Skip to main content
Log in

Mass Spectrum of Pseudo-Scalar Glueballs from a Bethe–Salpeter Approach with the Rainbow–Ladder Truncation

  • Published:
Few-Body Systems Aims and scope Submit manuscript

Abstract

We suggest a framework based on the rainbow approximation to the Dyson–Schwinger and Bethe–Salpeter equations with effective parameters adjusted to lattice QCD data to calculate the masses of the ground and excited states of pseudo-scalar glueballs. The structure of the truncated Bethe–Salpeter equation with the gluon and ghost propagators as solutions of the truncated Dyson–Schwinger equations is analyzed in Landau gauge. Both, the Bethe–Salpeter and Dyson–Schwinger equations, are solved numerically within the same rainbow–ladder truncation with the same effective parameters which ensure consistency of the approach. We found that with a set of parameters, which provides a good description of the lattice data within the Dyson–Schwinger approach, the solutions of the Bethe–Salpeter equation for the pseudo-scalar glueballs exhibit a rich mass spectrum which also includes the ground and excited states predicted by lattice calculations. The obtained mass spectrum contains also several intermediate excitations beyond the lattice approaches. The partial Bethe–Salpeter amplitudes of the pseudo-scalar glueballs are presented as well.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. In the present paper we use the four-dimensional Levi–Civita symbol \(\epsilon ^{\mu \nu \alpha \beta }\) normalized according to the Itzykson-Zuber convention (see Appendix A-10 in Ref. [57]), i.e. \(\epsilon ^{0123}=1\), \(\epsilon ^{\mu \nu \alpha \beta } \epsilon {_{\mu \nu }}^{\alpha '\beta '} =2 (g^{\alpha \beta '}g^{\beta \alpha '}-g^{\alpha \alpha '}g^{\beta \beta '}) \), \(\epsilon ^{\mu \nu \alpha \beta } \epsilon _{\mu \nu \alpha \beta }=-24\).

  2. It should be noted that the most recent publication [22] does not longer consider the excited states, so that the state \(M_{gg}^{(\mathrm{1st.})}=3640\pm 189\) MeV should be considered with some caution.

References

  1. H. Fritzsch, P. Minkowski, Nuovo Cim. A 30, 393 (1975)

    ADS  Google Scholar 

  2. R.L. Jaffe, K. Johnson, Phys. Lett. B 60, 201 (1976)

    ADS  Google Scholar 

  3. U. Wiedner, Prog. Part. Nucl. Phys. 66, 477 (2011)

    ADS  Google Scholar 

  4. S. Jia et al., Belle Collaboration. Phys. Rev. D 95, 012001 (2017)

    ADS  Google Scholar 

  5. D. Robson, Z. Phys, Z. Phys. C 3, 199 (1980)

    ADS  Google Scholar 

  6. N. Isgur, J.E. Paton, Phys. Rev. D 31, 2910 (1985)

    ADS  Google Scholar 

  7. C.E. Carlson, T.H. Hansson, C. Peterson, Phys. Rev. D 30, 1594 (1984)

    ADS  Google Scholar 

  8. M.S. Chanowitz, S.R. Sharpe, Nucl. Phys. B 222, 211 (1983). [Erratum Nucl. Phys. B 228, 588 (1983)]

    ADS  Google Scholar 

  9. J.M. Cornwall, A. Soni, Phys. Lett. B 120, 431 (1983)

    ADS  Google Scholar 

  10. Y.M. Cho, X.Y. Pham, P. Zhang, J.J. Xie, L.P. Zou, Phys. Rev. D 91, 114020 (2015)

    ADS  Google Scholar 

  11. N. Boulanger, F. Buisseret, V. Mathieu, C. Semay, Eur. Phys. J. A 38, 317 (2008)

    ADS  Google Scholar 

  12. J. Leutgeb, A. Rebhan, Phys. Rev. D 101, 014006 (2020)

    ADS  Google Scholar 

  13. L. Bellantuono, P. Colangelo, F. Giannuzzi, JHEP 10, 137 (2015)

    ADS  Google Scholar 

  14. Y. Chen, M. Huang, Chin. Phys. C 40, 123101 (2016)

    ADS  Google Scholar 

  15. F. Brunner, A. Rebhan, Phys. Lett. B 770, 124 (2017)

    ADS  Google Scholar 

  16. M.A. Shifman, A.I. Vainshtein, V.I. Zakharov, Nucl. Phys. B 147, 385 (1979)

    ADS  Google Scholar 

  17. E.V. Shuryak, Nucl. Phys. B 203, 116 (1982)

    ADS  Google Scholar 

  18. A. Pimikov, H.J. Lee, N. Kochelev, P. Zhang, V. Khandramai, Phys. Rev. D 96, 114024 (2017)

    ADS  Google Scholar 

  19. A. Pimikov, H.J. Lee, N. Kochelev, Phys. Rev. Lett. 119, 079101 (2017)

    ADS  Google Scholar 

  20. A. Pimikov, H.J. Lee, N. Kochelev, P. Zhang, Phys. Rev. D 95, 071501 (2017)

    ADS  Google Scholar 

  21. M. Albanese et al., Ape Collaboration. Phys. Lett. B 197, 400 (1987)

    ADS  Google Scholar 

  22. Y. Chen, A. Alexandru, S. Dong, T. Draper, I. Horvath et al., Phys. Rev. D 73, 014516 (2006)

    ADS  Google Scholar 

  23. C.J. Morningstar, M.J. Peardon, Phys. Rev. D 60, 034509 (1999)

    ADS  Google Scholar 

  24. G. Gabadadze, Phys. Rev. D 58, 055003 (1998)

    ADS  Google Scholar 

  25. W. Ochs, J. Phys. G 40, 043001 (2013)

    ADS  Google Scholar 

  26. H. Noshad, S.M. Zebarjad, S. Zarepour, Nucl. Phys. B 934, 408 (2018)

    ADS  Google Scholar 

  27. H. Sanchis-Alepuz, C.S. Fischer, C. Kellermann, L. von Smekal, Phys. Rev. D 92, 034001 (2015)

    ADS  Google Scholar 

  28. J. Meyers, E.S. Swanson, Phys. Rev. D 87, 036009 (2013)

    ADS  Google Scholar 

  29. E.V. Souza, M.N. Ferreira, A.C. Aguilar, J. Papavassiliou, C.D. Roberts, S.-S. Xu, Eur. Phys. J. A 56, 25 (2020)

    ADS  Google Scholar 

  30. P. Maris, C.D. Roberts, Phys. Rev. C 56, 3369 (1997)

    ADS  Google Scholar 

  31. S.M. Dorkin, T. Hilger, L.P. Kaptari, B. Kämpfer, Few Body Syst. 49, 247 (2011)

    ADS  Google Scholar 

  32. P. Maris, C.D. Roberts, Int. J. Mod. Phys. E 12, 297 (2003)

    ADS  Google Scholar 

  33. R. Alkofer, P. Watson, H. Weigel, Phys. Rev. D 65, 094026 (2002)

    ADS  Google Scholar 

  34. C.S. Fischer, P. Watson, W. Cassing, Phys. Rev. D 72, 094025 (2005)

    ADS  Google Scholar 

  35. M.R. Frank, C.D. Roberts, Phys. Rev. C 53, 390 (1996)

    ADS  Google Scholar 

  36. S.M. Dorkin, L.P. Kaptar, B. Kämpfer, Phys. Rev. C 91, 055201 (2015)

    ADS  Google Scholar 

  37. S.M. Dorkin, L.P. Kaptari, T. Hilger, B. Kampfer, Phys. Rev. C 89, 034005 (2014)

    ADS  Google Scholar 

  38. V.B. Berestetskii, E.V. Lifshitz, L.P. Pitaevskii, Quantum Electrodynamics (Pergamon Press, Oxford, 1982), p. 29

    Google Scholar 

  39. L.P. Kaptari, B. Kämpfer, P.-M. Zhang, Eur. Phys. J. Plus 134, 383 (2019)

    Google Scholar 

  40. A. Hauck, L. von Smekal, R. Alkofer, Comput. Phys. Commun. 112, 149 (1998)

    ADS  Google Scholar 

  41. S. Mandelstam, Phys. Rev. D 20, 3223 (1979)

    ADS  Google Scholar 

  42. K. Buttner, M.R. Pennington, Phys. Rev. D 52, 5220 (1995)

    ADS  Google Scholar 

  43. D. Atkinson, J.C.R. Bloch, Phys. Rev. D 58, 094036 (1998)

    ADS  Google Scholar 

  44. L. von Smekal, A. Hauck, R. Alkofer, Phys. Rev. Lett. 79, 3591 (1997)

    ADS  Google Scholar 

  45. C.S. Fischer, e-Print: hep-ph/0304233 (Univ. of Tübingen, PhD-thesis, Nov 2002)

  46. L. von Smekal, A. Hauck, R. Alkofer, Ann. Phys. 267, 1 (1998). [Erratum: Ann. Phys. 269, 282 (1998)]

    ADS  Google Scholar 

  47. C.S. Fischer, R. Alkofer, H. Reinhardt, Phys. Rev. D 65, 094008 (2002)

    ADS  Google Scholar 

  48. C.S. Fischer, A. Maas, J.M. Pawlowski, Ann. Phys. 324, 2408 (2009)

    ADS  Google Scholar 

  49. A.C. Aguilar, D. Binosi, J. Papavassiliou, Phys. Rev. D 78, 025010 (2008)

    ADS  Google Scholar 

  50. Ph Boucaud, J.P. Leroy, A. Le Yaouanc, J. Micheli, O. Pene, J. Rodriguez-Quintero, Few-Body Syst. 53, 387 (2012)

    ADS  Google Scholar 

  51. C.S. Fischer, J. Phys. G 32, R253 (2006)

    ADS  Google Scholar 

  52. P.O. Bowman, U.M. Heller, D.B. Leinweber, M.B. Parappilly, A. Sternbeck, L. von Smekal, A.G. Williams, J. Zhang, Phys. Rev. D 76, 094505 (2007)

    ADS  Google Scholar 

  53. A. Sternbeck, M. Müller-Preussker, Phys. Lett. B 276, 396 (2013)

    ADS  Google Scholar 

  54. M. Huber, arXiv:2003.13703 [hep-ph]

  55. V.G. Bornyakov, V.K. Mitrjushkin, M. Müller-Preussker, Phys. Rev. D 81, 054503 (2010)

    ADS  Google Scholar 

  56. V.G. Bornyakov, E.-M. Ilgenfritz, C. Litwinski, M. Müller-Preussker, V.K. Mitrjushkin, Phys. Rev. D 92, 074505 (2015)

    ADS  Google Scholar 

  57. C. Itzykson, J.-B. Zuber, Quantum Field Theory (McGraw-Hill, New York, 1980)

    MATH  Google Scholar 

  58. D. Zwanziger, Nucl. Phys. B 323, 513 (1989)

    ADS  Google Scholar 

  59. M. Stingl, Phys. Rev. D 34, 3863 (1986). [Erratum ibid. D 36, 651 (1987)]

    ADS  Google Scholar 

  60. A. Cucchieri, D. Dudal, T. Mendes, N. Vandersickel, Phys. Rev. D 85, 094513 (2012)

    ADS  Google Scholar 

  61. S. Strauss, C.S. Fischer, C. Kellermann, Prog. Part. Nucl. Phys. 67, 239 (2012)

    ADS  Google Scholar 

  62. S. Strauss, C.S. Fischer, C. Kellermann, Phys. Rev. Lett. 109, 252001 (2012)

    ADS  Google Scholar 

  63. P. Maris, Phys. Rev. D 52, 6087 (1995)

    ADS  MathSciNet  Google Scholar 

  64. J. Meyers, E.S. Swanson, Phys. Rev. D 87, 036009 (2013)

    ADS  Google Scholar 

  65. S.M. Dorkin, M. Beyer, S.S. Semikh, L.P. Kaptari, Few Body Syst. 42, 1 (2008)

    ADS  Google Scholar 

  66. S.M. Dorkin, L.P. Kaptari, C. Ciofi degli Atti, B. Kämpfer, Few Body Syst. 49, 233 (2011)

    ADS  Google Scholar 

  67. T. Hilger, M. Gomez-Rocha, A. Krassnigg, Eur. Phys. J. C 77, 625 (2017)

    ADS  Google Scholar 

  68. V.A. Karmanov, J. Carbonell, H. Sazdjian, EPJ Web Conf. 204, 01014 (2019)

    Google Scholar 

  69. V.A. Karmanov, J. Carbonell, H. Sazdjian, e-Print: arXiv:1903.02892

  70. V.A. Karmanov, J. Carbonell, H. Sazdjian, e-Print: arXiv:2001.00401

Download references

Acknowledgements

This work was supported in part by the Heisenberg—Landau program of the JINR—FRG collaboration, GSI-FE and BMBF. LPK appreciates the warm hospitality at the Helmholtz-Zentrum Dresden-Rossendorf.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to L. P. Kaptari.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Partial Decomposition of the Rainbow Kernel

The spatial dependence of the integrand on \(\varOmega _{{\mathbf {k}}}\) is contained in the rainbow exponents and in the scalar product \((p\cdot k)^\delta = ({\tilde{p}}\ {\tilde{k}} \ x_{kp})^\delta \),

$$\begin{aligned} \exp (\alpha x_{kp}) x_{kp}^\delta = \sum _{M_v} W_{M_v}^{(\delta )}({\tilde{p}},{\tilde{k}}) G_{M_v}^{(1)}(x_{kp}), \end{aligned}$$
(26)

where \(\alpha =2{\tilde{k}}{\tilde{p}}/\omega ^2\). The partial coefficients \(W_{M_v}^{(\delta )}({\tilde{p}},{\tilde{k}}) \) can be computed explicitly as

$$\begin{aligned} W_{M_v}^{(\delta )}({\tilde{p}},{\tilde{k}})= & {} \frac{2}{\pi }\int _{-1}^1 \sqrt{1-x_{kp}^2} \exp (\alpha x_{kp}) x_{kp}^\delta G_{M_v}^{(1)}(x_{kp})d x_{kp} \nonumber \\= & {} \frac{2}{\pi }\frac{d^\delta }{d\alpha ^\delta }\left[ \int _{-1}^1 \sqrt{1-x_{kp}^2} \mathrm{e}^{\alpha x_{kp}} G_{M_v}^{(1)}(x_{kp})\right] d x_{kp} = 2(M_v+1) \frac{d^\delta }{d\alpha ^\delta } \left[ \frac{1}{\alpha }I_{M_v+1}(\alpha )\right] , \end{aligned}$$
(27)

where \(I_{M_v+1}(\alpha )\) are the modified Bessel functions of second kind (of the imaginary argument) yielding

$$\begin{aligned} \mathrm{e}^{\alpha x_{kp}} x_{kp}^\delta =2 \sum _{M_v} (M_v+1) \frac{d^\delta }{d\alpha ^\delta }\left[ \frac{1}{\alpha }I_{M_v+1}(\alpha )\right] G_{M_v}^{(1)}(x_{kp}). \end{aligned}$$
(28)

The dependence on the spatial angles of the vectors p and k enters via \(G_{M_v}^{(1)}(x_{kp}\equiv \cos \xi _{kp})\), where \(\cos \xi _{kp}=\cos \xi _p\cos \xi _k + \sin \xi _p\sin \xi _k\cos \theta _{{\mathbf {kp}}}\). Explicitly, such a dependence can be written by using an addition theorem for Gegenbauer polynomials

$$\begin{aligned} G_{M_v}^{(1)}(x)=\frac{2\pi ^2}{M_v+1} \sum _{l\mu } Z_{M_v l\mu }^*(p) Z_{M_v l\mu }(k) \end{aligned}$$
(29)

with \( Z_{M_v l\mu }(k)=Z_{M_v l\mu }(\xi _k,\theta _{{\mathbf {k}}},\phi _{{\mathbf {k}}})\) as hyper-spherical harmonics, to obtain

$$\begin{aligned} \mathrm{e}^{\alpha x_{kp}} x_{kp}^\delta =4\pi ^2 \sum _{M_v,l,\mu } \frac{d^\delta }{d\alpha ^\delta } \left[ \frac{1}{\alpha }I_{M_v+1}(\alpha )\right] Z_{M_v l\mu }^*(p) Z_{M_v l\mu }(k), \end{aligned}$$
(30)

where the normalized hyperspherical harmonics are \(Z_{M_v l\mu }(p) = X_{M_v l}(\xi _p) \mathrm{Y}_{l\mu }({\mathbf {p}})\) with

$$\begin{aligned} X_{M_v l}(\xi _p) = 2^l l!\sqrt{\frac{2}{\pi }} \sqrt{\frac{(M_v+1)(M_v-l)!}{(M_v+l+1)!}} \sin ^l\xi _p G_{M_v-l}^{l+1}(\cos \xi _p). \end{aligned}$$
(31)

At a first glance, Eqs. (26)–(31) seemingly even complicate the integration. However, by observing that the dependence of the integrand in (6) on the spatial angles \(\varOmega _{{\mathbf {k}}}\) is only through the interaction kernel and trough \(x_{kp}^\delta \), Eq. (30), i.e. only trough the spatial harmonics \(\mathrm{Y}_{l\mu }({\mathbf {k}})\), the integration over \(d\varOmega _{{\mathbf {k}}}\) is trivial and eventually we have

$$\begin{aligned} \int \mathrm{e}^{\alpha x_{kp}} x_{kp}^\delta d\varOmega _{{\mathbf {k}}} = 8\pi \sum _{M_v}\frac{d^\delta }{d\alpha ^\delta } \left[ \frac{1}{\alpha }I_{M_v+1}(\alpha )\right] G_{M_v}^{(1)}(x_k) G_{M_v}^{(1)}(x_p). \end{aligned}$$
(32)

Appendix B: Integration over \(x_k\)

Here, we present some details of the integration over the hyper angle \(x_k\) and the resulting explicit expressions of selection rules. The corresponding angular integral is of the form

$$\begin{aligned} {{\mathcal {K}}}_{M_v,M_k}^{L}= \int \limits _{-1}^1\sqrt{1-x_k^2} x_k^L G_{M_k}^{(1)}(x)G_{M_v}^{(1)}(x_k) dx_k. \end{aligned}$$
(33)

Due to parity restrictions, the partial amplitudes \(F_{M_k}\) contain only even values of the Gegenbauer polynomials, i.e. \(M_k=[0,2,4,.. M^{max}]\), where \( M^{max}\) is the maximum number of polynomials taken into account in concrete calculations. The Gegenbauers \(G_{M_v}^{(1)}(x_k)\), which come from the interaction kernel (32), may contain both, even and odd values of \(M_v\), and formally the summation is extended to infinite, \(M_v=[0..\infty ]\). However, not all values in this interval contribute to (33). The symmetrical limits of integration restrict the Gegenbauer polynomials in (33) to obey the condition (\(L+M_k+M_v\))=even. Other restrictions originate from the explicit expression of the integral, see below. From a standard math handbook one infers that

$$\begin{aligned}&\int \limits _{-1}^1\sqrt{1-x^2} x^L G_{M_k}^{(\lambda )}(x)G_{M_v}^{(\lambda )}(x) dx = 2^{M_k+M_v}\frac{(2\lambda )_{M_k} (2\lambda )_{M_v} L!}{m_k!M_v!(L+M_k+M_v)!}\left( \frac{L-M_k-M_v}{2}+1\right) _{M_k+M_v} \nonumber \\&\times B\left( \lambda +\frac{1}{2},\frac{L+M_k+M_v+1}{2}\right) {}_3F_2\left( -M_k,-M_v,1;2\lambda ,\frac{L-M_k-M_v}{2}+1;1\right) , \end{aligned}$$
(34)

where \({}_3F_2\) is the generalized hypergeometric function and \((a)_k=a (a+1)(a+2)\ldots (a+k-1)\) (with \(a_0=1\)) is the known Pochhammer symbol and \(B(x,y)=\frac{\varGamma (x)\varGamma (y)}{\varGamma (x+1)}\) and \(\varGamma (x)\) are the familiar Euler \(\beta -\) and \(\varGamma -\) functions, respectively. Despite the integral (34) is finite, at some values of \(L,M_k\) and \(M_v\) the product of the Pochhammer symbol and hypergeometric function can be of the type \(0 \cdot \infty \), which implies that Eq. (34) cannot be implemented directly into numerical calculations. One needs to handle zeros and singularities manually. We use the obvious properties

$$\begin{aligned} (-m)_k=(-1)^k\frac{ m!}{(m-k)!}; \qquad (a)_k=\frac{\varGamma (a+k)}{\varGamma (a)} \end{aligned}$$
(35)

to obtain

$$\begin{aligned}&\left( \frac{L-M_k-M_v}{2}+1\right) _{M_k+M_v}=\frac{\kappa !}{\kappa _1!},\end{aligned}$$
(36)
$$\begin{aligned}&B\left( \frac{3}{2},\frac{L+M_k+M_v+1}{2}\right) =\frac{\pi }{2^{\kappa +1}}\frac{(2\kappa )!!}{(\kappa +1)!}, \end{aligned}$$
(37)
$$\begin{aligned} {}_3F_2\left( -M_k,-M_v,1;2,\frac{L-M_k-M_v}{2}+1;1\right) = \sum _{k=0}^\infty \frac{1}{(k+1)!} \frac{M_k!M_v!}{(M_k-k)!(M_v-k)!} \frac{\kappa _1!}{(\kappa _1+k)!}, \end{aligned}$$

where, for brevity, we introduce the shorthand notation \(\kappa =\frac{L+M_k+M_v}{2}\), \(\kappa _1=\frac{L-M_k-M_v}{2}\).

With this notation, the integral (34) reads

$$\begin{aligned} \int \limits _{-1}^1\sqrt{1-x^2} x^L G_{M_k}^{(1)}(x)G_{M_v}^{(1)}(x) dx= & {} \pi \frac{2^{M_k+M_v-\kappa -1} (M_k+1)!(M_v+1)!L!}{(\kappa +1)(2\kappa )!!} \nonumber \\\times & {} \sum _{k=0}^\infty \frac{1}{(k+1)!(M_k-k)!(M_v-k)!(\kappa _1+k)!}. \end{aligned}$$
(38)

In Eq. (38) the summation is restricted by those values of k which ensure non-negative factorials, i.e. in the above sum \(k\le M_k\) and \(k\le M_v\) and \((M_k+M_v-L+2k)\ge 0\). Together with the condition \((L+M_k+M_v)\)-even, these restrictions form the selecting rules for the integral (33). Actually, in practice the summation in (38) consists only of one, or, at maximum, two terms. Consequently, the integrals (33) turn out to be extremely simple, being expressed in form of the fractional parts of \(\pi \). For instance, the value L=0 results in the orthogonal condition for the Gegenbauer polynomials i.e. \({{\mathcal {K}}}_{M_k,M_v}^{L=0}=\frac{\pi }{2}\delta _{M_k,M_v}\). For \(L=1\) one has \(M_v=1,3,5,7\) and for even \(M_k=0,2,4,6,8\) the integral (33) is always \(\pi /4\). Here we present the corresponding explicit expressions of the integrals (33) for \(M_k^{max}=8\):

$$\begin{aligned}&{{\mathbf {L=0}}}:\qquad {{\mathcal {K}}}_{M_k,M_v}^{L=0}=\frac{\pi }{2} \delta _{M_k,M_v},\end{aligned}$$
(39)
$$\begin{aligned}&{{\mathbf {L=1}}}: \qquad M_k=0 \rightarrow {{\mathcal {K}}}_{0,1}^{L=1}=\frac{\pi }{4};\quad {{\mathcal {K}}}_{0,3}^{L=1}=0,\end{aligned}$$
(40)
$$\begin{aligned}&M_k=2 \rightarrow {{\mathcal {K}}}_{2,1}^{L=1}={{\mathcal {K}}}_{2,3}^{L=1}=\frac{\pi }{4},\end{aligned}$$
(41)
$$\begin{aligned}&M_k=4 \rightarrow {{\mathcal {K}}}_{4,3}^{L=1}={{\mathcal {K}}}_{4,5}^{L=1}=\frac{\pi }{4},\end{aligned}$$
(42)
$$\begin{aligned}&M_k=6 \rightarrow {{\mathcal {K}}}_{6,5}^{L=1}={{\mathcal {K}}}_{6,7}^{L=1}=\frac{\pi }{4},\end{aligned}$$
(43)
$$\begin{aligned}&M_k=8 \rightarrow {{\mathcal {K}}}_{8,7}^{L=1}={{\mathcal {K}}}_{8,9}^{L=1}=\frac{\pi }{4},\end{aligned}$$
(44)
$$\begin{aligned}&{{\mathbf {L=2}}}:\qquad M_k=0 \rightarrow {{\mathcal {K}}}_{0,0}^{L=2}={{\mathcal {K}}}_{0,2}^{L=2}=\frac{\pi }{8},\end{aligned}$$
(45)
$$\begin{aligned}&M_k=2 \rightarrow {{\mathcal {K}}}_{2,0}^{L=2}=\frac{\pi }{8}; \quad {{\mathcal {K}}}_{2,2}^{L=2}=\frac{\pi }{4};\quad {{\mathcal {K}}}_{2,4}^{L=2}=\frac{\pi }{8},\end{aligned}$$
(46)
$$\begin{aligned}&M_k=4 \rightarrow {{\mathcal {K}}}_{4,2}^{L=2}=\frac{\pi }{8}; \quad {{\mathcal {K}}}_{4,4}^{L=2}=\frac{\pi }{4};\quad {{\mathcal {K}}}_{4,6}^{L=2}=\frac{\pi }{8},\end{aligned}$$
(47)
$$\begin{aligned}&M_k=6 \rightarrow {{\mathcal {K}}}_{6,4}^{L=2}=\frac{\pi }{8}; \quad {{\mathcal {K}}}_{6,6}^{L=2}=\frac{\pi }{4};\quad {{\mathcal {K}}}_{6,8}^{L=2}=\frac{\pi }{8},\end{aligned}$$
(48)
$$\begin{aligned}&M_k=8 \rightarrow {{\mathcal {K}}}_{8,6}^{L=2}=\frac{\pi }{8}; \quad {{\mathcal {K}}}_{8,8}^{L=2}=\frac{\pi }{4};\quad {{\mathcal {K}}}_{8,10}^{L=2}=\frac{\pi }{8},\end{aligned}$$
(49)
$$\begin{aligned}&{{\mathbf {L=3}}}:\qquad M_k=0 \rightarrow {{\mathcal {K}}}_{0,1}^{L=3}=\frac{\pi }{8};\quad {{\mathcal {K}}}_{0,3}^{L=3}=\frac{\pi }{16},\end{aligned}$$
(50)
$$\begin{aligned}&M_k=2 \rightarrow {{\mathcal {K}}}_{2,1}^{L=3}=\frac{3\pi }{16};\quad {{\mathcal {K}}}_{2,3}^{L=3}=\frac{3\pi }{16};\quad {{\mathcal {K}}}_{2,5}^{L=3}=\frac{\pi }{16},\end{aligned}$$
(51)
$$\begin{aligned}&M_k=4 \rightarrow {{\mathcal {K}}}_{4,1}^{L=3}=\frac{\pi }{16};\quad {{\mathcal {K}}}_{4,3}^{L=3}=\frac{3\pi }{16};\quad {{\mathcal {K}}}_{4,5}^{L=3}=\frac{3\pi }{16}; \quad K_{4,7}^{L=3}=\frac{\pi }{16},\end{aligned}$$
(52)
$$\begin{aligned}&M_k=6 \rightarrow {{\mathcal {K}}}_{6,3}^{L=3}=\frac{\pi }{16};\quad {{\mathcal {K}}}_{6,5}^{L=3}=\frac{3\pi }{16};\quad {{\mathcal {K}}}_{6,7}^{L=3}=\frac{3\pi }{16}; \quad {{\mathcal {K}}}_{6,9}^{L=3}=\frac{\pi }{16},\end{aligned}$$
(53)
$$\begin{aligned}&M_k=8 \rightarrow {{\mathcal {K}}}_{8,5}^{L=3}=\frac{\pi }{16};\quad {{\mathcal {K}}}_{8,7}^{L=3}=\frac{3\pi }{16},\quad {{\mathcal {K}}}_{8,9}^{L=3}=\frac{3\pi }{16}; \quad {{\mathcal {K}}}_{8,11}^{L=3}=\frac{\pi }{16}. \end{aligned}$$
(54)

Appendix C: Explicit Expressions for the Coefficients \(C_{L,\delta }({\tilde{k}},{\tilde{p}},x_p)\) in Eq. (22)

Recall that \(M_p\) comes from the decomposition of the amplitude in l.h.s., \(M_k\) is the decomposition of the BS amplitude under the integral in the r.h.s. of Eq. (22), and \(M_v\) comes from decomposition of the rainbow interaction. Eventually, \(\delta \) corresponds to the power of the scalar product \((k\cdot p)^\delta \) and L is the power of \(k_4^L\) which we meet after contractions over the Lorentz indices. Results of the contraction provide \(L=0,1,2,3\) and \(\delta =0,1,2,3)\) and that \(L+\delta _{max}=3\). It implies that not all possible combination of \(L,\delta \) in \(C_{L,\delta }({\tilde{k}},{\tilde{p}})\) contribute to the r.h.s. There are 10 possible combinations: \((L,\delta ) =(0,0)\), \((L,\delta ) =(0,1)\), \((L,\delta ) =(0,2)\), \((L,\delta ) =(0,3)\), \((L,\delta ) =(1,0)\), \((L,\delta ) =(1,1)\), \((L,\delta ) =(1,2)\), \((L,\delta ) =(2,0)\), \((L,\delta ) =(2,1)\), \((L,\delta ) =(3,0)\). Explicitly one has

$$\begin{aligned} C_{0,0}({\tilde{p}},{\tilde{k}},x_p)&= 96 {\tilde{k}}^2{\tilde{p}}^2 (1 -x_p^2) M_{gg}^2 ({\tilde{k}}^2 +{\tilde{p}}^2); \quad C_{0,1}({\tilde{p}},{\tilde{k}},x_p) -12 M_{gg}^2 {\tilde{k}} {\tilde{p}}\nonumber \\&\qquad \qquad \times (-M_{gg}^ 2 {\tilde{k}}^2 - M_{gg}^2 {\tilde{p}}^2 + M_{gg}^2 {\tilde{p}}^2 x_p^2 + 12{\tilde{k}}^2 {\tilde{p}}^2 - 16 {\tilde{k}}^2{\tilde{p}}^2 x_p^2),\nonumber \\ C_{0,2}({\tilde{p}},{\tilde{k}},x_p)&= -24 M_{gg}^2 {\tilde{k}}^2{\tilde{p}}^2 \ (M_{gg}^2 + 4 {\tilde{k}}^2 + 4{\tilde{p}}^2);\quad C_{0,3}({\tilde{p}},{\tilde{k}},x_p)= 144 M_{gg}^2{\tilde{k}}^3{\tilde{p}}^3, \end{aligned}$$
(55)
$$\begin{aligned} C_{1,0}({\tilde{p}},{\tilde{k}},x_p)&=12 M_{gg}^2{\tilde{k}}{\tilde{p}} x_p \ (-M_{gg}^2 {\tilde{k}}^2-M_{gg}^2 {\tilde{p}}^2+M_{gg}^2 {\tilde{p}}^2 x_p^2 -4{\tilde{k}}^2{\tilde{p}}^2), \nonumber \\&\qquad \qquad C_{1,1}({\tilde{p}},{\tilde{k}},x_p) =48 M_{gg}^2{\tilde{k}}^2{\tilde{p}}^2 x_p, \ (M_{gg}^2 + 4 {\tilde{k}}^2 + 4{\tilde{p}}^2), \nonumber \\ C_{1,2}({\tilde{p}},{\tilde{k}},x_p)&= -336{\tilde{p}}^3 x_p M_{gg}^2 {\tilde{k}}^3, \end{aligned}$$
(56)
$$\begin{aligned} C_{2,0}({\tilde{p}},{\tilde{k}},x_p)&=-24 M_{gg}^2{\tilde{k}}^2 \ (M_{gg}^2 {\tilde{p}}^2*x_p^2+4 {\tilde{k}}^2* {\tilde{p}}^2 + 4{\tilde{p}}^4),\quad C_{2,1}({\tilde{p}},{\tilde{k}},x_p) = -12 {\tilde{p}} M_{gg}^2 {\tilde{k}}^3 (M_{gg}^2 - 16 {\tilde{p}}^2), \end{aligned}$$
(57)
$$\begin{aligned} C_{3,0}({\tilde{p}},{\tilde{k}},x_p)&=12 M_{gg}^4 {\tilde{k}}^3 {\tilde{p}} \ x_p. \end{aligned}$$
(58)

Actually, expressions (5558) imply that the summation \(\sum _{L,\delta }\) in Eq. (22) is restricted to \(\sum _{L=0}^3\sum _{\delta =0}^{\delta _{max}}\), where \(L+\delta _{max}=3\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kaptari, L.P., Kämpfer, B. Mass Spectrum of Pseudo-Scalar Glueballs from a Bethe–Salpeter Approach with the Rainbow–Ladder Truncation. Few-Body Syst 61, 28 (2020). https://doi.org/10.1007/s00601-020-01562-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00601-020-01562-4

Navigation