Skip to main content
Log in

Diffraction at the LHC

  • Regular Article - Theoretical Physics
  • Published:
The European Physical Journal C Aims and scope Submit manuscript

Abstract

We show that the diffractive pp (and \(p\bar{p}\)) data (on σ tot, el/dt, proton dissociation into low-mass systems, \(\sigma^{\mathrm{D}}_{\mathrm{low}M}\), and high-mass dissociation, /dη)) in a wide energy range from CERN-ISR to LHC energies, may be described in a two-channel eikonal model with only one ‘effective’ pomeron. By allowing the pomeron coupling to the diffractive eigenstates to depend on the collider energy (as is expected theoretically) we are able to explain the low value of \(\sigma^{\mathrm{D}}_{\mathrm{low}M}\) measured at the LHC. We calculate the survival probability, S 2, of a rapidity gap to survive ‘soft rescattering’. We emphasise that the values found for S 2 are particularly sensitive to the detailed structure of the diffractive eigenstates.

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

Similar content being viewed by others

Notes

  1. This vertex factor is denoted V(pN ) below. We use N as a generic name for low-mass nucleon resonances and other low-mass excitations.

  2. Actually the TOTEM result is based on the difference between the total rate of inelastic events, obtained using optical theorem, and the observed rate of events with at least one charged particle with |η|<6.5. According to Monte Carlo simulations, this difference corresponds to processes where the mass of the dissociating system is less than 3.4 GeV. These processes should mainly originate from the hadronisation of the GW eigenstates. As a rule, particles coming from the fragmentation of a low-mass system (M diss∼2–3 GeV produced around mean rapidity y∼8) are spread out over a |η|∼1.5 rapidity interval, that is, just down to η=6–6.5 starting from the rapidity y p =8.9 of an incoming 3.5 TeV proton.

  3. Besides the constant slope, we insert the π-loop contribution as proposed in [29], implemented as in [30].

  4. Here we neglect the size, Δb h , of the ‘hard matrix element’ in comparison with the size of the proton. That is, we assume that the amplitude of the hard subprocess is point-like in b-space.

  5. In processes mediated by γ-exchange, the value of S 2 is not small, since the process is dominated by contributions from large distances in impact parameter space, b. Nevertheless the corrections are not negligible, and may be about 20–30 %.

  6. This is because at larger optical density Ω we have a larger probability of interactions.

References

  1. G. Antchev et al. (TOTEM Collaboration), Europhys. Lett. 95, 41001 (2011)

    Article  ADS  Google Scholar 

  2. G. Antchev et al. (TOTEM Collaboration), Europhys. Lett. 96, 21002 (2011)

    Article  ADS  Google Scholar 

  3. G. Antchev et al. (TOTEM Collaboration), Europhys. Lett. 101, 21002 (2013)

    Article  ADS  Google Scholar 

  4. G. Antchev et al. (TOTEM Collaboration), Europhys. Lett. 101, 21004 (2013)

    Article  ADS  Google Scholar 

  5. G. Aad et al. (ATLAS Collaboration), Eur. Phys. J. C 72, 1926 (2012)

    Article  ADS  Google Scholar 

  6. CMS PAS FSQ-12-005

  7. B. Abelev et al. (ALICE Collaboration), arXiv:1208.4968 [hep-ex]

  8. A. Donnachie, P.V. Landshoff, Phys. Lett. B 296, 227 (1992)

    Article  ADS  Google Scholar 

  9. A.B. Kaidalov, Sov. J. Nucl. Phys. 13, 226 (1971)

    Google Scholar 

  10. M.G. Ryskin, A.D. Martin, V.A. Khoze, Eur. Phys. J. C 71, 1617 (2011)

    Article  ADS  Google Scholar 

  11. For a recent review see E. Gotsman, arXiv:1304.7627 [hep-ph] and references therein

  12. U. Maor, arXiv:1305.0299 [hep-ph]

  13. S. Ostapchenko, Phys. Rev. D 81, 114028 (2010)

    Article  ADS  Google Scholar 

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

    Article  ADS  Google Scholar 

  15. L. Baksay et al., Phys. Lett. B 53, 484 (1975)

    Article  ADS  Google Scholar 

  16. R. Webb et al., Phys. Lett. B 55, 331 (1975)

    Article  ADS  Google Scholar 

  17. L. Baksay et al., Phys. Lett. B 61, 405 (1976)

    Article  ADS  Google Scholar 

  18. H. de Kerret et al., Phys. Lett. B 63, 477 (1976)

    Article  ADS  Google Scholar 

  19. G.C. Mantovani et al., Phys. Lett. B 64, 471 (1976)

    Article  ADS  Google Scholar 

  20. A.B. Kaidalov, Phys. Rep. 50, 157 (1979)

    Article  ADS  Google Scholar 

  21. N. Kwak et al., Phys. Lett. B 58, 233 (1975)

    Article  ADS  Google Scholar 

  22. U. Amaldi et al., Phys. Lett. B 66, 390 (1977)

    Article  ADS  Google Scholar 

  23. L. Baksay et al., Nucl. Phys. B 141, 1 (1978)

    Article  ADS  Google Scholar 

  24. B.L. Ioffe, V.S. Fadin, L.N. Lipatov, Quantum Chromodynamics: Perturbative and Nonperturbative Aspects (Cambridge University Press, Cambridge, 2010)

    Book  MATH  Google Scholar 

  25. M.G. Ryskin, A.D. Martin, V.A. Khoze, A.G. Shuvaev, J. Phys. G 36, 093001 (2009)

    Article  ADS  Google Scholar 

  26. F.E. Low, Phys. Rev. D 12, 163 (1975)

    Article  ADS  Google Scholar 

  27. S. Nussinov, Phys. Rev. Lett. 34, 1286 (1975)

    Article  ADS  Google Scholar 

  28. T. Sjostrand, S. Mrenna, P.Z. Skands, Comput. Phys. Commun. 178, 852 (2008). arXiv:0710.3820 [hep-ph]

    Article  ADS  Google Scholar 

  29. A.A. Anselm, V.N. Gribov, Phys. Lett. B 40, 487 (1972)

    Article  ADS  Google Scholar 

  30. V.A. Khoze, A.D. Martin, M.G. Ryskin, Eur. Phys. J. C 18, 167 (2000)

    Article  ADS  Google Scholar 

  31. M.G. Ryskin, A.D. Martin, V.A. Khoze, Eur. Phys. J. C 72, 1937 (2012)

    Article  ADS  Google Scholar 

  32. V.A. Khoze et al., Eur. Phys. J. C 69, 85 (2010)

    Article  ADS  Google Scholar 

  33. M.G. Ryskin, A.D. Martin, V.A. Khoze, Eur. Phys. J. C 60, 265 (2009)

    Article  ADS  Google Scholar 

  34. E.G.S. Luna, V.A. Khoze, A.D. Martin, M.G. Ryskin, Eur. Phys. J. C 59, 1 (2009)

    Article  ADS  Google Scholar 

  35. TOTEM Collaboration, Europhys. Lett. 96, 21002 (2011)

    Article  ADS  Google Scholar 

  36. UA4 Collaboration, Phys. Lett. B 147, 385 (1984)

    Article  Google Scholar 

  37. UA4/2 Collaboration, Phys. Lett. B 316, 448 (1993)

    Article  Google Scholar 

  38. UA1 Collaboration, Phys. Lett. B 128, 336 (1982)

    Google Scholar 

  39. E710 Collaboration, Phys. Lett. B 247, 127 (1990)

    Article  Google Scholar 

  40. CDF Collaboration, Phys. Rev. D 50, 5518 (1994)

    Article  Google Scholar 

  41. N. Kwak et al., Phys. Lett. B 58, 233 (1975)

    Article  ADS  Google Scholar 

  42. U. Amaldi et al., Phys. Lett. B 66, 390 (1977)

    Article  ADS  Google Scholar 

  43. L. Baksay et al., Nucl. Phys. B 141, 1 (1978)

    Article  ADS  Google Scholar 

  44. U. Amaldi et al., Nucl. Phys. B 166, 301 (1980)

    Article  ADS  Google Scholar 

  45. M. Bozzo et al., Phys. Lett. B 155, 197 (1985)

    Article  ADS  Google Scholar 

  46. V.M. Abazov et al. (D0 Collaboration), Phys. Rev. D 86, 012009 (2012)

    Article  ADS  Google Scholar 

  47. J. Orear et al., Phys. Rev. 152, 1162 (1966)

    Article  ADS  Google Scholar 

  48. M. Ciafaloni, D. Colferai, G. Salam, Phys. Rev. D 60, 114036 (1999)

    Article  ADS  Google Scholar 

  49. G. Salam, J. High Energy Phys. 9807, 019 (1998)

    Article  ADS  Google Scholar 

  50. G. Salam, Acta Phys. Pol. B 30, 3679 (1999)

    ADS  Google Scholar 

  51. V.A. Khoze, A.D. Martin, M.G. Ryskin, W.J. Stirling, Phys. Rev. D 70, 074013 (2004)

    Article  ADS  Google Scholar 

  52. A. Donnachie, P.V. Landshoff, Nucl. Phys. B 231, 189 (1984)

    Article  ADS  Google Scholar 

  53. S. Chatrchyan et al. (CMS Collaboration), arXiv:1305.5596 [hep-ex]

  54. LHCb Collaboration, J. Phys. G 40, 045001 (2013)

    Article  Google Scholar 

  55. A.B. Kaidalov, V.A. Khoze, Y.F. Pirogov, N.L. Ter-Isaakyan, Phys. Lett. B 45, 493 (1973)

    ADS  Google Scholar 

Download references

Acknowledgements

We thank Lucian Harland-Lang for discussions. MGR thanks the IPPP at the University of Durham for hospitality. This work was supported by the grant RFBR 11-02-00120-a and by the Federal Program of the Russian State RSGSS-4801.2012.2.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to V. A. Khoze.

Appendices

Appendix A: Observables in terms of GW eigenstates

In the Good–Walker approach [14], low-mass diffractive dissociation is described in terms of so-called diffractive (or GW) eigenstates, |ϕ i 〉 with i=1,n, that diagonalise the T-matrix, and so only undergo ‘elastic’ scattering. On the other hand high-mass dissociation is described in terms of multi-pomeron diagrams. We discuss our treatment of high-mass dissociation in Appendix B.

In this appendix we recall the GW formalism. First, the incoming ‘beam’ proton wave function is written as a superposition of the diffractive eigenstates:

$$ |p\rangle= \sum a_i |\phi_i\rangle, $$
(19)

and similarly for the incoming ‘target’ proton. In this paper we use two diffractive eigenstates, i=1,2. In terms of this 2-channel eikonal model, the pp elastic cross section has the form

$$\begin{aligned} &{\frac{d\sigma_{\mathrm{el}}}{dt} = \frac{1}{4\pi} \biggl \vert \int d^2b\,e^{i \boldsymbol {q}_t \cdot \boldsymbol {b}} \sum_{i,k}|a_i|^2 |a_k|^2\bigl(1-e^{-\varOmega _{ik}(b)/2}\bigr) \biggr \vert ^2,} \\ \end{aligned}$$
(20)

where \(-t=q_{t}^{2}\), and the opacity Ω ik (b) corresponds to one-pomeron exchange between states ϕ i and ϕ k written in the b-representation. Also we have

$$\begin{aligned} &{\sigma_{\mathrm{el}} = \int d^2b \,\biggl \vert \sum_{i,k}|a_i|^2 |a_k|^2 \bigl(1-e^{-\varOmega_{ik}(b)/2}\bigr) \biggr \vert ^2,} \end{aligned}$$
(21)
$$\begin{aligned} &{\sigma_{\mathrm{tot}} = 2\int d^2b\,\sum _{i,k}|a_i|^2 |a_k|^2 \bigl(1-e^{-\varOmega _{ik}(b)/2}\bigr)} \end{aligned}$$
(22)

and the ‘total’ low-mass diffractive cross section

$$ \sigma_{\rm el+SD+DD} = \int d^2b ~\sum _{i,k}|a_i|^2 |a_k|^2 \bigl \vert \bigl(1-e^{-\varOmega_{ik}(b)/2}\bigr) \bigr \vert ^2, $$
(23)

where SD includes the single dissociation of both protons. So the low-mass diffractive dissociation cross section is

$$ \sigma^{\rm D}_{\mathrm{low}M} = \sigma_{\rm el+SD+DD}- \sigma_{\mathrm{el}}, $$
(24)

where \(\sigma_{\rm el+SD+DD}\) corresponds to all possible low-mass dissociation caused by the dispersion of the Good–Walker eigenstate scattering amplitudes. As mentioned in footnote 2, this corresponds to M diss∼2–3 GeV for the TOTEM data [14].

Appendix B: Formulae for high-mass dissociation

The process ppX+p, where one proton dissociates into a system X of high-mass M is conventionally studied in terms of the triple-pomeron coupling, shown as the dot between the dashed lines in Fig. 6(a). In the absence of absorptive corrections, the corresponding cross section is given by

$$\begin{aligned} &{\frac{M^2 \,d\sigma}{dt\,dM^2}} \\ &{\quad = g_{3P}(t)\beta(0)\beta^2(t) \biggl( \frac {s}{M^2} \biggr)^{2\alpha(t)-2} \biggl(\frac{M^2}{s_0} \biggr)^{\alpha(0)-1},} \end{aligned}$$
(25)

where β(t) is the coupling of the pomeron to the proton and g 3P (t) is the triple-pomeron coupling. The coupling g 3P is obtained from a fit to lower energy data. Mainly it is the data on proton dissociation taken at the CERN-ISR with energies from 23.5→62.5 GeV.

Fig. 6
figure 6

(a) A schematic diagram showing the notation of the impact parameters arising in the calculation of the screening corrections to the triple-pomeron contributions to the cross section; (b) a symbolic diagram of multi-pomeron effects

The problem, in the above determination of g 3P , is that this is an effective vertex with coupling

$$ g_{\mathrm{eff}} = g_{3P}*S^2 $$
(26)

which already includes the suppression S 2—the probability that no other secondaries, simultaneously produced in the same pp interaction, populate the rapidity gap region. Recall that the survival factor S 2 depends on the energy of the collider. Since the opacity Ω increases with energy, the number of multiple interactions, NΩ, grows,Footnote 6 leading to a smaller S 2. Thus, we have to expect that the naive triple-pomeron formula with the coupling [20, 55], measured at relatively low collider energies will appreciably overestimate the cross section for high-mass dissociation at the LHC. A more precise analysis [34] accounts for the survival effect \(S^{2}_{\mathrm{eik}}\) caused by the eikonal rescattering of the fast ‘beam’ and ‘target’ partons. In this way, a coupling g 3P about a factor of 3 larger than g eff is obtained, namely g 3P ≃0.2g N , where g N is the coupling of the pomeron to the proton. The analysis of Ref. [34] enables us to better take account of the energy dependence of \(S^{2}_{\mathrm{eik}}\).

To account for the absorptive effect, it is easier to work in the impact parameter, b, representation. To do this we follow the procedure of Ref. [34]. We first take Fourier transforms with respect to the impact parameters specified in Fig. 6(a). Then (25) becomes

$$\begin{aligned} \frac{M^2 \,d\sigma_{ik}}{dt\,dM^2} =& A\int\frac{d^2b_2}{2\pi}e^{i\boldsymbol {q}_t \cdot \boldsymbol {b}_2} F_i(b_2)\int\frac{d^2b_3}{2\pi}e^{i\boldsymbol {q}_t \cdot \boldsymbol {b}_3} F_i(b_3) \\ &{}\times \int\frac{d^2b_1}{2\pi} F_k(b_1), \end{aligned}$$
(27)

where F i (b) is described by the opacity corresponding to the interaction of eigenstate ϕ i with an intermediate parton placed at the position of the triple-pomeron vertex, while F k (b) describes the opacity of eigenstate ϕ k from the proton which dissociates and interacts with the same intermediate parton. After integrating (27) over t, the cross section becomes

$$\begin{aligned} &{\frac{M^2 \,d\sigma_{ik}}{dM^2}} \\ &{\quad = A\int\frac{d^2b_2}{\pi}\int\frac {d^2b_1}{2\pi} \big|F_i(b_2)\big|^2 F_k(b_1) \cdot S_{ik}^2(\boldsymbol {b}_2-\boldsymbol {b}_1),} \end{aligned}$$
(28)

where here we have included the screening correction \(S_{ik}^{2}\), which depends on the separation in impact parameter space, (b 2b 1), of states ϕ i ,ϕ k coming from the incoming protons

$$ S_{ik}^2(\boldsymbol {b}_2-\boldsymbol {b}_1)~ \equiv~\exp\bigl(-\varOmega_{ik}(\boldsymbol {b}_2- \boldsymbol {b}_1)\bigr). $$
(29)

If we now account for more complicated multi-pomeron vertices, coupling m to n pomerons, and assume an eikonal form of the vertex with coupling

$$ g^m_n=(g_N\lambda)^{m+n-2}, $$
(30)

then we have to replace F i by the eikonal elastic amplitude and F k by the inelastic interaction probability. That is, instead of F i =Ω i (b 2) and F k =Ω k (b 1), we put

$$ F_i \to2\bigl(1-e^{-\varOmega_i(b_2)/2}\bigr),\qquad F_k \to\bigl(1-e^{-\varOmega_k(b_1)}\bigr). $$
(31)

Figure 6(b) symbolically indicates multi-pomeron couplings. In (30), g N is the proton-pomeron coupling and λ determines the strength of the triple-pomeron coupling.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Khoze, V.A., Martin, A.D. & Ryskin, M.G. Diffraction at the LHC. Eur. Phys. J. C 73, 2503 (2013). https://doi.org/10.1140/epjc/s10052-013-2503-x

Download citation

  • Received:

  • Revised:

  • Published:

  • DOI: https://doi.org/10.1140/epjc/s10052-013-2503-x

Keywords

Navigation