Skip to main content
Log in

Scaling properties of exclusive vector meson production cross section from gluon saturation

  • Regular Article
  • Published:
The European Physical Journal Plus Aims and scope Submit manuscript

Abstract

It is already known from phenomenological studies that in exclusive deep-inelastic scattering off nuclei there appears to be a scaling behavior of vector meson production cross section in both nuclear mass number, A, and photon virtuality, \(Q^{2}\), which is strongly modified due to gluon saturation effects. In this work we continue those studies in a realistic setup based upon using the Monte Carlo event generator Sartre. We make quantitative predictions for the kinematics of the Electron-Ion Collider, focusing on this A and \(Q^{2}\) scaling picture, along with establishing a small region of squared momentum transfer, t, where there are signs of this scaling that may potentially be observed at the EIC. Our results are represented as pseudo-data of vector meson production diffractive cross section and/or their ratios, which are obtained by parsing data collected by the event generator through smearing functions, emulating the proposed detector resolutions for the future EIC.

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
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

Data Availability Statement

This manuscript has associated data in a data repository. [Authors’ comment: The Monte Carlo generated pseudo-data sets obtained and analyzed during the current study, including the analysis codes, root files, plot-making scripts, are available at the repository [81]].

Notes

  1. References [16,17,18,19,20,21] discuss the first hints of the onset of gluon saturation stemming from \(p(d)+A\) collisions at RHIC energies, meanwhile, there are also alternative explanations [22,23,24] to consider.

  2. There is also another Monte Carlo generator, STARlight [52], which simulates a wide variety of vector meson final states, produced in \(e+A\) scattering. The improved version of this generator, dubbed as eSTARlight, has been used for exclusive vector meson production studies at the EIC kinematics [53].

  3. In any case the program has an imposed cutoff on the dipole size, r, against its any unphysical increase. For nuclei it is \(r < 3R_{N}\), where \(R_{N}\) is the nuclear radius given in the Woods–Saxon parametrization. For the proton it is \(r <3\) fm. This cutoff does not show any changes in final simulated cross sections, though it can be altered in a broad kinematic range.

  4. The phase factor \(\exp \!{[i(1 - z) {\mathbf {r_{T}}}\cdot {\varvec{\Delta }}]}\) standing in Eq. (4) stems from the difference between forward and non-forward wavefunctions. Although used in many phenomenological applications, the validity of the exponent \((1 - z) {\mathbf {r_{T}}}\cdot {\varvec{\Delta }}\) is challenged in [66], nevertheless, we keep it in our framework because it is anyway expected to have a negligible effect at small-x region [44].

  5. This cross section is Fourier transformed into momentum space with \({\varvec{\Delta }}\), which is the Fourier conjugate to the center-of-mass of the dipole, \({\mathbf {b_{T}}}- (1-z){\mathbf {r_{T}}}\), relative to the proton’s center (see Fig. 2 for details).

  6. In a dipole-target scattering there is no exchange of color charge, which is already mentioned in the introduction.

  7. It is shown in [72] how the incoherent cross section is affected by saturation effects that is not negligible for \(J/\Psi \) production.

  8. The ion beam energy here represents that of per nucleon.

  9. The procedure is bottomed on an assumption leading to using a constraint when the invariant mass of the outgoing nucleus should be taken as \(M_{A^{\prime }}^{2}\), to find the incoming electron’s longitudinal momentum, instead of just assuming the nominal electron beam momentum (see Sect. 8.4.6 of [30]).

  10. The \(\rho \) meson might be more sensitive to saturation effects at moderate \(Q^2\) than \(\phi \), nevertheless, there are large theoretical uncertainties in the knowledge of its wavefunction, making perturbative calculations less reliable at smaller \(Q^2\) values. Perhaps it also somewhat applies to \(\phi \)’s wavefunction but we prefer to work with \(\phi \) and \(J/\Psi \) following Accardi et al. [28] and Toll and Ullrich [50].

  11. The major motivation, which the EIC and LHeC proposals are based upon, is to search for gluon many-body nonlinear dynamics, and the gluon saturation in particular.

  12. Currently, the look-up amplitude tables in Sartre have maximum reach of \(Q^{2}\) as 20 GeV\(^{2}\) for \(e + Au\) and \(e + Ca\), and as 200 GeV\(^{2}\) for \(e + p\). We plan to update those tables to have a higher \(Q^{2}\) reach, along with making others for various \(e + A\) beams that can be found in Table 10.3 of [30].

  13. It should be emphasized that a better calculation gives an A-dependent parameter, standing in front of \(A^{4/3}\) in the r.h.s of Eq. (24), which, e.g., for the Gold nucleus is \(C[197] \approx 1/2\) [14].

  14. This effect is absent in the lower plot of Fig. 6 as the events are analyzed within a much narrower range of |t|.

  15. It is anticipated that full-scale studies of gluon saturation at EIC will largely benefit from collisions, which include light-, medium- and large-sized nuclei. Therefore, it would be highly desirable if the EIC community considers of having, e.g., d, Ca and Cu as parts of the baseline beam species, such as p and Au.

  16. Since our analysis was focused on and the scaling results for “indirectly observing” some hints of gluon saturation were produced in the momentum range of \(1< Q^{2} < 20\,\mathrm{GeV}^{2}\), then we would think of having a detector PID for such measurements in the future at EIC, at least, in that specified range of \(Q^{2}\).

  17. In the original Figure 54 of the EIC White Paper [28], the final-state particle pseudorapidity is restricted between \(-4.0< \eta < 4.0\). In our case, the Handbook smearing functions from [59] are implemented to cover tracking up to \(-3.5< \eta < 3.5\).

  18. The bin size in these figures changes as a function of \(Q^2\). When the pseudo-data are used to fill those asymmetrically sized bins, the error bars in each bin are calculated accordingly.

References

  1. F.D. Aaron, [H1 and ZEUS Collaborations] et al., JHEP 1001, 109 (2010)

  2. H. Abramowicz, [H1 and ZEUS Collaborations] et al., Eur. Phys. J. C 75, 580 (2015)

  3. L.V. Gribov, E.M. Levin, M.G. Ryskin, Phys. Rept. 100, 1 (1983)

    Article  ADS  Google Scholar 

  4. A.H. Mueller, J.W. Qiu, Nucl. Phys. B 268, 427 (1986)

    Article  ADS  Google Scholar 

  5. L.D. McLerran, R. Venugopalan, Phys. Rev. D 49, 2233 (1994)

    Article  ADS  Google Scholar 

  6. L.D. McLerran, R. Venugopalan, Phys. Rev. D 49, 3352 (1994)

    Article  ADS  Google Scholar 

  7. L.D. McLerran, R. Venugopalan, Phys. Rev. D 50, 2225 (1994)

    Article  ADS  Google Scholar 

  8. A. Ayala, J. Jalilian-Marian, L.D. McLerran, R. Venugopalan, Phys. Rev. D 53, 458 (1996)

    Article  ADS  Google Scholar 

  9. E. Iancu, A. Leonidov, L. McLerran, Nucl. Phys. A 692, 583 (2001)

    Article  ADS  Google Scholar 

  10. E. Ferreiro, E. Iancu, A. Leonidov, L. McLerran, Nucl. Phys. A 703, 489 (2002)

    Article  ADS  Google Scholar 

  11. F. Gelis, T. Lappi, R. Venugopalan, Int. J. Mod. Phys. E 16, 2595 (2007)

    Article  ADS  Google Scholar 

  12. F. Gelis, E. Iancu, J. Jalilian-Marian, R. Venugopalan, Ann. Rev. Nucl. Part. Sci. 60, 463 (2010)

    Article  ADS  Google Scholar 

  13. J.L. Albacete, N. Armesto, J.G. Milhano, P. Quiroga-Arias, C.A. Salgado, Eur. Phys. J. C 71, 1705 (2011)

    Article  ADS  Google Scholar 

  14. H. Mántysaari, P. Zurita, Phys. Rev. D 98, 036002 (2018)

    Article  ADS  Google Scholar 

  15. T. Lappi, H. Mäntysaari, Phys. Rev. D 88, 114020 (2013)

    Article  ADS  Google Scholar 

  16. I. Arsene, [BRAHMS Collaboration] et al., Phys. Rev. Lett. 93, 242303 (2004)

  17. J. Adams, [STAR Collaboration] et al., Phys. Rev. Lett. 97, 152302 (2006)

  18. E. Braidot, [STAR Collaboration], Nucl. Phys. A 854, 168 (2011)

  19. A. Adare, [PHENIX Collaboration] et al., Phys. Rev. Lett. 107, 172301 (2011)

  20. J.L. Albacete, C. Marquet, Phys. Rev. Lett. 105, 162301 (2010)

    Article  ADS  Google Scholar 

  21. T. Lappi, H. Mantysaari, Nucl. Phys. A 908, 51 (2013)

    Article  ADS  Google Scholar 

  22. M. Strikman, W. Vogelsang, Phys. Rev. D 83, 034029 (2011)

    Article  ADS  Google Scholar 

  23. Z.-B. Kang, I. Vitev, H. Xing, Phys. Rev. D 85, 054024 (2012)

    Article  ADS  Google Scholar 

  24. Z.-B. Kang, I. Vitev, H. Xing, Phys. Lett. B 718, 482 (2012)

    Article  ADS  Google Scholar 

  25. H. Kowalski, T. Lappi, R. Venugopalan, Phys. Rev. Lett. 100, 022303 (2008)

    Article  ADS  Google Scholar 

  26. H. Kowalski, D. Teaney, Phys. Rev. D 68, 114005 (2003)

    Article  ADS  Google Scholar 

  27. D. Boer et al., Gluons and the quark sea at high energies: distributions, polarization, tomography. arXiv:1108.1713 [nucl-th]

  28. A. Accardi et al., Eur. Phys. J. A 52, 268 (2016)

    Article  ADS  Google Scholar 

  29. E.C. Aschenauer et al., Rept. Prog. Phys. 82, 024301 (2019)

    Article  ADS  Google Scholar 

  30. R. Abdul Khalek, A. Accardi, J. Adam, D. Adamiak, W. Akers, M. Albaladejo, A. Al-bataineh, M.G. Alexeev, F. Ameli and P. Antonioli, et al., Nucl. Phys. A 1026, 122447 (2022). arXiv:2103.05419 [physics.ins-det]

  31. J.L. Abelleira Fernandez, [LHeC Study Group Collaboration] et al., J. Phys. G 39, 075001 (2012)

  32. E. Abbas, [ALICE Collaboration] et al., Eur. Phys. J. C 73, 2617 (2013)

  33. B. Abelev, [ALICE Collaboration] et al., Phys. Lett. B 718, 1273 (2013)

  34. J. Adam, [ALICE Collaboration] et al., JHEP 1509, 095 (2015)

  35. V. Khachatryan, [CMS Collaboration] et al., Phys. Lett. B 772, 489 (2017)

  36. B.B. Abelev, [ALICE Collaboration] et al., Phys. Rev. Lett. 113, 232504 (2014)

  37. R. Chudasama, [CMS Collaboration] et al., PoS ICPAQGP 2015, 042 (2017)

  38. S. Acharya, [ALICE Collaboration] et al., Eur. Phys. J. C 79, 402 (2019)

  39. D.E. Sosnov, [CMS], Nucl. Phys. A 1005, 121857 (2021)

  40. S. Acharya, [ALICE Collaboration] et al., Phys. Lett. B 820, 136481 (2021)

  41. S. Acharya, [ALICE Collaboration], et al. arXiv:2101.04577 [nucl-ex]

  42. M.G. Ryskin, Z. Phys. C 57, 89 (1993)

    Article  ADS  Google Scholar 

  43. D. Bendova, J. Cepila, J.G. Contreras, V.P. Gonçalves, M. Matas, Eur. Phys. J. C 81(3), 211 (2021)

    Article  ADS  Google Scholar 

  44. H. Mäntysaari, R. Venugopalan, Phys. Lett. B 781, 664 (2018)

    Article  ADS  Google Scholar 

  45. K.J. Golec-Biernat, M. Wusthoff, Phys. Rev. D 59, 014017 (1998)

    Article  ADS  Google Scholar 

  46. K.J. Golec-Biernat, M. Wusthoff, Phys. Rev. D 60, 114023 (1999)

    Article  ADS  Google Scholar 

  47. J. Bartels, K.J. Golec-Biernat, H. Kowalski, Phys. Rev. D 66, 014001 (2002)

    Article  ADS  Google Scholar 

  48. H. Kowalski, L. Motyka, G. Watt, Phys. Rev. D 74, 074016 (2006)

    Article  ADS  Google Scholar 

  49. T. Toll, T. Ullrich, Comput. Phys. Commun. 185, 1835 (2014). https://sartre.hepforge.org/

  50. T. Toll, T. Ullrich, Phys. Rev. C 87, 024913 (2013)

    Article  ADS  Google Scholar 

  51. D. Bendova, J. Cepila, V.P. Gonçalves, C.R. Sena, Eur. Phys. J. C 82, 99 (2022)

    Article  ADS  Google Scholar 

  52. S.R. Klein, J. Nystrand, J. Seger, Y. Gorbunov, J. Butterworth, Comput. Phys. Commun. 212, 258–268 (2017)

    Article  ADS  Google Scholar 

  53. M. Lomnitz, S. Klein, Phys. Rev. C 99, 015203 (2019)

    Article  ADS  Google Scholar 

  54. B. Sambasivam, T. Toll, T. Ullrich, Phys. Lett. B 803, 135277 (2020)

    Article  Google Scholar 

  55. H. Mäntysaari, B. Schenke, Phys. Rev. Lett. 117, 052301 (2016)

    Article  ADS  Google Scholar 

  56. H. Mäntysaari, B. Schenke, Phys. Rev. D 94, 034042 (2016)

    Article  ADS  Google Scholar 

  57. H. Mäntysaari, B. Schenke, Phys. Lett. B 772, 832 (2017)

    Article  ADS  Google Scholar 

  58. H. Mäntysaari, Rept. Prog. Phys. 83(8), 082201 (2020)

    Article  ADS  Google Scholar 

  59. Electron-Ion Collider Detector Requirements and R &D Handbook. http://web.archive.org/web/20220120214345/. https://www.eicug.org/web/sites/default/files/EIC_HANDBOOK_v1.1.pdf

  60. H. Kowalski, T. Lappi, C. Marquet, R. Venugopalan, Phys. Rev. C 78, 045201 (2008)

    Article  ADS  Google Scholar 

  61. M.L. Good, W.D. Walker, Phys. Rev. 120, 1857 (1960)

    Article  ADS  Google Scholar 

  62. Y.V. Kovchegov, L.D. McLerran, Phys. Rev. D 60, 054025 (1999). Erratum: [Phys. Rev. D 62, 019901 (2000)]

  63. L. Frankfurt, M. Strikman, D. Treleani, C. Weiss, Phys. Rev. Lett. 101, 202003 (2008)

    Article  ADS  Google Scholar 

  64. A. Caldwell, H. Kowalski, Phys. Rev. C 81, 025203 (2010)

    Article  ADS  Google Scholar 

  65. H.I. Miettinen, J. Pumplin, Phys. Rev. D 18, 1696 (1978)

    Article  ADS  Google Scholar 

  66. Y. Hatta, B.W. Xiao, F. Yuan, Phys. Rev. D 95, 114026 (2017)

    Article  ADS  Google Scholar 

  67. A.D. Martin, M.G. Ryskin, Phys. Rev. D 57, 6692 (1998)

    Article  ADS  Google Scholar 

  68. A.G. Shuvaev, K.J. Golec-Biernat, A.D. Martin, M.G. Ryskin, Phys. Rev. D 60, 014015 (1999)

    Article  ADS  Google Scholar 

  69. A.D. Martin, M.G. Ryskin, T. Teubner, Phys. Rev. D 62, 014022 (2000)

    Article  ADS  Google Scholar 

  70. M. Krelina, V.P. Goncalves, J. Cepila, Nucl. Phys. A 989, 187–200 (2019)

    Article  ADS  Google Scholar 

  71. W. Chang, E.C. Aschenauer, M.D. Baker, A. Jentsch, J.H. Lee, Z. Tu, Z. Yin, L. Zheng, Phys. Rev. D 104(11), 114030 (2021)

    Article  ADS  Google Scholar 

  72. T. Lappi, H. Mantysaari, Phys. Rev. C 83, 065202 (2011). [arXiv:1011.1988 [hep-ph]]

    Article  ADS  Google Scholar 

  73. M. Defurne, [Jefferson Lab Hall A] et al., Phys. Rev. Lett. 117, 262001 (2016)

  74. P.A. Zyla, [Particle Data Group], et al., Prog. Theor. Exp. Phys. 2020, 083C01 (2020) and 2021 update. https://pdg.lbl.gov/2021/listings/rpp2021-list-J-psi-1S.pdf

  75. P.A. Zyla, [Particle Data Group], et al., Prog. Theor. Exp. Phys. 2020, 083C01 (2020) and 2021 update. https://pdg.lbl.gov/2021/listings/rpp2021-list-phi-1020.pdf

  76. I. Angeli, K.P. Marinova, At. Data Nucl. Data Tables 99, 59 (2013)

    Article  ADS  Google Scholar 

  77. W. Xiong et al., Nature 575(7781), 147–150 (2019)

    Article  ADS  Google Scholar 

  78. F.G. Ben, M.V.T. Machado, W.K. Sauter, Phys. Rev. D 96, 054015 (2017)

    Article  ADS  Google Scholar 

  79. Design of the ECCE Detector for the Electron Ion Collider. https://www.researchgate.net/publication/363331598_Design_of_the_ECCE_Detector_for_the_Electron_Ion_Collider

  80. J. Adam [ATHENA], et al., JINST 17(10), P10019 (2022). arXiv:2210.09048 [physics.ins-det]

  81. G. Matousek, Analysis codes, data files and plotting scripts for the paper. arXiv:2202.05981 [nucl-th]. https://github.com/Gregtom3/vlad_dvmp_plots/

  82. J. Breitweg, [ZEUS Collaboration] et al., Eur. Phys. J. C 6, 603–627 (1999)

  83. C. Adloff, [H1 Collaboration] et al., Eur. Phys. J. C 13, 371–396 (2000)

  84. J.R. Smith, An experimentalist’s guide to photon flux calculations. UCD/IIRPA 92-24 (1992) and H1 Internal Note H1-259 (1992). https://lss.fnal.gov/archive/other/ucd-iirpa-92-24.pdf

  85. J.R. Smith, Polarization Decomposition of Fluxes and Kinematics in ep Reaction. UCD/IIRPA 93-10 (1993) and H1 Internal Note H1-282 (1993). https://lss.fnal.gov/archive/other/ucd-iirpa-93-10.pdf

  86. L.N. Hand, Phys. Rev. 129, 1834 (1963)

    Article  ADS  Google Scholar 

  87. S.R. Klein, J. Nystrand, Phys. Rev. Lett. 84, 2330 (2000)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We are very grateful to Heikki Mäntysaari and Björn Schenke for the reading of the manuscript and giving us extremely useful comments, based on which the results of the paper have been improved. We are also thankful to Abhay Deshpande, Haiyan Gao, Barak Schmookler, Tobias Toll, Thomas Ullrich and Raju Venugopalan for fruitful and informative discussions on the subject matter. The work of Gregory Matousek is supported in part by Duke University. The work of Vladimir Khachatryan is supported in part by the U.S. Department of Energy, Office of Science, Offices of Nuclear Physics under contract DE-FG02-03ER41231. The work of Jinlong Zhang is supported by the Qilu Youth Scholar Funding of Shandong University.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vladimir Khachatryan.

Appendices

Appendix A: Calculating diffractive differential cross sections

An extremely complicated task is to calculate and generate total cross sections, for which one has to evaluate complex four-dimensional integrals at each phase-space point. But Sartre uses an approach based on computing the first and second moments of the scattering amplitudes separately, and then stores the results in three-dimensional look-up tables, in terms of \(Q^{2}\), \(W^{2}\) and t independent variables. The ultimate outcome is a set of four look-up tables for each nuclear species, each final-state vector meson (and DVCS photon), each polarization, and each dipole model (either IPSat or IPNonSat).

$$\begin{aligned}&\bigg<\mathcal {A}_{T}(Q^2, W^2, t)\bigg>_{\Omega } ,\quad \bigg<\mathcal {A}_{L}(Q^2, W^2, t)\bigg>_{\Omega } , \nonumber \\&\bigg<|\mathcal {A}_{T}(Q^2, W^2, t)|^2\bigg>_{\Omega } ,\quad \bigg <|\mathcal {A}_{L}(Q^2, W^2, t)|^2\bigg >_{\Omega } , \end{aligned}$$
(A.1)

These look-up tables contain all the physics information from both dipole models. The program also provides tables for calculating the phenomenological corrections described in Sect. 2.2.2.

The master equation of Sartre is the total diffractive differential cross section, which for electron-nucleus scattering has the following form:

$$\begin{aligned} \left( \frac{\mathrm{d}^{3}\sigma ^{\gamma ^{*}A\,\rightarrow \,V A}}{\mathrm{d}Q^{2}\,\mathrm{d}W^{2}\,\mathrm{d}t} \right)_{\mathrm{tot. diff.}}= & \sum _{T, L}\frac{R_{g}^{2}(1 + B^{2})}{16\pi }\frac{\mathrm{d}\Gamma _{T, L}}{\mathrm{d}Q^{2}\,\mathrm{d}W^{2}}\nonumber \\&\times \,\bigg < \left| \mathcal {A}_{T,L}^{\gamma ^{*}A\,\rightarrow \,V A}(x_{I\!\!P},Q^{2},t,\Omega )\right| ^{2} \bigg >_{\Omega } , \end{aligned}$$
(A.2)

where B is given in Eq. (18), and \(R_{g}\) in Eq. (19). The quantity \(\Gamma _{T, L}\) is the flux of transversely T and longitudinally L polarized virtual photons [82,83,84,85,86], given by

$$\begin{aligned}&\Gamma _{T} = \frac{\alpha }{2\pi } \left( \frac{1 + (1 - y)^{2}}{y\,Q^{2}} \right) ,\nonumber \\&\Gamma _{L} = \frac{\alpha }{2\pi } \left( \frac{2(1 - y)}{y\,Q^{2}} \right) , \end{aligned}$$
(A.3)

where y is the inelasticity defined as the fraction of the electron’s energy lost in the nucleon rest frame. The averaging over configurations \(\Omega \) is defined as

$$\begin{aligned}&\bigg < \left| \mathcal {A}_{T,L}^{\gamma ^{*}A\,\rightarrow \,V A}(x_{I\!\!P},Q^{2},t,\Omega )\right| ^{2} \bigg >_\Omega \nonumber \\&\quad = \frac{1}{N_{\mathrm{max}}}\sum _{j=1}^{N_{\mathrm{max}}} \left| \mathcal {A}_{T,L}^{\gamma ^{*}A\,\rightarrow \,V A} (x_{I\!\!P},Q^{2},t,\Omega _{j})\right| ^{2} , \end{aligned}$$
(A.4)

where \(N_{\mathrm{max}}\) is the number of configurations. If it is large enough, then the sum in Eq. (A.4) converges to a true average. It is already shown in [50] that \(2 \times 500\) configurations, 500 for T polarized \(\gamma ^{*}\) and 500 for L polarized \(\gamma ^{*}\), give a good convergence. Such that there are 1000 such integrals for each \((Q^{2}, W^{2}, t)\) phase-space point.

The photon flux may emanate from electrons, as in the case of \(e+p\) and \(e+A\) scatterings, however, it may be also radiated from protons or nuclei [87], as in the case of \(A+A\) or \(p+A\) UPC. A phase-space point together with a given beam energy fully determines the final state of a produced vector meson, except for its azimuthal angle, which is distributed uniformly.

The second moment of the amplitude in Eq. (A.4) for the nucleon configurations \(\Omega _{j}\) can be calculated based upon [49, 50]:

$$\begin{aligned} \bigg < \left| \mathcal {A}_{T,L}^{\gamma ^{*}A\,\rightarrow \,V A}(x_{I\!\!P},Q^{2},t,\Omega ) \right| ^{2}\bigg >_{\Omega }= & \sum _{j=1}^{N_{\mathrm{max}}}\,\left| \int \mathrm{d}^{2}{} \mathbf{r_{T}} \int \mathrm{d}^{2}{\mathbf {b_{T}}}\int \frac{\mathrm{d} z}{4\pi } \,\times \right. \nonumber \\&\times \,\left. \left[ \left( \Psi ^{*}\Psi _{V} \right) _{T,L}\!(Q^{2},{\mathbf {r_{T}}}, z)\right] J_{0}([1 - z]r_{T}\Delta ) \right. \nonumber \\&\quad \times \,\left. e^{-i{\mathbf {b_{T}}}\cdot {\varvec{\Delta }}}\, \frac{\mathrm{d}\sigma _{q\bar{q}}^{A}}{\mathrm{d}^{2}{\mathbf {b_{T}}}}(x_{I\!\!P}, {\mathbf {r_{T}}}, {\mathbf {b_{T}}}, \Omega _{j})\right| ^{2} , \end{aligned}$$
(A.5)

where the last term \(\mathrm{d}\sigma _{q\bar{q}}^{A}/\mathrm{d}^{2}{\mathbf {b_{T}}}\) is defined in Eq. (16).

Thus, Eqs. (A.2) and (A.5) determine the total diffractive differential cross section. Its coherent part is given by

$$\begin{aligned} \left( \frac{\mathrm{d}^{3}\sigma ^{\gamma ^{*}A\,\rightarrow \,V A}}{\mathrm{d}Q^{2}\,\mathrm{d}W^{2}\,\mathrm{d}t} \right) _{\mathrm{coh. diff.}}= & \sum _{T, L} \frac{R_{g}^{2}(1 + B^{2})}{16\pi }\frac{\mathrm{d}\Gamma _{T, L}}{\mathrm{d}Q^{2}\,\mathrm{d}W^{2}}\nonumber \\&\times \,\left| \bigg <\mathcal {A}_{T,L}^{\gamma ^{*}A\,\rightarrow \,V A}(x_{I\!\!P},Q^{2},t,\Omega )\bigg >_{\Omega } \right| ^{2} , \end{aligned}$$
(A.6)

For the first moment of the amplitude, the integral to calculate will be

$$\begin{aligned} \,\left| \bigg<\mathcal {A}_{T,L}^{\gamma ^{*}A\,\rightarrow \,V A}(x_{I\!\!P},Q^{2},t,\Omega )\bigg>_{\Omega } \right| ^{2}= & \left| \int \mathrm{d}^{2}{} \mathbf{r_{T}} \int \mathrm{d}^{2}{\mathbf {b_{T}}}\right. \nonumber \\&\left. \times \,\int \frac{\mathrm{d} z}{4\pi } \left[ \left( \Psi ^{*}\Psi _{V} \right) _{T,L}\!(Q^{2},{\mathbf {r_{T}}}, z)\right] J_{0}([1 - z]r_{T}\Delta ) \right. \nonumber \\&\left. \times \,J_{0}(b_{T}\Delta ) \bigg <\frac{\mathrm{d}\sigma _{q\bar{q}}^{A}}{\mathrm{d}^{2}{\mathbf {b_{T}}}}(x_{I\!\!P}, {\mathbf {r_{T}}}, {\mathbf {b_{T}}}, \Omega )\bigg >_{\Omega } \right| ^{2} , \end{aligned}$$
(A.7)

where the average \(< \mathrm{d}\sigma _{q\bar{q}}^{A}/\mathrm{d}^{2}{\mathbf {b_{T}}}>_{\Omega }\) in the last term is defined in Eq. (17).

Thereby, as in Eq. (14), the incoherent part of the total diffractive differential cross section is taken to be the difference between the total and coherent cross sections:

$$\begin{aligned} \left( \frac{\mathrm{d}^{3}\sigma ^{\gamma ^{*}A\,\rightarrow \,V A}}{\mathrm{d}Q^{2}\,\mathrm{d}W^{2}\,\mathrm{d}t} \right) _{\mathrm{incoh. diff.}}&= \left( \frac{\mathrm{d}^{3}\sigma ^{\gamma ^{*}A\,\rightarrow \,V A}}{\mathrm{d}Q^{2}\,\mathrm{d}W^{2}\,\mathrm{d}t} \right)_{\mathrm{tot. diff.}} \nonumber \\ &\quad - \,\left( \frac{\mathrm{d}^{3}\sigma ^{\gamma ^{*}A\,\rightarrow \,V A}}{\mathrm{d}Q^{2}\,\mathrm{d}W^{2}\,\mathrm{d}t} \right) _{\mathrm{coh. diff.}} . \end{aligned}$$
(A.8)

The incoherent part directly gives the probability for the nuclear breakup.

Appendix B: Error analysis of the pseudo-data using the EIC Detector Handbook

We use Sartre 1.33, an exclusive event generator, to produce vector meson production data. A user-edited runcard is called at the initialization of a simulation to set beam energies, decay modes, and ranges on event kinematics such as \(Q^{2}\) and |t|. The generator outputs the kinematics of final-state particles, such as their momentum and pseudorapidity. Additionally, important event information such as \(Q^{2}\), \({x_{\mathbb {P}}}\), and y are recorded. We refer to the event generator output as truth data. This data, while unobtainable in a physical experiment, allows us to perform perfect event identification and to create pseudo-data. Using detector resolutions outlined in the EIC Detector Handbook [59], true particle kinematics are smeared to create pseudo-data. For instance, according to the handbook, the barrel (\(|\eta |<1\)) tracking resolution for electrons is \(\sigma _p/p = 0.05\%p + 0.5\%\). By having such resolutions written out explicitly for the relevant final-state particles from vector meson production simulations allows us to calculate pseudo-data based on the handbook’s projections. The smearing functions immediately produce smeared kinematics, such as momentum and energy. Event-by-event, we manually smear the final-state kinematics according to these functions. Subsequently, the pseudo-data reconstruction of the scattered electron is used to determine event kinematics such as \(Q^{2}\). The event |t| is the only quantity which we smear independently. One of the immediate effects of generating pseudo-data is the reduction of statistics. Final-state particles, which exit at pseudorapidities beyond the coverage of the detector system, render the exclusive event unrecoverable. Systematic errors introduced through particle/event mis-identification are not accounted for when analyzing the pseudo-data. Instead, we opt to use true, event generator information to identify the scattered electron and decay particles. Lastly, we use truth information to separate coherent vector meson production from the incoherent one, as well as events with transversely polarized virtual photons from longitudinally polarized virtual photons.

After the pseudo-data are generated, we are left with the reconstructed particle and event kinematics for a pile of production events. Then, the data are stored event-by-event in a ROOT TTree, which is read and analyzed to produce plots. From the TTree, the analyzed data are used to fill ROOT histogram objects. These histogram objects, initially storing the number of events per bin, are scaled in a variety of ways. This includes scaling to the correct decay mode branching ratios, scaling to the desired luminosity, and scaling by bin size (to create differential cross sections). What follows is a detailed look at how the pseudo-data are analyzed, how this analysis is piped into histograms, how they are scaled, and how the error is propagated.

Pseudo-data and error propagation for Figs. 3 and 4. After the pseudo-data are generated, the following cuts are placed event-by-event:

  • Reconstructed \({x_{\mathbb {P}}}< 0.01\).

  • Reconstructed \(1.0< Q^{2} < 10.0\,\mathrm{GeV}^{2}\).

  • True event coherent vector meson production.

  • Reconstructed final-state particle pseudorapidity between \(-3.5< \eta < 3.5\)Footnote 17.

  • Reconstructed final-state particle momentum greater than 1 GeV.

Then, using true event |t| information outputted by Sartre, we generate the reconstructed |t| using the Gaus() function from the ROOT TRandom class. For each event, we select a random variable from a Gaussian centered at true |t| with spread equal to true \(|t|\times tsmear \). This random variable represents the reconstructed |t| of the event. The quantity \( tsmear \) is changed depending on our desired resolution in |t|. It ranges from 0 all the way to \(0.3\,\mathrm{GeV}^{2}\), or is parameterized by the “Method L” in [30]. Once all of the events have been parsed through and either added or discarded, a TGraphErrors object is created. This object will pick up information from the filled histogram and produce a plot.

Below are the individual modifications we make to the entries of the filled histogram, given in order. Beforehand, we define \(\mathrm {N}_{\mathrm {entries}}\) to be the number of total entries in the histogram, and w to be the amount of entries in an arbitrary bin.

  1. 1.

    First, we multiply each bin of the histogram by the quantity \(\sigma \cdot L/A\), where \(\sigma \) is the total event cross section (in nanobarns) outputted by Sartre, and L is the integrated luminosity. We also divide by the atomic number A of a nucleus to isolate the scattering cross section of a single nucleon for a given nucleus beam.

  2. 2.

    Then, we multiply each bin by the branching ratio (BR) of the truth decay mode.

  3. 3.

    Next, we divide each bin by \(\mathrm {N}_{\mathrm {entries}}\). And now for each bin, we are left with the number of events expected in an experiment with integrated luminosity L, within the reconstructed |t| bin range:

    $$\begin{aligned} \mathrm {N}_{\mathrm {expected}} \equiv w' = \frac{\mathrm {BR} \times L\times \sigma \times w}{\mathrm {N}_{\mathrm {entries}} \times A}. \end{aligned}$$
    (B.9)
  4. 4.

    We further calculate the statistical uncertainty for each bin as the square root of its entries, \(\sigma _{w'} = \sqrt{w'}\).

  5. 5.

    Afterward, we divide each bin quantity w by the branching ratio, the bin width \(\Delta t\), and integrated luminosity. We repeat this for \(\sigma _{w}\) as well.

    $$\begin{aligned}&\frac{d\sigma }{dt}{\text{[pseudo-data] }} = \frac{A}{\mathrm {BR} \times \Delta t \times L} \mathrm {N}_{\mathrm {expected}} ,\nonumber \\&\frac{d\sigma }{dt}{\text{[pseudo-data } \text{ statistical } \text{ uncertainty] }}= \frac{A}{\mathrm {BR}\times \Delta t \times L} \sqrt{\mathrm {N}_{\mathrm {expected}}} . \end{aligned}$$
    (B.10)

This step completes the pseudo-data and error analysis for Figs. 3 and 4.

Fig. 16
figure 16

Cross-sectional conversion scheme for obtaining the scaling plots in Figs. 5, 6, 7, 8, 9, 10, 11, 12, 13, 14 and 15 based upon Sartre 1.33 simulations.

Pseudo-data and error propagation for Figs. 5, 6, 7, 8, 9, 10, 11, 12, 13, 14 and 15. The analysis of these figures differ from the above method in subtle ways.Footnote 18 For each collision type (\(e+p\) and/or \(e+A\)), vector meson produced (\(J/\Psi \) or \(\phi \)), and dipole model used (IPSat or IPNonSat), we generate \(N = 10^{8}\) exclusive events using Sartre. At the beginning of the simulation, we save the total cross section of the simulation’s phase space. Using this cross section, the projected EIC luminosities (\(100\,\mathrm{fb}^{-1}/A\) for \(e+p\) at \(10 \times 100\,\mathrm{GeV}\), \(10\,\mathrm{fb}^{-1}/A\) for \(e+A\) at \(10 \times 110\,\mathrm{GeV}\)), and the branching ratios of the studied vector mesons’ decay modes, we calculate the number of expected events produced at the EIC. This number, we call n, can be divided by N to tell us how much we should scale our simulation size to reflect the specific luminosity. We call this parameter \(\text{ Scale } = n/N\).

Event by event, we use the true event generator output to determine if the event was coherent or incoherent. Skipping incoherent events, we then store the virtual photon polarization for an event, also given by Sartre. We place a hard cut on the true event y to minimize the impact of poor smearing in that kinematic regime. By using detector uncertainties projected in [59], we smear the energy, momentum, and polar angle of the final-state electron and decay products. We skip events in which any one of these three-final state particles true pseudorapidity falls outside the reach of the detector system’s coverage. After the smearing is completed, we use the true beam energy of the incoming electron and nuclear/proton target to calculate event kinematics such as \(Q^{2}\), y, and \({x_{\mathbb {P}}}\). We smear the event’s t in a way described in Sect. 3, skipping events with \(y < 0.05\).

When filling our histograms, we weigh the fill by the factor

$$\begin{aligned} \text{ Scale } \times \frac{d{x_{\mathbb {P}}}}{dy} \frac{Q^{m}}{\Gamma _{T, L}} , \end{aligned}$$
(B.11)

where m is determined by the specific \(Q^{2}\) scaling being analyzed, the photon flux factor \(\Gamma _{T, L}\) depends on the polarization of the event, and the derivative is given by

$$\begin{aligned} \frac{d{x_{\mathbb {P}}}}{dy} = \frac{s \times {x_{\mathbb {P}}}}{M_{V}^{2} + Q^{2} - t} . \end{aligned}$$
(B.12)

where s is the center-of-mass energy squared. The derivative \(d{x_{\mathbb {P}}}/dy\) and \(\Gamma _{T, L}\) factor effectively divide out the virtual photon flux element of the event cross section, leaving us with (see Fig. 16) a beam-independent cross section \(\sigma \left( \gamma ^{*} + p/A\,\rightarrow \,V + p^{\prime }/A^{\prime } \right) \). The scale factor is included to reflect the number of events expected, given the experimental luminosity. If we had run Sartre and generated exactly the number of events expected in a given luminosity, then the Scale would equal 1. Each time a histogram’s bin is filled, ROOT updates the statistical error of the bin to be the summation in quadrature of all the bin’s individual weights.

After all events have been simulated, we divide each bin of each histogram by the factor \(\mathrm {BR} \times L/A\). We then divide out by the \(Q^{2}\) bin width, \({x_{\mathbb {P}}}\) range, and proper nuclear size \(A^{k}\) scaling, which leaves the bin content being equivalent to \(A^{-k}\,Q^{m}\,\sigma ^{T, L}\). To obtain the differential cross section, we finally divide by the t binning, which is equal 0.001. The proper nuclear size A scaling and is applied to each histogram depending on x. For the nonzero |t| scaling plots in Figs. 10, 11, 12 and 13, we additionally divide the vertical axis by G(At) (see Eq. (33)).

Appendix C. Addendum to Sect. 3.2.1

In this appendix, we show qualitative comparisons between model calculations and simulated pseudo-data on the \(A^{2}\) normalized cross-sectional ratio and normalized \(A^{2}/Q^{6}\) cross section shown in Sect. 3.2.1, in Figs. 6 and 8, respectively. The pseudo-data are obtained by passing events generated by Sartre through particle kinematics smearing functions with detector acceptance. In these plots, the high \({x_{\mathbb {P}}}\) denotes the range of pseudo-data made in \({x_{\mathbb {P}}}= [0.005 - 0.009]\), and the low \({x_{\mathbb {P}}}\) denotes the range of \({x_{\mathbb {P}}}= [0.001 - 0.005\)]. So, Figs. 17 and 18 have their top and bottom panels being the same as in Figs. 6 and 8, respectively. We also display truth curves corresponding to all sets of the shown pseudo-data that are calculated using the IPSat model in Sartre. The “kink” structures seen in the curves around \(\sim 3~\mathrm{GeV^{2}}\) in the top high-\({x_{\mathbb {P}}}\) plots, which describe the \(\phi \) normalized cross-sectional ratio and cross section, partially come from the fact that these curves are actually calculated at fixed value of \({x_{\mathbb {P}}}= 0.008\). This region of \({x_{\mathbb {P}}}\) is close to 0.01, the point at which Sartre’s reliability starts to fail, especially for light vector meson production. Lastly, the low-\({x_{\mathbb {P}}}\) curves are calculated at 0.0012.

Fig. 17
figure 17

In the top and bottom panels, the pseudo-data and labeling are exactly the same as in Fig. 6. In addition, there are also overlaid curves obtained directly from Sartre’s IPSat model, where any detector smearing/resolution and experimental acceptance considerations are excluded in the normalized cross-sectional ratio computations

Fig. 18
figure 18

In the top and bottom panels, the pseudo-data and labeling are exactly the same as in Fig. 8. The explanation for the overlaid curves is similar as in Fig. 17

Overall, the need for additional reconstruction algorithms is evidenced by these figures. The comparison between the truth curves and pseudo-data should thus at this point stay qualitative. A future analysis will consider bin-by-bin corrections, such as detection efficiency and bin migration, to alleviate the discrepancies between the Monte Carlo and pseudo-data cross-sectional reconstructions. Although one may expect these systematic effects to “divide themselves out” in the ratio plots, the differing center-of-mass energies planned for the \(e+p\) and \(e+A\) runs at the future EIC makes this assumption false. During this paper’s composition, the EIC collaboration has begun narrowing down a single detector design for the first interaction region. Coined EPIC, the detector system and relevant software is currently in development, making our future analysis of detector effects for the continuation of the study in this paper extremely timely.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Matousek, G., Khachatryan, V. & Zhang, J. Scaling properties of exclusive vector meson production cross section from gluon saturation. Eur. Phys. J. Plus 138, 113 (2023). https://doi.org/10.1140/epjp/s13360-023-03729-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjp/s13360-023-03729-4

Navigation