc+(3900) in the 1+− channel on the lattice, Phys. Lett. B 727, 172 (2013)

[644]

Y. Ikeda, . Fate of the Tetraquark candidate Zc(3900) from lattice QCD, Phys. Rev. Lett. 117, 242001 (2016)

[645]

X. L. Wang, . The study of γγ→γψ(2S) at Belle, Phys. Rev. D 105, 112011 (2022)

[646]

C. Patrignani, . Review of particle physics, Chin. Phys. C 40, 100001 (2016)

[647]

Y. H. Chen, L. Y. Dai, F. K. Guo, and B. Kubis, Nature of the Y(4260): A light-quark perspective, Phys. Rev. D 99, 074016 (2019)

[648]

Z. G. Wang, Comment on “Zc-like spectra from QCD Laplace sum rules at NLO” arXiv:

[649]

F. K. Guo, C. Hanhart, U. G. Meissner, Evidence that the Y(4660) is a f0(980)ψ′ bound state, Phys. Lett. B 665, 26 (2008)

[650]

F. K. Guo, J. Haidenbauer, C. Hanhart, and U. G. Meissner, Reconciling the X(4630) with the Y(4660), Phys. Rev. D 82, 094008 (2010)

[651]

T. Mehen and J. W. Powell, Heavy quark symmetry predictions for weakly bound B-meson molecules, Phys. Rev. D 84, 114013 (2011)

[652]

Q. Wang, V. Baru, A. A. Filin, C. Hanhart, A. V. Nefediev, and J. L. Wynen, The line shapes of the Zb(10610) and Zb(10650) in the elastic and inelastic channels revisited, Phys. Rev. D 98, 074023 (2018)

[653]

S. Ohkoda, Y. Yamaguchi, S. Yasui, K. Sudoh, and A. Hosaka, Exotic mesons with hidden bottom near thresholds, Phys. Rev. D 86, 014004 (2012)

[654]

V. Baru, E. Epelbaum, A. A. Filin, C. Hanhart, A. V. Nefediev, and Q. Wang, Spin partners WbJ from the line shapes of the Zb(10610) and Zb(10650), Phys. Rev. D 99, 094013 (2019)

[655]

M. Karliner and S. Nussinov, The doubly heavies: Q¯Qq¯q and QQq¯q¯ tetraquarks and QQq baryons, JHEP 07, 153 (2013)

[656]

X. W. Wang and Z. G. Wang, Search for the Charmed Baryonium and Dibaryon Structures via the QCD Sum Rules, Adv. High Energy Phys. 2022, 6224597 (2022)

[657]

B. D. Wan, L. Tang, and C. F. Qiao, Hidden-bottom and -charm hexaquark states in QCD sum rules, Eur. Phys. J. C 80, 121 (2020)

[658]

Z. G. Wang, Analysis of the Ξ¯ccΞcc hexaquark molecular state with the QCD sum rules, Phys. Lett. B 819, 136464 (2021)

[659]

Z. Y. Di and Z. G. Wang, Analysis of the DD¯∗K system with QCD sum rules, Adv. High Energy Phys. 2019, 8958079 (2019)

[660]

M. L. Du, V. Baru, X. K. Dong, A. Filin, F. K. Guo, C. Hanhart, A. Nefediev, J. Nieves, and Q. Wang, Coupled-channel approach to Tcc+ including three-body effects, Phys. Rev. D 105, 014024 (2022)

[661]

A. Feijoo, W. H. Liang, and E. Oset, D0D0π+ mass distribution in the production of the Tcc exotic state, Phys. Rev. D 104, 114015 (2021)

[662]

M. Albaladejo, Tcc+ coupled channel analysis and predictions, Phys. Lett. B 829, 137052 (2022)

[663]

X. Z. Ling, M. Z. Liu, L. S. Geng, E. Wang, and J. J. Xie, Can we understand the decay width of the Tcc+ state?, Phys. Lett. B 826, 136897 (2022)

[664]

L. Meng, G. J. Wang, B. Wang, and S. L. Zhu, Probing the long-range structure of the Tcc+ with the strong and electromagnetic decays, Phys. Rev. D 104, 051502 (2021)

[665]

C. R. Deng and S. L. Zhu, Tcc+ and its partners, Phys. Rev. D 105, 054015 (2022)

[666]

S. Fleming, R. Hodges, and T. Mehen, Tcc+ decays: Differential spectra and two-body final states, Phys. Rev. D 104, 116010 (2021)

[667]

M. J. Yan and M. P. Valderrama, Subleading contributions to the decay width of the Tcc+ tetraquark, Phys. Rev. D 105, 014007 (2022)

[668]

R. Chen, Q. Huang, X. Liu, and S. L. Zhu, Predicting another doubly charmed molecular resonance Tcc′+(3876), Phys. Rev. D 104, 114042 (2021)

[669]

H. Ren, F. Wu, and R. L. Zhu, Hadronic molecule interpretation of Tcc+ and its beauty partners, Adv. High Energy Phys. 2022, 9103031 (2022)

[670]

Z. G. Wang, Does vacuum saturation work for the higher-dimensional vacuum condensates in the QCD sum rules?, Int. J. Mod. Phys. A 36, 2150246 (2021)

[671]

Z. G. Wang, Triply-charmed hexaquark states with the QCD sum rules, Int. J. Mod. Phys. A 35, 2050073 (2020)

[672]

N. Kodama, M. Oka, and T. Hatsuda, H dibaryon in the QCD sum rule, Nucl. Phys. A 580, 445 (1994)

[673]

Z. G. Wang and S. L. Wan, X(1835) as a baryonium state with QCD sum rules, J. Phys. G 34, 505 (2007)

[674]

T. Iritani, . dibaryon from lattice QCD near the physical point, Phys. Lett. B 792, 284 (2019)

[675]

S. Gongyo, . Most strange dibaryon from lattice QCD, Phys. Rev. Lett. 120, 212001 (2018)

[676]

P. Junnarkar and N. Mathur, Deuteronlike Heavy Dibaryons from Lattice Quantum Chromodynamics, Phys. Rev. Lett. 123, 162003 (2019)

[677]

Y. Lyu, . Dibaryon with highest charm number near unitarity from lattice QCD, Phys. Rev. Lett. 127, 072003 (2021)

[678]

N. Mathur, M. Padmanath, and D. Chakraborty, Strongly bound dibaryon with maximal beauty flavor from lattice QCD, Phys. Rev. Lett. 130, 111901 (2023)

[679]

P. M. Junnarkar and N. Mathur, Spectrum of two-flavored spin-zero heavy dibaryons in lattice QCD, Phys. Rev. D 111, 014512 (2025)

[680]

Y. W. Pan, M. Z. Liu, F. Z. Peng, M. S. Sanchez, L. S. Geng, and M. P. Valderrama, Model independent determination of the spins of the Pc(4440) and Pc(4457) from the spectroscopy of the triply charmed dibaryons, Phys. Rev. D 102, 011504 (2020)

[681]

Y. Chung, H. G. Dosch, M. Kremer, and D. Schall, Baryon sum rules and chiral symmetry breaking, Nucl. Phys. B 197, 55 (1982)

[682]

E. Bagan, M. Chabab, H. G. Dosch, and S. Narison, Baryon sum rules in the heavy quark effective theory, Phys. Lett. B 301, 243 (1993)

[683]

D. Jido, N. Kodama, and M. Oka, Negative parity nucleon resonance in the QCD sum rule, Phys. Rev. D 54, 4532 (1996)

[684]

Z. G. Wang, Reanalysis of the heavy baryon states Ωb, Ωc, Ξb′, Ξc′, Σb and Σc with QCD sum rules, Phys. Lett. B 685, 59 (2010)

[685]

Z. G. Wang, Analysis of the 12+ doubly heavy baryon states with QCD sum rules, Eur. Phys. J. A 45, 267 (2010)

[686]

Z. G. Wang, Analysis of the 32+ heavy and doubly heavy baryon states with QCD sum rules, Eur. Phys. J. C 68, 459 (2010)

[687]

Z. G. Wang, Analysis of the 12± antitriplet heavy baryon states with QCD sum rules, Eur. Phys. J. C 68, 479 (2010)

[688]

Z. G. Wang, Analysis of the 12− and 32− heavy and doubly heavy baryon states with QCD sum rules, Eur. Phys. J. A 47, 81 (2011)

[689]

Z. G. Wang, Analysis of the triply heavy baryon states with QCD sum rules, Commun. Theor. Phys. 58, 723 (2012)

[690]

S. L. Zhu, Understanding pentaquark states in QCD, Phys. Rev. Lett. 91, 232002 (2003)

[691]

J. Sugiyama, T. Doi, and M. Oka, Pentaquark baryon from the QCD sum rule, Phys. Lett. B 581, 167 (2004)

[692]

R. D. Matheus, F. S. Navarra, M. Nielsen, R. Rodrigues da Silva, and S. H. Lee, Are Θ+ and the Roper resonance diquark-diquark-antiquark states?, Phys. Lett. B 578, 323 (2004)

[693]

L. Maiani, A. D. Polosa, and V. Riquer, The new pentaquarks in the diquark model, Phys. Lett. B 749, 289 (2015)

[694]

A. Ali and A. Y. Parkhomenko, Interpretation of the narrow J/ψp peaks in Λb→J/ψpK− decay in the compact diquark model, Phys. Lett. B 793, 365 (2019)

[695]

R. F. Lebed, The pentaquark candidates in the dynamical diquark picture, Phys. Lett. B 749, 454 (2015)

[696]

R. L. Zhu and C. F. Qiao, Pentaquark states in a diquark−triquark model, Phys. Lett. B 756, 259 (2016)

[697]

A. Ali, I. Ahmed, M. J. Aslam, and A. Rehman, Heavy quark symmetry and weak decays of the b-baryons in pentaquarks with a cc¯ component, Phys. Rev. D 94, 054001 (2016)

[698]

L. Maiani, A. D. Polosa, and V. Riquer, From pentaquarks to dibaryons in Λb(5620) decays, Phys. Lett. B 750, 37 (2015)

[699]

Z. G. Wang, Analysis of the Λc(2860), Λc(2880), Ξc(3055) and Ξc(3080) as D-wave baryon states with QCD sum rules, Nucl. Phys. B 926, 467 (2018)

[700]

Q. Xin, Z. G. Wang, and F. Lu, The Λ-type P-wave bottom baryon states via the QCD sum rules, Chin. Phys. C 47, 093106 (2023)

[701]

G. L. Yu, Z. G. Wang, and X. W. Wang, The 1D, 2D Ξb and Λb baryons, Chin. Phys. C 46, 093102 (2022)

[702]

Shi-Zhong Huang, Free Particles and Fields of High Spins, Anhui Peoples Publishing House, 2006 (in Chinese)

[703]

Z. G. Wang and T. Huang, Analysis of the 12± pentaquark states in the diquark model with QCD sum rules, Eur. Phys. J. C 76, 43 (2016)

[704]

Z. G. Wang, Analysis of the 12± pentaquark states in the diquark-diquark-antiquark model with QCD sum rules, Eur. Phys. J. C 76, 142 (2016)

[705]

Z. G. Wang, Analysis of the 32± pentaquark states in the diquark−diquark−antiquark model with QCD sum rules, Nucl. Phys. B 913, 163 (2016)

[706]

J. X. Zhang, Z. G. Wang, and Z. Y. Di, Analysis of the 32± pentaquark states in the diquark model with QCD sum rules, Acta Phys. Polon. B 48, 2013 (2017)

[707]

G. N. Li, X. G. He, and M. He, Some Predictions of Diquark Model for Hidden Charm Pentaquark Discovered at the LHCb, JHEP 12, 128 (2015)

[708]

H. Y. Cheng and C. K. Chua, Bottom Baryon Decays to Pseudoscalar Meson and Pentaquark, Phys. Rev. D 92, 096009 (2015)

[709]

Z. G. Wang, Analysis of the triply-charmed pentaquark states with QCD sum rules, Eur. Phys. J. C 78, 300 (2018)

[710]

H. X. Chen, W. Chen, X. Liu, T. G. Steele, and S. L. Zhu, Towards exotic hidden-charm pentaquarks in QCD, Phys. Rev. Lett. 115, 172001 (2015)

[711]

H. X. Chen, E. L. Cui, W. Chen, T. G. Steele, X. Liu, and S. L. Zhu, QCD sum rule study of hidden-charm pentaquarks, Eur. Phys. J. C 76, 572 (2016)

[712]

K. Azizi, Y. Sarac, and H. Sundu, Analysis of Pc+(4380) and Pc+(4450) as pentaquark states in the molecular picture with QCD sum rules, Phys. Rev. D 95, 094016 (2017)

[713]

K. Azizi, Y. Sarac, and H. Sundu, Strong Decay of Pc(4380) Pentaquark in a Molecular Picture, Phys. Lett. B 782, 694 (2018)

[714]

H. X. Chen, W. Chen, and S. L. Zhu, Possible interpretations of the Pc(4312), Pc(4440), and Pc(4457), Phys. Rev. D 100, 051501 (2019)

[715]

J. R. Zhang, Exploring a ΣcD¯ state: with focus on Pc(4312)+, Eur. Phys. J. C 79, 1001 (2019)

[716]

H. X. Chen, W. Chen, X. Liu, and X. H. Liu, Establishing the first hidden-charm pentaquark with strangeness, Eur. Phys. J. C 81, 409 (2021)

[717]

K. Azizi, Y. Sarac, and H. Sundu, Properties of the Pc(4312) pentaquark and its bottom partner, Chin. Phys. C 45, 053103 (2021)

[718]

K. Azizi, Y. Sarac, and H. Sundu, Hidden bottom pentaquark states with spin 32 and 52, Phys. Rev. D 96, 094030 (2017)

[719]

J. B. Xiang, H. X. Chen, W. Chen, X. B. Li, X. Q. Yao, and S. L. Zhu, Revisiting hidden-charm pentaquarks from QCD sum rules, Chin. Phys. C 43, 034104 (2019)

[720]

X. W. Wang, Z. G. Wang, G. L. Yu, and Q. Xin, Isospin eigenstates of the color singlet-singlet-type pentaquark states, Sci. China Phys. Mech. Astron. 65, 291011 (2022)

[721]

X. W. Wang and Z. G. Wang, Study of isospin eigenstates of the pentaquark molecular states with strangeness, Int. J. Mod. Phys. A 37, 2250189 (2022)

[722]

X. W. Wang and Z. G. Wang, Analysis of Pcs(4338) and related pentaquark molecular states via QCD sum rules, Chin. Phys. C 47, 013109 (2023)

[723]

U. Ozdem, Investigation of magnetic moment of Pcs(4338) and Pcs(4459) pentaquark states, Phys. Lett. B 836, 137635 (2023)

[724]

U. Ozdem, Electromagnetic properties of D¯(∗)Ξc′, D¯(∗)Λc, D¯s(∗)Λc and D¯s(∗)Ξc pentaquarks, Phys. Lett. B 846, 138267 (2023)

[725]

U. Ozdem, Analysis of the isospin eigenstate D¯Σc, D¯∗Σc, and D¯Σc∗ pentaquarks by their electromagnetic properties, Eur. Phys. J. C 84, 769 (2024)

[726]

R. Chen, X. Liu, X. Q. Li, and S. L. Zhu, Identifying exotic hidden-charm pentaquarks, Phys. Rev. Lett. 115, 132002 (2015)

[727]

F. K. Guo, U. G. Meissner, W. Wang, and Z. Yang, How to reveal the exotic nature of the Pc(4450), Phys. Rev. D 92, 071502 (2015)

[728]

R. Chen, Z. F. Sun, X. Liu, and S. L. Zhu, Strong LHCb evidence supporting the existence of the hidden-charm molecular pentaquarks, Phys. Rev. D 100, 011502 (2019)

[729]

C. J. Xiao, Y. Huang, Y. B. Dong, L. S. Geng, and D. Y. Chen, Exploring the molecular scenario of Pc(4312), Pc(4440), and Pc(4457), Phys. Rev. D 100, 014022 (2019)

[730]

T. J. Burns, Phenomenology of Pc(4380)+, Pc(4450)+ and related states, Eur. Phys. J. A 51, 152 (2015)

[731]

F. K. Guo, H. J. Jing, U. G. Meissner, and S. Sakai, Isospin breaking decays as a diagnosis of the hadronic molecular structure of the Pc(4457), Phys. Rev. D 99, 091501 (2019)

[732]

H. X. Chen, L. S. Geng, We. H. Liang, E. Oset, E. Wang, and J. J. Xie, Looking for a hidden-charm pentaquark state with strangeness S=−1 from Ξb− decay into J/ψK−Λ, Phys. Rev. C 93, 065203 (2016)

[733]

T. J. Burns and E. S. Swanson, Molecular interpretation of the Pc(4440) and Pc(4457) states, Phys. Rev. D 100, 114033 (2019)

[734]

Y. Yamaguchi, H. Garcia-Tecocoatzi, A. Giachino, A. Hosaka, E. Santopinto, S. Takeuchi, and M. Takizawa, Pc pentaquarks with chiral tensor and quark dynamics, Phys. Rev. D 101, 091502 (2020)

[735]

Y. Shimizu, D Suenaga, and M. Harada, Coupled channel analysis of molecule picture of Pc(4380), Phys. Rev. D 93, 114003 (2016)

[736]

Y. Yamaguchi and E. Santopinto, Hidden-charm pentaquarks as a meson-baryon molecule with coupled channels for D¯(∗)Λc and D¯(∗)Σc(∗), Phys. Rev. D 96, 014018 (2017)

[737]

M. L. Du, V. Baru, F. K. Guo, C. Hanhart, U. G. Meissner, J. A. Oller, and Q. Wang, Interpretation of the LHCb Pc States as Hadronic Molecules and Hints of a Narrow Pc(4380), Phys. Rev. Lett. 124, 072001 (2020)

[738]

M. L. Du, V. Baru, F. K. Guo, C. Hanhart, U. G. Meissner, J. A. Oller, and Q. Wang, Revisiting the nature of the Pc pentaquarks, JHEP 08, 157 (2021)

[739]

A. Pimikov, H. J. Lee, and P. M. Zhang, Hidden charm pentaquarks with color-octet substructure in QCD Sum Rules, Phys. Rev. D 101, 014002 (2020)

[740]

Z. G. Wang, Light tetraquark state candidates, Adv. High Energy Phys. 2020, 6438730 (2020)

[741]

X. S. Yang, Q. Xin, and Z. G. Wang, Analysis of the Tcs¯(2900) and related tetraquark states with the QCD sum rules, Int. J. Mod. Phys. A 38, 2350056 (2023)

[742]

G. J. Wang, L. Meng, L. Y. Xiao, M. Oka, and S. L. Zhu, Mass spectrum and strong decays of tetraquark c¯s¯qq states, Eur. Phys. J. C 81, 188 (2021)

[743]

Y. Huang, J. X. Lu, J. J. Xie, and L. S. Geng, Strong decays of D¯∗K∗ molecules and the newly observed X0,1 states, Eur. Phys. J. C 80, 973 (2020)

[744]

C. J. Xiao, D. Y. Chen, Y. B. Dong, and G. W. Meng, A study of the decays of S-wave D¯∗K∗ hadronic molecules: The scalar X0(2900) and its spin partners XJ(J=1,2), Phys. Rev. D 103, 034004 (2021)

[745]

B. Wang and S. L. Zhu, How to understand the X(2900)?, Eur. Phys. J. C 82, 419 (2022)

[746]

T. J. Burns and E. S. Swanson, Kinematical cusp and resonance interpretations of the X(2900), Phys. Lett. B 813, 136057 (2021)

[747]

T. J. Burns and E. S. Swanson, Discriminating among interpretations for the X(2900) states, Phys. Rev. D 103, 014004 (2021)

[748]

M. E. Bracco, A. Lozea, R. D. Matheus, F. S. Navarra, and M. Nielsen, Disentangling two- and four-quark state pictures of the charmed scalar mesons, Phys. Lett. B 624, 217 (2005)

[749]

H. Kim and Y. Oh, Ds(2317) as a four-quark state in QCD sum rules, Phys. Rev. D 72, 074012 (2005)

[750]

Z. G. Wang and S. L. Wan, Ds0(2317) as a tetraquark state with QCD sum rules in heavy quark limit, Nucl. Phys. A 778, 22 (2006)

[751]

T. J. Burns and E. S. Swanson, Interpreting the X(5568), Phys. Lett. B 760, 627 (2016)

[752]

F. K. Guo, U. G. Meissner, and B. S. Zou, How the X(5568) challenges our understanding of QCD, Commun. Theor. Phys. 65, 593 (2016)

[753]

R. Aaij, . Observation of five new narrow Ωc0 states decaying to Ξc+K−, Phys. Rev. Lett. 118, 182001 (2017)

[754]

J. Yelton, . Observation of excited Ωc charmed baryons in e+e− collisions, Phys. Rev. D 97, 051102 (2018)

[755]

R. Aaij, . Observation of new Ωc0 states decaying to the Ξc+K− final state, Phys. Rev. Lett. 131, 131902 (2023)

[756]

R. Mizuk, . Observation of an isotriplet of excited charmed baryons decaying to Λc+π, Phys. Rev. Lett. 94, 122002 (2005)

[757]

B. Aubert, . Measurements of B(B¯0→Λc+p¯) and B(B−→Λc+p¯π−) and studies of Λc+π− resonances, Phys. Rev. D 78, 112003 (2008)

[758]

B. Aubert, . Observation of a charmed baryon decaying to D0p at a mass near 2.94GeV/c2, Phys. Rev. Lett. 98, 012001 (2007)

[759]

K. Abe, . Experimental constraints on the spin and parity of the Λc(2880), Phys. Rev. Lett. 98, 262001 (2007)

[760]

R. Aaij, . Study of the D0p amplitude in Λb0→D0pπ− decays, JHEP 05, 030 (2017)

[761]

Y. B. Li, . Evidence of a new excited charmed baryon decaying to Σc(2455)0,++π±, Phys. Rev. Lett. 130, 031901 (2023)

[762]

S. Sakai, F. K. Guo, and B. Kubis, Extraction of ND scattering lengths from the Λb→π−pD0 decay and properties of the Σc(2800)+, Phys. Lett. B 808, 135623 (2020)

[763]

L. F. Zhao, H. X. Huang, and J. L. Ping, ND and NB systems in quark delocalization color screening model, Eur. Phys. J. A 53, 28 (2017)

[764]

Z. Y. Wang, J. J. Qi, X. H. Guo, and K. W. Wei, Study of molecular ND bound states in the Bethe-Salpeter equation approach, Phys. Rev. D 97, 094025 (2018)

[765]

M. J. Yan, F. Z. Peng, and M. P. Valderrama, Molecular charmed baryons and pentaquarks from light-meson exchange saturation, Phys. Rev. D 109, 014023 (2024)

[766]

Y. Yan, X. H. Hu, Y. H. Wu, H. X. Huang, and J. L. Ping, Molecular state interpretation of charmed baryons in the quark model, Eur. Phys. J. C 83, 524 (2023)

[767]

J. R. Zhang, S-wave D(∗)N molecular states: Σc(2800) and Λc(2940)+?, Phys. Rev. D 89, 096006 (2014)

[768]

Z. L. Zhang, Z. W. Liu, S. Q. Luo, F. L. Wang, B. Wang, and H. Xu, Λc(2910) and Λc(2940) as the conventional baryons dressed with the D∗N channel, Phys. Rev. D 107, 034036 (2023)

[769]

B. Wang, L. Meng, and S. L. Zhu, D(∗)N interaction and the structure of Σc(2800) and Λc(2940) in chiral effective field theory, Phys. Rev. D 101, 094035 (2020)

[770]

Y. Huang, J. He, J. J. Xie, and L. S. Geng, Production of the Λc(2940) by kaon-induced reactions on a proton target, Phys. Rev. D 99, 014045 (2019)

[771]

Q. Xin, X. S. Yang, and Z. G. Wang, The singly charmed pentaquark molecular states via the QCD sum rules, Int. J. Mod. Phys. A 38, 2350123 (2023)

[772]

K. L. Wang and X. H. Zhong, Toward establishing the low-lying P-wave excited Σc baryon states, Chin. Phys. C 46, 023103 (2022)

[773]

K. L. Wang, Y. X. Yao, X. H. Zhong, and Q. Zhao, Strong and radiative decays of the low-lying S- and P-wave singly heavy baryons, Phys. Rev. D 96, 116016 (2017)

[774]

H. M. Yang and H. X. Chen, P-wave charmed baryons of the SU(3) flavor 6F, Phys. Rev. D 104, 034037 (2021)

[775]

D. J Jia, W. N. Liu, and A. Hosaka, Regge behaviors in orbitally excited spectroscopy of charmed and bottom baryons, Phys. Rev. D 101, 034016 (2020)

[776]

G. L. Yu, Z. Y. Li, Z. G. Wang, J. Lu, and M. Yan, Systematic analysis of single heavy baryons ΛQ, ΣQ and ΩQ, Nucl. Phys. B 990, 116183 (2023)

[777]

H. Y. Cheng and C. W. Chiang, Quantum numbers of Ωc states and other charmed baryons, Phys. Rev. D 95, 094018 (2017)

[778]

S. Q. Luo, B. Chen, Z. W. Liu, and X. Liu, Resolving the low mass puzzle of Λc(2940)+, Eur. Phys. J. C 80, 301 (2020)

[779]

Q. F. Lu, L. Y. Xiao, Z. Y. Wang, and X. H. Zhong, Strong decay of Λc(2940) as a 2P state in the Λc family, Eur. Phys. J. C 78, 599 (2018)

[780]

K. Azizi, Y. Sarac, and H. Sundu, Interpretation of the Λc(2910)+ baryon newly seen by Belle Collaboration and its possible bottom partner, Eur. Phys. J. C 82, 920 (2022)

[781]

W. J. Wang, L. Y. Xiao, and X. H. Zhong, The strong decays of the low-lying ρ-mode 1P-wave singly heavy baryons, Phys. Rev. D 106, 074020 (2022)

[782]

G. L. Yu, M. Yan, Z. Y. Li, Z. G. Wang, and J. Lu, Strong decay properties of single heavy baryons ΛQ, ΣQ and ΩQ, Int. J. Mod. Phys. A 38, 2350082 (2023)

[783]

M. Karliner and J. L. Rosner, Excited Ωc baryons as 2S states, Phys. Rev. D 108, 014006 (2023)

[784]

S. Q. Luo and X. Liu, The newly observed Ωc(3327): A good candidate for a D-wave charmed baryon, Phys. Rev. D 107, 074041 (2023)

[785]

Z. G. Wang, F. Lu, and Y. Liu, Analysis of the D-wave Σ-type charmed baryon states with the QCD sum rules, Eur. Phys. J. C 83, 689 (2023)

[786]

H. J. Wang, Z. Y. Di, and Z. G. Wang, Analysis of the Ξb(6227) as the 12± Pentaquark Molecular States with QCD Sum Rules, Int. J. Theor. Phys. 59, 3124 (2020)

[787]

H. J. Wang, Z. Y. Di, and Z. G. Wang, Analysis of the excited Ωc states as the 12± pentaquark states with QCD sum rules, Commun. Theor. Phys. 73, 035201 (2021)

[788]

Z. G. Wang and J. X. Zhang, Possible pentaquark candidates: new excited Ωc states, Eur. Phys. J. C 78, 503 (2018)

[789]

Z. G. Wang, Strong decays of the Y(4660) as a vector tetraquark state in solid quark-hadron duality, Eur. Phys. J. C 79, 184 (2019)

[790]

F. S. Navarra and M. Nielsen, X(3872)→J/ψπ+π− and X(3872)→J/ψπ+π−π0 decay widths from QCD sum rules, Phys. Lett. B 639, 272 (2006)

[791]

J. M. Dias, K. P. Khemchandani, A. M. Torres, M. Nielsen, and C. M. Zanetti, A QCD sum rule calculation of the X±(5568)→Bs0π± decay width, Phys. Lett. B 758, 235 (2016)

[792]

W. Chen, T. G. Steele, H. X. Chen, and S. L. Zhu, Zc(4200)+ decay width as a charmonium-like tetraquark state, Eur. Phys. J. C 75, 358 (2015)

[793]

K. Azizi, Y. Sarac, and H. Sundu, Investigation of Pcs(4459)0 pentaquark via its strong decay to ΛJ/ψ, Phys. Rev. D 103, 094033 (2021)

[794]

K. Azizi, Y. Sarac, and H. Sundu, Investigation of a candidate spin-12 hidden-charm triple strange pentaquark state Pcsss, Phys. Rev. D 107, 014023 (2023)

[795]

K. Azizi, Y. Sarac, and H. Sundu, Investigation of hidden-charm double strange pentaquark candidate Pcss via its mass and strong decays, Eur. Phys. J. C 82, 543 (2022)

[796]

K. Azizi, Y. Sarac, and H. Sundu, Investigation of the strange pentaquark candidate PψsΛ(4338)0 recently observed by LHCb, Phys. Rev. D 108, 074010 (2023)

[797]

Z. G. Wang, Analysis of strong decays of the Zc(4600) with the QCD sum rules, Int. J. Mod. Phys. A 34, 1950110 (2019)

[798]

Z. G. Wang and X. Wang, Analysis of the strong decays of the Pc(4312) as a pentaquark molecular state with QCD sum rules, Chin. Phys. C 44, 103102 (2020)

[799]

Z. G. Wang, H. J. Wang, and Q. Xin, The hadronic coupling constants of the lowest hidden-charm pentaquark state with the QCD sum rules in rigorous quark-hadron duality, Chin. Phys. C 45, 063104 (2021)

[800]

Y. J. Xu, Y. L. Liu, and M. Q. Huang, The magnetic moment of Pc(4312) as a D¯Σc molecular state, Eur. Phys. J. C 81, 421 (2021)

[801]

Z. G. Wang, Decay widths of Zcs(3985/4000) based on rigorous quark-hadron duality, Chin. Phys. C 46, 103106 (2022)

[802]

Z. G. Wang and X. S. Yang, The two-body strong decays of the fully-charm tetraquark states, AAPPS Bull. 34, 5 (2024)

[803]

X. W. Wang and Z. G. Wang, Strong decays of the Pc(4312) and its isospin cousin via the QCD sum rules, Chin. Phys. C 48, 053102 (2024)

[804]

X. W. Wang and Z. G. Wang, Strong decays of Pcs(4338) and its high isospin cousin via QCD sum rules, Phys. Rev. D 110, 014008 (2024)

[805]

Y. J. Xu, C. Y. Cui, Y. L. Liu, and M. Q. Huang, Partial decay widths of Pc(4312) as a D¯Σc molecular state, Phys. Rev. D 102, 034028 (2020)

[806]

Z. G. Wang, The magnetic moment of the Zc(3900) as an axialvector tetraquark state with QCD sum rules, Eur. Phys. J. C 78, 297 (2018)

[807]

S. S. Agaev, K. Azizi, and H. Sundu, Exploring the resonances X(4140) and X(4274) through their decay channels, Phys. Rev. D 95, 114003 (2017)

[808]

S. S. Agaev, K. Azizi, and H. Sundu, Spectroscopic parameters and decays of the resonance Zb(10610), Eur. Phys. J. C 77, 836 (2017)

[809]

H. Sundu, S. S. Agaev, and K. Azizi, Resonance Y(4660) as a vector tetraquark and its strong decay channels, Phys. Rev. D 98, 054021 (2018)

[810]

Y. Xie and H. Sun, Investigate the strong coupling of gXJ/ψϕ in X(4500)→J/ψϕ by using the three-point sum rules and the light-cone sum rules, Nucl. Phys. B 1000, 116471 (2024)

[811]

U. Ozdem and K. Azizi, Electromagnetic multipole moments of the Pc+(4380) pentaquark in light-cone QCD, Eur. Phys. J. C 78, 379 (2018)

[812]

B. L. Ioffe and A. V. Smilga, Nucleon magnetic moments and magnetic properties of vacuum in QCD, Nucl. Phys. B 232, 109 (1984)

[813]

V. M. Belyaev, V. M. Braun, A. Khodjamirian, and R. Ruckl, D∗Dπ and B∗Bπ couplings in QCD, Phys. Rev. D 51, 6177 (1995)

[814]

P. Ball, Theoretical update of pseudoscalar meson distribution amplitudes of higher twist: The Nonsinglet case, JHEP 01, 010 (1999)

[815]

V. M. Braun and I. E. Filyanov, Conformal invariance and pion wave functions of nonleading twist, Z. Phys. C 48, 239 (1990)

[816]

Z. G. Wang, Three-body strong decays of the Y(4230) via the light-cone QCD sum rules, Int. J. Mod. Phys. A 38, 2350175 (2023)

RIGHTS & PERMISSIONS

Higher Education Press

AI Summary AI Mindmap
PDF (12138KB)

957

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/