Skip to main content

Beyond Supergravity in AdS-CFT: An Application to Jet Quenching

  • Chapter
  • First Online:
Modifications of Einstein's Theory of Gravity at Large Distances

Part of the book series: Lecture Notes in Physics ((LNP,volume 892))

  • 1987 Accesses

Abstract

These notes are dedicated to the study of jet quenching in the strongly coupled limit using gauge/string duality. We are interested in corrections to the infinite coupling λ =  result for the jet stopping, in powers of λ −1∕2. To estimate these corrections we need to go beyond supergravity in AdS-CFT, and include all higher-derivative corrections to the supergravity action which arise from the string α′ expansion. For the particular type of “jets” that we study, the expansion in λ −1∕2 is well behaved for jets whose stopping distance stop is in the range \(\lambda ^{-1/6}\ell_{\mathrm{max}} \ll \ell_{\mathrm{stop}} \lesssim \ell_{\mathrm{max}}\), but the expansion breaks down for jets created in such a way that \(\ell_{\mathrm{stop}} \ll \lambda ^{-1/6}\ell_{\mathrm{max}}\). The reason for the breakdown of the λ −1∕2 expansion is caused by the excitation of massive string states. In particular, consider “jets” which are dual to high-momentum gravitons. In the black brane background the gravitons, which are closed string states, get stretched into relatively large classical strings by tidal forces. These stringy excitations of the graviton are not contained in the supergravity approximation, but the jet stopping problem can nonetheless still be solved by drawing on various string-theory methods (the eikonal approximation, the Penrose limit, string quantization in pp-wave backgrounds) to obtain a probability distribution for the late-time classical string loops.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 99.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 129.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    In a weak-coupling analysis, the two running couplings relevant to jet stopping are, roughly, α s(T) and α s(Q  ⊥ ), where \(Q_{\perp }\sim (\hat{q}E)^{1/4}\) grows slowly with energy and is the scale of the typical relative momentum between two daughter partons when a high energy parton splits through hard bremsstrahlung or pair production. (\(\hat{q} \sim \alpha _{\mathrm{s}}^{2}T^{3}\) is a scale characteristic of the plasma that parametrizes transverse momentum diffusion of high-energy partons.) A third limiting case of interest, not addressed here, is where α s(T) is large but α s(Q  ⊥ ) is small. See, for example, Liu, Rajagopal, and Wiedemann [2].

  2. 2.

    See Sin and Zahed [9] for the earliest attempt we are aware of to discuss jet stopping in the context of gauge-gravity duality. See also [10]. In our work, we will only consider analogs of light-particle jets and will not study the heavy-particle case. For λ =  analysis of the latter, see, for example, [11, 12] and references therein.

  3. 3.

    For a discussion of one potential source of 1∕N c corrections to jet propagation, see Shuryak, Yee, and Zahed [14].

  4. 4.

    At a technical level, we define where the jet stops [6, 7] following Chesler et al. [5, 16] as the location where the jet’s energy and momentum and charge first begin to evolve hydrodynamically. Since hydrodynamics is an effective theory only on distance scales ≫ 1∕T at strong coupling, it does not make sense to apply this definition to stopping distances small compared to 1∕T.

  5. 5.

    For a nice summary of higher-dimensional gravitational corrections in Type II supergravity generated by tree-level string amplitudes (i.e. in the N c =  limit), see Table 1 of Stieberger [17]. Though not relevant to the N c =  case we are discussing, a nice discussion of corrections generated from one-loop string amplitudes may be found in Richards [18].

  6. 6.

    See [7] for a discussion in the context of the present paper, but this correspondence is implicit in the earlier work of [35].

  7. 7.

    For more discussion of why the distance the excitation travels before falling into the horizon should be identified with the stopping distance in the 3 + 1 dimensional field theory problem, see the discussion in [7], as well as earlier discussions in the context of falling classical strings [3, 16].

  8. 8.

    The coordinate used in [6, 7] is u = z 2z h 2, which is u = (z∕2)2 when working in the units 2π T = 1 used there.

  9. 9.

    It is important to note that the masses of five-dimensional fields in the gravity dual have nothing to do with the masses of four-dimensional excitations in the \(\mathcal{N}=4\) SYM field theory. The five-dimensional mass m is not the “mass of a jet.”

  10. 10.

    See [7] for this explicit result, but the parametric behavior stop ∼ (E 2∕−q 2)1∕4, within its range of validity, was found earlier by Hatta, Iancu and Mueller [4].

  11. 11.

    For a very brief summary of the relevant scales for the coupling, see, for example, [24].

  12. 12.

    See, for example, Table 7 of the review by D’Hoker and Freedman [26]. Here X k is shorthand for any symmetric product \(X^{(i_{1}}X^{i_{2}}\cdots X^{i_{k})}\) of k factors of the three complex scalar fields X 1, X 2, X 3 of \(\mathcal{N}=4\) SYM.

  13. 13.

    Equation (14.24) is nicely summarized in Eqs. (3.1–3) of [27] and originates from [15, 28].

  14. 14.

    This is a statement about the uncorrected, i.e. λ = , (AdS5-Schwarzschild) × S 5 background, and does not account for corrections to that background due to α3 C 4. But this is good enough for figuring out the leading correction to the ϕ equation of motion.

  15. 15.

    See [7] for a λ =  discussion of when the wave packet is small enough to treat as a particle. The summary is that L can be chosen appropriately so that everything is fine at z ∼ z when by convolving with an appropriate localized boundary source function stop ≪  max.

  16. 16.

    \((C_{\mathsf{0101}},C_{\mathsf{1212}},C_{\mathsf{0505}},C_{\mathsf{1515}}) = (f, 1,-3,-f^{-1}) \times \mathbf{R}^{2}/z_{\mathrm{h}}^{4}\), with all other components determined by symmetry.

  17. 17.

    Penrose limits have previously been studied in AdS5-Schwarzschild by Pando Zayas and Sonnenschein [35], but the null geodesic studied was different. Their geodesic fell straight toward the horizon in AdS5-Schwarzschild, corresponding to \(\boldsymbol{q}=0\) in our problem rather than \(\vert \boldsymbol{q}\vert \simeq E\). Their geodesics also have non-trivial motion on the 5-sphere S 5. In our application, no dynamical evolution of the S 5 degrees of freedom takes place, and the S 5 string degrees of freedom simply remain in a quantum state given by the S 5 harmonic of the supergravity field of interest.

  18. 18.

    This u should not be confused with the coordinate u ≡ (zz h)2 used in earlier work by some of the authors [6, 7].

  19. 19.

    For example, at late times the exponential in the wavepacket (14.98) becomes exp[iS], where S ∝ x 2∕(−τ). The WKB condition |  x 2 S | ≪ ( x S)2 is satisfied as τ → 0 for x ∝ (−τ)−1∕3. We will see shortly that the proportionality constant in (14.111) is of order 1 for the modes \(n \lesssim n_{\star }\) of interest. If one keeps track parametrically of all the proportionality constants in the exponential exp[iS], one finds more specifically that the WKB condition is satisfied when \(-\bar{\tau }\ll 1\) (i.e. \(\bar{z} \gg 1\)).

  20. 20.

    For numerical work, it is mildly convenient to eliminate \(\bar{\tau }\) and express all of the relevant equations solely in terms of \(\bar{z}\), giving \(\bar{z}^{4}(1 +\bar{ z}^{4})\chi '' + 2\bar{z}^{3}(1 + 2\bar{z}^{4})\chi ' = -4(\xi ^{6} -\bar{ z}^{6})\chi\) and \(\chi '(\bar{z}_{0}) = 2i\xi ^{3}/\bar{z}_{0}^{2}\) (with \(\bar{z}_{0} \rightarrow 0\)) and \(\vert \chi (\bar{z})\vert \rightarrow 3^{1/3}\,C(\xi )\,\bar{z}\) (as \(\bar{z} \rightarrow \infty \)).

  21. 21.

    More precisely, Gubser et al. first considered a folded open string that stretched out from beyond the horizon, as in Fig. 1 of [3]. But in actual calculations, they focused on the trailing infinite folded string, as in Fig. 2 of that reference.

  22. 22.

    We ignore here logarithmic energy dependence in the prefactor of the exponential. The small-λ scaling is really (E∕lnE)1∕2, which is equivalent to including a log-of-log energy dependence in the exponent f: \((E/\ln E)^{1/2} =\exp [\tfrac{1} {2} -\tfrac{1} {2}\ln \ln E]\).

References

  1. P.B. Arnold, S. Cantrell, W. Xiao, Stopping distance for high energy jets in weakly-coupled quark-gluon plasmas. Phys. Rev. D 81, 045017 (2010) [arXiv:0912.3862 [hep-ph]]

    Google Scholar 

  2. H. Liu, K. Rajagopal, U.A. Wiedemann, Calculating the jet quenching parameter from AdS/CFT. Phys. Rev. Lett. 97, 182301 (2006) [hep-ph/0605178]

    Google Scholar 

  3. S.S. Gubser, D.R. Gulotta, S.S. Pufu, F.D. Rocha, Gluon energy loss in the gauge-string duality. JHEP 0810, 052 (2008) [arXiv:0803.1470 [hep-th]]

    Google Scholar 

  4. Y. Hatta, E. Iancu, A.H. Mueller, Jet evolution in the \(\mathcal{N}=4\) SYM plasma at strong coupling. JHEP 0805, 037 (2008) [arXiv:0803.2481 [hep-th]]

    Google Scholar 

  5. P.M. Chesler, K. Jensen, A. Karch, L.G. Yaffe, Light quark energy loss in strongly-coupled \(\mathcal{N}=4\) supersymmetric Yang-Mills plasma. Phys. Rev. D 79, 125015 (2009) [arXiv:0810.1985 [hep-th]]

    Google Scholar 

  6. P. Arnold, D. Vaman, Jet quenching in hot strongly coupled gauge theories revisited: 3-point correlators with gauge-gravity duality. JHEP 1010, 099 (2010) [arXiv:1008.4023 [hep-th]]

    Google Scholar 

  7. P. Arnold, D. Vaman, Jet quenching in hot strongly coupled gauge theories simplified. JHEP 1104, 027 (2011) [arXiv:1101.2689 [hep-th]]

    Google Scholar 

  8. P.M. Chesler, Y.-Y. Ho, K. Rajagopal, Shining a gluon beam through quark-gluon plasma. Phys. Rev. D 85, 126006 (2012) [arXiv:1111.1691 [hep-th]]

    Google Scholar 

  9. S.-J. Sin, I. Zahed, Holography of radiation and jet quenching. Phys. Lett. B 608, 265 (2005) [hep-th/0407215]

    Article  ADS  Google Scholar 

  10. M. Spillane, A. Stoffers, I. Zahed, Jet quenching in shock waves. JHEP 1202, 023 (2012) [arXiv:1110.5069 [hep-th]]; A. Stoffers, I. Zahed, arXiv:1110.2943 [hep-th]

    Google Scholar 

  11. C.P. Herzog, A. Karch, P. Kovtun, C. Kozcaz, L.G. Yaffe, Energy loss of a heavy quark moving through N = 4 supersymmetric Yang-Mills plasma. JHEP 0607, 013 (2006) [hep-th/0605158]

    Article  ADS  MathSciNet  Google Scholar 

  12. J. Casalderrey-Solana, D. Teaney, Heavy quark diffusion in strongly coupled \(\mathcal{N}=4\) Yang-Mills. Phys. Rev. D 74, 085012 (2006) [hep-ph/0605199]

    Google Scholar 

  13. P. Arnold, P. Szepietowski, D. Vaman, Coupling dependence of jet quenching in hot strongly-coupled gauge theories. JHEP 1207, 024 (2012) [arXiv:1203.6658 [hep-th]]

    Google Scholar 

  14. E. Shuryak, H.-U. Yee, I. Zahed, Self-force and synchrotron radiation in odd space-time dimensions. Phys. Rev. D 85, 104007 (2012) [arXiv:1111.3894 [hep-th]]

    Google Scholar 

  15. D.J. Gross, E. Witten, Superstring modifications of Einstein’s equations. Nucl. Phys. B 277, 1 (1986)

    Article  ADS  MathSciNet  Google Scholar 

  16. P.M. Chesler, K. Jensen, A. Karch, Jets in strongly-coupled \(\mathcal{N}=4\) super Yang-Mills theory. Phys. Rev. D 79, 025021 (2009) [arXiv:0804.3110 [hep-th]]

    Google Scholar 

  17. S. Stieberger, Constraints on tree-level higher order gravitational couplings in superstring theory. Phys. Rev. Lett.  106, 111601 (2011) [arXiv:0910.0180 [hep-th]]

    Google Scholar 

  18. D.M. Richards, The one-loop five-graviton amplitude and the effective action. JHEP 0810, 042 (2008) [arXiv:0807.2421 [hep-th]]

    Google Scholar 

  19. G. D’Appollonio, P. Di Vecchia, R. Russo, G. Veneziano, High-energy string-brane scattering: Leading eikonal and beyond. JHEP 1011, 100 (2010) [arXiv:1008.4773 [hep-th]]

    Google Scholar 

  20. G. Papadopoulos, J.G. Russo, A.A. Tseytlin, Solvable model of strings in a time dependent plane wave background. Class. Quant. Grav. 20, 969 (2003) [hep-th/0211289]

    Article  ADS  MATH  MathSciNet  Google Scholar 

  21. G.T. Horowitz, A.R. Steif, Space-time singularities in string theory. Phys. Rev. Lett. 64, 260 (1990); Strings in strong gravitational fields. Phys. Rev. D 42, 1950 (1990)

    Google Scholar 

  22. E.G. Gimon, L.A. Pando Zayas, J. Sonnenschein, Penrose limits and RG flows. JHEP 0209, 044 (2002) [hep-th/0206033]

    Article  ADS  MathSciNet  Google Scholar 

  23. H.J. de Vega, N.G. Sanchez, Strings falling into space-time singularities. Phys. Rev. D 45, 2783 (1992); H.J. de Vega, M. Ramon Medrano, N.G. Sanchez, Classical and quantum strings near space-time singularities: Gravitational plane waves with arbitrary polarization. Class. Quant. Grav. 10, 2007 (1993)

    Google Scholar 

  24. P. Arnold, D. Vaman, Some new results for ‘jet’ stopping in AdS/CFT. arXiv:1106.1680, An abridged version appears in J. Phys. G 38, 124175 (2011)

    Google Scholar 

  25. N. Armesto, J.D. Edelstein, J. Mas, Jet quenching at finite ‘t Hooft coupling and chemical potential from AdS/CFT. JHEP 0609, 039 (2006) [hep-ph/0606245]

    Google Scholar 

  26. E. D’Hoker, D.Z. Freedman, (author) Supersymmetric gauge theories and the AdS / CFT correspondence, in Strings, Branes and Extra Dimensions: TASI 2001: Proceedings, ed. by Steven S. Gubser, Joseph D. Lykken (River Edge, NJ, World Scientific, 2004) [hep-th/0201253]

    Google Scholar 

  27. A. Buchel, J.T. Liu, A.O. Starinets, Coupling constant dependence of the shear viscosity in \(\mathcal{N}=4\) supersymmetric Yang-Mills theory. Nucl. Phys. B 707, 56 (2005) [hep-th/0406264]

    Article  ADS  MATH  MathSciNet  Google Scholar 

  28. M.T. Grisaru, D. Zanon, Sigma model superstring corrections to the Einstein-Hilbert action. Phys. Lett. B 177, 347 (1986); M.D. Freeman, C.N. Pope, M.F. Sohnius, K.S. Stelle, Higher order sigma model counterterms and the effective action for superstrings. Phys. Lett. B 178, 199 (1986); Q-H. Park, D. Zanon, More on σ-model β functions and low-energy effective actions. Phys. Rev. D 35, 4038 (1987)

    Google Scholar 

  29. O. Aharony, S.S. Gubser, J.M. Maldacena, H. Ooguri, Y. Oz, Large N field theories, string theory and gravity. Phys. Rept. 323, 183 (2000) [hep-th/9905111]

    Article  ADS  MathSciNet  Google Scholar 

  30. M.B. Green, P. Vanhove, Low-energy expansion of the one loop type II superstring amplitude. Phys. Rev. D 61, 104011 (2000) [hep-th/9910056]

    Article  ADS  MathSciNet  Google Scholar 

  31. D. Amati, M. Ciafaloni, G. Veneziano, Superstring collisions at planckian energies. Phys. Lett. B 197, 81 (1987); Higher order gravitational deflection and soft bremsstrahlung in planckian energy superstring collisions. Nucl. Phys. B 347, 550 (1990). Effective action and all order gravitational eikonal at Planckian energies. Nucl. Phys. B 403, 707 (1993); D.N. Kabat, M. Ortiz, Eikonal quantum gravity and planckian scattering. Nucl. Phys. B 388, 570 (1992) [hep-th/9203082]; L. Cornalba, M.S. Costa, J. Penedones, R. Schiappa, Eikonal approximation in AdS/CFT: from shock waves to four-point functions. JHEP 0708, 019 (2007) [hep-th/0611122]; L. Cornalba, M.S. Costa, J. Penedones, Eikonal approximation in AdS/CFT: resumming the gravitational loop expansion. JHEP 0709, 037 (2007) [arXiv:0707.0120 [hep-th]]; S.B. Giddings, R.A. Porto, The gravitational S-matrix. Phys. Rev. D 81, 025002 (2010) [arXiv:0908.0004 [hep-th]]

    Google Scholar 

  32. M. Blau, J.M. Figueroa-O’Farrill, G. Papadopoulos, Penrose limits, supergravity and brane dynamics. Class. Quant. Grav. 19, 4753 (2002) [hep-th/0202111]

    Google Scholar 

  33. M. Blau, Penrose waves and penrose limits. Lecture given at 2004 Saalburg/Wolfesdorf Summer School, http://www.blau.itp.unibe.ch/Lecturenotes.html

  34. M. Blau, D. Frank, S. Weiss, Fermi coordinates and Penrose limits. Class. Quant. Grav. 23, 3993 (2006) [hep-th/0603109]

    Article  ADS  MATH  MathSciNet  Google Scholar 

  35. L.A. Pando Zayas, J. Sonnenschein, On Penrose limits and gauge theories. JHEP 0205, 010 (2002) [hep-th/0202186]

    Article  ADS  Google Scholar 

  36. S.P. Kim, C.H. Lee, Nonequilibrium quantum dynamics of second order phase transitions. Phys. Rev. D 62, 125020 (2000) [hep-ph/0005224]

    Google Scholar 

  37. P. Arnold, P. Szepietowski, D. Vaman, G. Wong, Tidal stretching of gravitons into classical strings: application to jet quenching with AdS/CFT. JHEP 1302, 130 (2013) [arXiv:1212.3321 [hep-th]]

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Diana Vaman .

Editor information

Editors and Affiliations

Appendices

Appendix 1: What Happens for \(\boldsymbol{z \gg z_{\star }}\)?

By making the reasonable assumptions that we outlined in Sect. 14.2.3, we have managed to analyze the question of when corrections become important by focusing on particle trajectories at z ∼ z . As z increases beyond this scale, the forward progress of the trajectory slows to a stop.We previously asserted that at some scale z bad ≫ z the expansion in higher-derivative corrections would eventually break downeven if the expansion was well behaved at z .

Start by considering the cost of an αD 2 factor, which we analyzed in Sect. 14.4.3 for z ∼ z . At larger z, with \(z_{\star } \lesssim z \ll z_{\mathrm{h}}\), (14.62) gives

$$\displaystyle{ q_{\mathsf{5}}\bigr |_{z\gtrsim z_{\star }} \sim \frac{z^{2}E} {z_{\mathrm{h}}^{2}}. }$$
(14.140)

Then the cost (14.61) of each αD 2 factor is

$$\displaystyle{ \alpha 'D^{2}\vert _{ z\gtrsim z_{\star }} \sim \frac{\alpha 'zq_{\mathsf{5}}} {\mathbf{R}^{2}} \sim \frac{z^{3}ET^{2}} {\lambda ^{1/2}} \,. }$$
(14.141)

This cost is unsuppressed for \(z \gtrsim z_{\mathrm{bad}}\), where

$$\displaystyle{ z_{\mathrm{bad}} \sim \frac{\lambda ^{1/6}} {E^{1/3}T^{2/3}}\,. }$$
(14.142)

The same constraint arises from the other important corrections that we analyzed. For instance, the cost of adding an α2 D 2 C factor was z 6 E 2λ z h 4, which also becomes unsuppressed at the same \(z \gtrsim z_{\mathrm{bad}}\).

Note that the requirement \(\ell_{\mathrm{stop}} \gg \lambda ^{-1/6}\ell_{\mathrm{max}}\) for the expansion in corrections to be well-behaved in Fig. 14.3 is the same condition as requiring \(z_{\mathrm{bad}} \gg z_{\star }\).

Appendix 2: Why (14.46) Cannot Precisely Determine \(\boldsymbol{\varDelta \ell_{\mathrm{stop}}}\)

Consider the safe region \(\ell_{\mathrm{stop}} \gg \lambda ^{-1/6}\ell_{\mathrm{max}}\) of Fig. 14.3, where the effects of higher-dimensional supergravity interactions should be suppressed. The R 4 corrections then dominate the corrections at z ∼ z . We might then be tempted to use the explicit form (14.25) of the R 4 correction, combined with the particle-based formula (14.38) for the stopping distance, to explicitly calculate the first correction to the λ =  result (14.18) for the stopping distance. In this appendix, we discuss why that does not work.

We start with (14.45),

$$\displaystyle{ \ell_{\mathrm{stop}} \simeq \int _{0}^{z_{\mathrm{h}} }\mathit{dz} \frac{\vert \boldsymbol{q}\vert \left [1 -\frac{2\varepsilon z^{10}} {z_{\mathrm{h}}^{8}} \vert \boldsymbol{q}\vert ^{2}\right ]} {\sqrt{-q^{2 } + \frac{z^{4 } } {z_{\mathrm{h}}^{4}} \vert \boldsymbol{q}\vert ^{2} + \frac{\varepsilon z^{10}} {z_{\mathrm{h}}^{8}} \vert \boldsymbol{q}\vert ^{4}f}}. }$$
(14.143)

First we explain why the potentially sign-changing behavior of the numerator correction in the R 4-corrected formula (14.143) for the stopping distance could be ignored. The disturbing features of this correction arise in the z range given by (14.48),

$$\displaystyle{ z \gg z_{\mathrm{disturbing}} \sim \left (\frac{\lambda ^{3/4}T} {E} \right )^{1/5}z_{\mathrm{ h}}. }$$
(14.144)

This difficulty only arises at all if z disturbing < z h, which requires

$$\displaystyle{ E \gg \lambda ^{3/4}T. }$$
(14.145)

Now compare (14.144) and (14.142):

$$\displaystyle{ z_{\mathrm{disturbing}} \sim \left ( \frac{E} {\lambda ^{1/8}T}\right )^{2/15}z_{\mathrm{ bad}}. }$$
(14.146)

The inequality (14.145) then gives

$$\displaystyle{ z_{\mathrm{disturbing}} \gg z_{\mathrm{bad}}, }$$
(14.147)

and so the numerator correction cannot be believed in the range of z for which it becomes disturbing.

Dropping the numerator correction from (14.143) leaves

$$\displaystyle{ \ell_{\mathrm{stop}} \simeq \int _{0}^{z_{\mathrm{h}} }\mathit{dz} \frac{\vert \boldsymbol{q}\vert } {\sqrt{-q^{2 } + \frac{z^{4 } } {z_{\mathrm{h}}^{4}} \vert \boldsymbol{q}\vert ^{2} + \frac{\varepsilon z^{10}} {z_{\mathrm{h}}^{8}} \vert \boldsymbol{q}\vert ^{4}f}}, }$$
(14.148)

For simplicity, in what follows we will just analyze the case E ≫ λ 3∕4 T.

In the integrand, look at the expression under the square root in the denominator. The relative importance of the \(\varepsilon z^{10}\) term grows with increasing z. For z ≫ z , the z 4 term under the square root dominates over the − q 2 term, so we should compare the \(\varepsilon z^{10}\) term to the z 4 term. These are the same size at a scale \(z_{\star \star } \gg z_{\star }\) given by

$$\displaystyle{ z_{\star \star } \sim \frac{z_{\mathrm{h}}^{2/3}} {\varepsilon ^{1/6}\vert \boldsymbol{q}\vert ^{1/3}} \sim \frac{\lambda ^{1/4}} {E^{1/3}T^{2/3}} \sim \frac{\lambda ^{1/4}} {l_{\mathrm{max}}T^{2}}\,, }$$
(14.149)

assuming that z ⋆⋆ ≪ z h so that f ≃ 1. But z ⋆⋆ ≪ z h follows from (14.149) and our consideration of E ≫ λ 3∕4 T.

Now calculate the correction Δ ℓ to the stopping distance by subtracting the λ =  result (14.15) from (14.148),

$$\displaystyle{ \varDelta \ell\equiv \ell_{\mathrm{stop}} -\ell_{\mathrm{stop}}^{\lambda =\infty }\simeq \int _{ 0}^{z_{\mathrm{h}} }\mathit{dz}\left [ \frac{\vert \boldsymbol{q}\vert } {\sqrt{-q^{2 } + \frac{z^{4 } } {z_{\mathrm{h}}^{4}} \vert \boldsymbol{q}\vert ^{2} + \frac{\varepsilon z^{10}} {z_{\mathrm{h}}^{8}} \vert \boldsymbol{q}\vert ^{4}f}} - \frac{\vert \boldsymbol{q}\vert } {\sqrt{-q^{2 } + \frac{z^{4 } } {z_{\mathrm{h}}^{4}} \vert \boldsymbol{q}\vert ^{2}}}\right ]\!. }$$
(14.150)

This integral is dominated by z ∼ z ⋆⋆. So, to explicitly calculate Δ ℓ will require trusting the integrand at z ∼ z ⋆⋆ given by (14.149). Compare this to the z scale (14.142) where the expansion in supergravity corrections breaks down:

$$\displaystyle{ z_{\star \star } \sim \lambda ^{1/12}z_{\mathrm{ bad}} \gg z_{\mathrm{bad}}. }$$
(14.151)

So we cannot trust (14.150) in the range of z where we want to use it to get an explicit result for Δ ℓ.

Appendix 3: Other Higher-Derivative Terms

In Sect. 14.4.3, we addressed the αQ 5 D 5 piece of αD 2 when studying the importance of D 2n C 4. We also recycled our conclusion from that analysis when later considering applying extra powers of derivatives to D 2k C 4+k. The dominant terms involved αQ I D I where the D I hits a background Weyl tensor. We motivated focusing on Q 5 D 5 by noting that the background Weyl tensor depends only of the x 5 coordinate. If the D’s were ordinary derivatives instead of covariant derivatives, that would be the end of the story. However, the other components D μ of the covariant derivative do not vanish when applied to the background Weyl tensor. In fact, they are parametrically of order 1∕z, just like D 5 . As a result, for example,

$$\displaystyle{ \alpha 'Q^{\mathsf{3}}D_{\mathsf{ 3}} =\alpha 'Q_{\mathsf{3}}g^{\mathsf{3}\mathsf{3}}D_{\mathsf{ 3}} \sim \alpha '\times E \times \frac{z^{2}} {\mathbf{R}^{2}} \times \frac{1} {z} \sim \frac{Ez} {\lambda ^{1/2}} }$$
(14.152)

is actually parametrically larger than the derivative

$$\displaystyle{ \alpha 'Q^{\mathsf{5}}D_{\mathsf{ 5}} \sim \frac{q_{\mathsf{5}}z} {\lambda ^{1/2}} }$$
(14.153)

considered in the main text (14.61).

So why doesn’t this lead to much larger results for the importance of D 2n R 4 and other operators than shown in Fig. 14.3? Our answer requires thinking about how the indices of the background Weyl tensor C JKLM hit by αQ I D I contract with everything else.

Because C IJKL depends only on x 5, a non-zero value for Q μ D μ C IJKL arises only from the terms of D involving the Christoffel symbols:

$$\displaystyle{ Q^{\mu }D_{\mu }C_{\mathit{IJKL}} = -Q^{\mu }\varGamma ^{\bar{I}}_{ I\mu }C_{\bar{I}\mathit{JKL}} - Q^{\mu }\varGamma ^{\bar{J}}_{ J\mu }C_{I\bar{J}\mathit{KL}} -\cdots \,. }$$
(14.154)

Now write

$$\displaystyle{ \varGamma =\varGamma ^{\mathrm{(AdS)}}+\varDelta \varGamma, }$$
(14.155)

where Γ (AdS) is the zero-temperature, purely AdS expression for the connection Γ. The difference between AdS and AdS5-Schwarzschild is the difference between taking f = 1 and f = 1 − (zz h)4 in the metric (14.6). As a result, the Δ Γ piece of (14.155) is suppressed compared to the Γ (AdS) piece by order (zz h)4. For studying the dominant corrections at z ∼ z  ≪ z h, we should therefore focus on Γ (AdS). In particular,

$$\displaystyle{ \alpha 'Q^{\mu }\varDelta \varGamma ^{\bar{I}}_{ I\mu } \sim \alpha 'E \times \frac{z^{2}} {\mathbf{R}^{2}} \times \frac{1} {z}\left ( \frac{z} {z_{\mathrm{h}}}\right )^{4} \sim \frac{Ez^{5}} {\lambda ^{1/2}z_{\mathrm{h}}^{4}} }$$
(14.156)

is always less important at z ∼ z than the αQ 5 D 5 term (14.153) that we considered in the main text.

So now focus on Γ AdS:

$$\displaystyle{ Q^{\mu }D_{\mu }C_{\mathit{IJKL}} \simeq -Q^{\mu }(\varGamma ^{\bar{I}}_{ I\mu })^{\mathrm{AdS}}C_{\bar{ I}\mathit{JKL}} - Q^{\mu }(\varGamma ^{\bar{J}}_{ J\mu })^{\mathrm{AdS}}C_{ I\bar{J}\mathit{KL}} -\cdots \,. }$$
(14.157)

Because AdS space has four-dimensional Lorentz invariance, the μ index on Q μ above must pass through to contract with something else. For example,

$$\displaystyle\begin{array}{rcl} Q^{\mu }(\varGamma ^{\bar{I}}_{ I\mu })^{\mathrm{AdS}}& & C_{\bar{ I}\mathit{JKL}} \times (\mbox{ other stuff})^{\mathit{IJKL}} \\ & & \simeq \frac{1} {z}\,Q^{\mu }C_{\mu \mathit{JKL}}\times (\mbox{ other stuff})^{\mathsf{5}\mathit{JKL}}-\frac{1} {z}\,C^{\mathsf{5}}_{ \mathit{JKL}}Q_{\mu }\times (\mbox{ other stuff})^{\mu \mathit{JKL}}{}\end{array}$$
(14.158)

and

$$\displaystyle\begin{array}{rcl} Q^{\mu }(\varGamma ^{\bar{J}}_{ J\mu })^{\mathrm{AdS}}& & C_{ I\bar{J}\mathit{KL}} \times (\mbox{ other stuff})^{\mathit{IJKL}} \\ & & \simeq \frac{1} {z}\,Q^{\mu }C_{I\mu \mathit{KL}}\times (\mbox{ other stuff})^{I\mathsf{5}\mathit{KL}} -\frac{1} {z}\,C_{I}^{\mathsf{5}}{}_{ \mathit{KL}}Q_{\mu }\times (\mbox{ other stuff})^{I\mu \mathit{KL}}{}\end{array}$$
(14.159)

But now recall that our dominant terms already had every C contracted with two Q’s. So the “other stuff” above had the form

$$\displaystyle{ (\mbox{ other stuff})^{\mathit{IJKL}} \sim Q^{I}Q^{K}(\mbox{ something})^{\mathit{JL}}, }$$
(14.160)

and these terms were dominant because both Q I and Q K were parametrically of order E when contracted with the Weyl tensor C IJKL . We are currently worried about the possibility that the Q μ factor above is also of order E. Now look at the first term in (14.158). The Q μ Q I Q K  ∼ E 3 is contracted in such a way that it instead gives \(Q_{\mu }Q_{\mathsf{5}}Q_{K} \sim q_{\mathsf{5}}E^{2} \ll E^{3}\), which is not problematical. The second term in (14.158) contracts two Q’s together to give a factor of Q μ η μ ν Q ν  ∼ q 2 instead of an E 2, and so it also is suppressed. Next look at the first term in (14.159). There we have Q μ Q I Q K C I μ KL . Up to terms which are suppressed by q 5  ≪ E, this is the same as Q J Q I Q K C IJKL , which vanishes by the symmetry of the Weyl tensor. Finally, look at the second term of (14.159), which involves

$$\displaystyle{ Q_{\mu }(\mbox{ something})^{\mu L}. }$$
(14.161)

For the dominant terms analyzed in the main text of this paper, the “something” is made up of factors of Q and QQC. If (something)μ L gives a factor of Q μ, then two of our Q’s that were supposed to be giving factors of E will instead give a factor of − q 2 ≪ E. If (something)μ L gives a factor of Q N Q P C N μ P, then we’ll get a suppression as before because of the symmetry of C.

Appendix 4: Large \(\boldsymbol{\xi }\) Behavior of \(\boldsymbol{C(\xi )}\)

For large ξ, the \(\bar{z}^{6}\) term in the differential equation (14.104a) for ξ can be ignored until \(\bar{z} \gg 1\). At that point, however, we may use the simple large-\(\bar{z}\) result (14.110) for \(\bar{z}\). Substituting this into (14.104a) gives

$$\displaystyle{ \frac{d^{2}\chi } {d\bar{\tau }^{2}} = -4\left [\xi ^{6} - \frac{1} {(-3\bar{\tau })^{2}}\right ]\chi, }$$
(14.162)

whose solution is

$$\displaystyle{ \chi = (-\pi \xi ^{3}\bar{\tau })^{1/2}H_{ 5/6}^{(2)}{\bigl ( - 2\xi ^{3}\bar{\tau }\bigr )}. }$$
(14.163)

The late-time behavior τ → 0 is

$$\displaystyle{ \vert \chi \vert \simeq \frac{\varGamma (\tfrac{5} {6})} {\pi ^{1/2}\xi (-\bar{\tau })^{1/3}}\,. }$$
(14.164)
Fig. 14.15
figure 15

(a) Parallel geodesics in AdS5 in the Lorentz frame where the excitation is at rest in 3-space. These geodesics maintain a constant proper separation as they fall into the bulk, and this separation should be thought of as of order the characteristic size (\(\sim \sqrt{\alpha '}\)) of the closed quantum string loop describing the graviton (or other massless string mode). The narrow red loops are meant to be suggestive of the closed string loop. (b) The same picture boosted to the original Lorentz frame. (c) A picture of how those geodesics evolve in AdS5-Schwarzschild rather than AdS5. The early-time behavior is the same as (b). [For classical oscillating string solutions, the strings depicted in (a) may be thought of as snapshots at moments when the string’s proper extent in x 3 is at, say, maximum (or half-maximum or whatever). Such solutions would similarly oscillate in the z ≪ z part of (b) but not in the z ≫ z part, where tidal forces dominate over tension]

Appendix 5: A Back-of-the-Envelope Estimate

In this appendix we give a parametric estimate of the amount of tidal stretching of the string compared to the size of the stopping distance stop. Here the only thing we need to know is that the stopping distance given by following a null geodesic as in Fig. 14.5 is proportional to a power of the slope dx 3dx 5 of that geodesic where it starts, at the boundary. The more downward-directed one starts the trajectory in Fig. 14.5, the less distance it will travel in x 3 before reaching the horizon.

Now interpret the trajectory of Fig. 14.5 as a trajectory for the center of mass of a tiny, falling loop of string. Once the string gets far enough from the boundary that the tidal forces dominate over the string tension, then the string tension becomes ignorable, and different pieces of the string will fall independently along their own geodesics, the string stretching accordingly. Imagine plotting two such geodesics, for the two bits of the string loop that are most separated. The separation of those geodesics is a measure of the extent of the tidally-stretched loop of string as it falls towards the horizon. The proper size of the string should start out of order the quantum mechanical size Σ of the graviton, which is roughly set by dimensional analysis in terms of the string tension \(\mathbb{T}\) as

$$\displaystyle{ \varSigma _{\mathrm{graviton}} \sim \mathbb{T}^{-1/2} \sim \sqrt{\alpha '}, }$$
(14.165)

where \(\alpha ' = 1/2\pi \mathbb{T}\) is the string slope parameter.

Very close to the boundary, the tidal forces due to the black hole are negligible, and the closed loop of string is in its ground state. We can set up our two geodesics above so that, correspondingly, they maintain constant proper separation Σ graviton near the boundary, where the AdS5-Schwarzschild metric approaches a purely AdS5 metric. To see how this works, imagine making a four-dimensional boost from (i) the plasma rest frame, in which we create an excitation with large 4-momentum q μ = (ω, 0, 0, q 3 ) ≃ (E, 0, 0, E) and relatively small 4-virtuality − q 2 ≪ E 2, to (ii) the excitation’s initial rest frame, where the 4-momentum is instead \((\sqrt{-q^{2}},0,0,0)\). The Lorentz boost factor for this transformation is

$$\displaystyle{ \gamma = \sqrt{ \frac{\omega ^{2 } } {-q^{2}}} \simeq \sqrt{ \frac{E^{2 } } {-q^{2}}} \gg 1. }$$
(14.166)

In AdS5, the trajectory in the new frame will drop straight down away from the boundary, as depicted by the dashed line in Fig. 14.15a.

Now consider the graviton as an extended object with proper size Σ. The two straight solid null lines in Fig. 14.15a depict the extent of the graviton in AdS5 in the excitation’s rest frame at early times. In pure AdS5 null geodesics are straight lines. We parametrize the two solid lines of Fig. 14.15a as

$$\displaystyle{ x^{I} ={\bigl (\gamma _{ +},0,0,\pm \beta _{+}\gamma _{+},1\bigr )}\,z }$$
(14.167)

with β + ≪ 1 and γ + ≡ (1 −β + 2)−1∕2 ≃ 1. Because of the warp factor in the metric, these two lines are parallel and maintain constant proper separation \(\sqrt{\varDelta x^{\mathsf{3} } \,g_{\mathsf{3} \mathsf{3} } \,\varDelta x^{\mathsf{3}}} = 2\beta _{+}\gamma _{+}\mathbf{R} \simeq 2\beta _{+}\mathbf{R}\) as a function of the rest-frame time. Setting this proper separation to be of order the graviton size Σ given by (14.165) then gives

$$\displaystyle{ \beta _{+} \sim \frac{\varSigma _{\mathrm{graviton}}} {\mathbf{R}} \sim \frac{\sqrt{\alpha '}} {\mathbf{R}} \sim \lambda ^{-1/4}\,, }$$
(14.168)

and (14.167) gives

$$\displaystyle{ x^{I} \sim {\bigl ( 1,0,0,\pm \lambda ^{-1/4},1\bigr )}\,z. }$$
(14.169)

Now boost back to the original plasma frame using (14.166) to get the early-time trajectories depicted by solid lines in Fig. 14.15b: \(x^{I} \sim {\Bigl (\gamma (1 \pm \lambda ^{-1/4}),0,0,\gamma (1 \pm \lambda ^{-1/4}),1\Bigr )}\,z\), where we have used γ ≫ 1 (14.166). Then

$$\displaystyle{ \frac{\varDelta (\mathit{dx}^{\mathsf{3}}/\mathit{dz})} {\mathit{dx}^{\mathsf{3}}/\mathit{dz}} \Bigg\vert _{\mathrm{initial}} \sim \lambda ^{-1/4}. }$$
(14.170)

As discussed before, the stopping distance (which requires a calculation in the full AdS5-Schwarzschild metric) covered by a null geodesic is power-law related to this initial slope, and so the difference Δ ℓ stop in how far the two bits of string travel also has the same small size (14.170) relative to stop:

$$\displaystyle{ \frac{\varDelta \ell_{\mathrm{stop}}} {\ell_{\mathrm{stop}}} \sim \lambda ^{-1/4}. }$$
(14.171)

Appendix 6: Checking the Penrose Limit: Details

In order to check the validity of the Penrose limit, here we characterize the string by following null geodesics that roughly trace different bits of string and which deviate slightly from our reference geodesic. This approximation amounts to ignoring the tension in the string as in Appendix 5 (for an alternative check of the Penrose limit outside of this approximation see [37]).

From the null geodesic formula and the metric (14.6), the x 3 coordinate for such geodesics is given by

$$\displaystyle{ \frac{\mathit{dx}^{\mathsf{3}}} {\mathit{dz}} = \frac{\hat{q}_{\mathsf{3}}} {\sqrt{1 - f\hat{\boldsymbol{q}}^{2}}}, }$$
(14.172)

where

$$\displaystyle{ \hat{q}_{\mu } \equiv \frac{q_{\mu }} {\omega } = (-1,\hat{\boldsymbol{q}}). }$$
(14.173)

Remembering that \(\varDelta x^{\mu } \equiv x^{\mu } -\bar{ x}^{\mu }(z)\) is the deviation relative to the reference geodesic, we have

$$\displaystyle{ \frac{d\varDelta x^{\mathsf{3}}} {\mathit{dz}} = \frac{\hat{q}_{\mathsf{3}}} {\sqrt{1 - f\hat{\boldsymbol{q}}^{2}}} - \frac{\bar{\hat{q}}_{\mathsf{3}}} {\sqrt{1 - f\bar{\hat{\boldsymbol{q}}}^{2}}}. }$$
(14.174)

Expand to first order in \(\varDelta \hat{q}_{3} \equiv \hat{ q}_{3} -\bar{\hat{ q}}_{3}\):

$$\displaystyle{ \frac{d\varDelta x^{\mathsf{3}}} {\mathit{dz}} \simeq \frac{\varDelta \hat{q}_{\mathsf{3}}} {(1 - f\bar{\hat{\boldsymbol{q}}}^{2})^{3/2}}. }$$
(14.175)

Then using (14.85) (and defining u with respect to the reference geodesic \(\bar{x}\)),

$$\displaystyle{ \frac{f\mathbf{R}^{2}} {z^{2}} \,\frac{d\varDelta x^{\mathsf{3}}} {\mathit{du}} \simeq \frac{f\,\varDelta \hat{q}_{\mathsf{3}}} {1 - f\hat{\boldsymbol{q}}^{2}}. }$$
(14.176)

Since \(1 - f\hat{\boldsymbol{q}}^{2} \simeq (z_{\star }^{4} + z^{4})/z_{\mathrm{h}}^{4}\), the combination (14.176) is largest for \(z \lesssim z_{\star }\), and the d Δ x 3du condition in (14.128) requires

$$\displaystyle{ \varDelta \hat{q}_{\mathsf{3}} \ll \frac{z_{\star }^{4}} {z_{\mathrm{h}}^{4}} }$$
(14.177)

for the Penrose limit. Use (14.18) to relate this to the stopping distance:

$$\displaystyle{ \ell_{\mathrm{stop}} \sim \frac{z_{\mathrm{h}}^{2}} {z_{\star }} \sim z_{\mathrm{h}}\left ( \frac{E^{2}} {-q^{2}}\right )^{1/4} \sim \frac{z_{\mathrm{h}}} {(1 -\hat{ q}_{\mathsf{3}})^{1/4}}, }$$
(14.178)

so that

$$\displaystyle{ \varDelta \ell_{\mathrm{stop}} \sim \frac{z_{\mathrm{h}}\,\varDelta \hat{q}_{\mathsf{3}}} {(1 -\hat{ q}_{\mathsf{3}})^{5/4}} \sim \frac{\varDelta \hat{q}_{\mathsf{3}}\,\ell_{\mathrm{stop}}} {1 -\hat{ q}_{\mathsf{3}}} \sim \varDelta \hat{ q}_{\mathsf{3}}\,\ell_{\mathrm{stop}}\,\frac{z_{\mathrm{h}}^{4}} {z_{\star }^{4}}. }$$
(14.179)

Combining (14.177) and (14.179) gives the condition

$$\displaystyle{ \varDelta \ell_{\mathrm{stop}} \ll \ell_{\mathrm{stop}} }$$
(14.180)

quoted in (14.130).

Now turn to the condition on dvdu in (14.128). The definition (14.77) of v gives

$$\displaystyle{ \mathit{dv} =\bar{\hat{ q}}_{\mu }\,d(\varDelta x^{\mu }) = -d(\varDelta x^{\mathsf{0}}) +\bar{\hat{ q}}_{\mathsf{ 3}}\,d(\varDelta x^{\mathsf{3}}), }$$
(14.181)

and so we need a formula for d(Δ x 0). The analog of (14.172) is

$$\displaystyle{ \frac{\mathit{dx}^{\mathsf{0}}} {\mathit{dz}} = \frac{-f^{-1}\hat{q}_{\mathsf{0}}} {\sqrt{1 - f\hat{\boldsymbol{q}}^{2}}}, }$$
(14.182)

with expansion

$$\displaystyle{ \frac{d\varDelta x^{\mathsf{0}}} {\mathit{dz}} \simeq \frac{\hat{q}_{\mathsf{3}}\,\varDelta \hat{q}_{\mathsf{3}}} {(1 - f\bar{\hat{q}}^{2})^{3/2}}. }$$
(14.183)

Combining (14.175), (14.181), and (14.183), gives dvdu ≃ 0. We therefore have to go back and make our expansions to second-order in \(\varDelta \hat{\boldsymbol{q}}\). The result is

$$\displaystyle{ \frac{\mathit{dv}} {\mathit{dz}} \simeq - \frac{(\varDelta \hat{\boldsymbol{q}}_{\perp })^{2}} {2(1 - f\bar{\hat{q}}_{3}^{2})^{1/2}} - \frac{(\varDelta \hat{q}_{\mathsf{3}})^{2}} {2(1 - f\bar{\hat{q}}_{3}^{2})^{3/2}}\,, }$$
(14.184)

and so

$$\displaystyle{ \frac{f\mathbf{R}^{2}} {z^{2}} \left \vert \frac{\mathit{dv}} {\mathit{du}}\right \vert \simeq \frac{f(\varDelta \hat{\boldsymbol{q}}_{\perp })^{2}} {2} + \frac{f(\varDelta \hat{q}_{\mathsf{3}})^{2}} {2(1 - f\bar{\hat{q}}_{3}^{2})}\,. }$$
(14.185)

The corresponding condition on dvdu in (14.128) is then

$$\displaystyle{ f(\varDelta \hat{\boldsymbol{q}}_{\perp })^{2}\quad \mbox{ and}\quad \frac{f(\varDelta \hat{q}_{\mathsf{3}})^{2}} {(1 - f\bar{\hat{q}}_{3}^{2})}\quad \ll \quad 1. }$$
(14.186)

The first condition is strongest for z ≪ z h and the second for \(z \lesssim z_{\star }\), giving

$$\displaystyle{ \vert \varDelta \hat{\boldsymbol{q}}_{\perp }\vert \quad \mbox{ and}\quad \vert \varDelta \hat{q}_{\mathsf{3}}\vert \,\frac{z_{\mathrm{h}}^{2}} {z_{\star }^{2}} \quad \ll \quad 1. }$$
(14.187)

Using (14.179), the condition involving \(\varDelta \hat{q}_{\mathsf{3}}\) becomes

$$\displaystyle{ \varDelta \ell_{\mathrm{stop}} \ll \ell_{\mathrm{stop}}\frac{z_{\mathrm{h}}^{2}} {z_{\star }^{2}} \,. }$$
(14.188)

Since z  ≪ z h, this is weaker than the previous condition (14.180).

Lastly, consider the other condition, \(\vert \varDelta \boldsymbol{q}_{\perp }\vert \ll 1\) in (14.187). To estimate \(\vert \varDelta \boldsymbol{q}_{\perp }\vert \), return to the arguments of Appendix 5, but now, in the rest frame, include an initial proper displacement of the two geodesics in \(\boldsymbol{x}^{\perp }\) of the same parametric size as the initial proper displacement in x 3. Following through the argument, one finds \(x^{I} \simeq {\Bigl (\gamma (1 +\beta \beta _{+}),\boldsymbol{\beta }_{\perp },\gamma (\beta +\beta _{+}),1\Bigr )}\,z\) with β  ⊥  ∼ β +. Then

$$\displaystyle{ \varDelta \hat{q}_{\mathsf{3}} =\varDelta \frac{q_{\mathsf{3}}} {q_{\mathsf{0}}} =\varDelta \frac{\mathit{dx}^{\mathsf{3}}/\mathit{dz}} {\mathit{dx}^{\mathsf{0}}/\mathit{dz}} \simeq \varDelta \frac{\gamma (\beta +\beta _{+})} {\gamma (1 +\beta \beta _{+})} \simeq \frac{\beta _{+}} {\gamma ^{2}} }$$
(14.189)

and

$$\displaystyle{ \varDelta \hat{\boldsymbol{q}}_{\perp } =\varDelta \frac{\boldsymbol{q}_{\perp }} {q_{\mathsf{0}}} =\varDelta \frac{d\boldsymbol{x}^{\perp }/\mathit{dz}} {\mathit{dx}^{\mathsf{0}}/\mathit{dz}} \simeq \varDelta \frac{\boldsymbol{\beta }_{\perp }} {\gamma (1 +\beta \beta _{+})} \simeq \frac{\boldsymbol{\beta }_{\perp }} {\gamma }, }$$
(14.190)

so that

$$\displaystyle{ \frac{\vert \varDelta \hat{\boldsymbol{q}}_{\perp }\vert } {\vert \varDelta \hat{q}_{\mathsf{3}}\vert } \sim \gamma \sim \sqrt{ \frac{E^{2 } } {-q^{2}}} \sim \frac{z_{\mathrm{h}}^{2}} {z_{\star }^{2}} \,. }$$
(14.191)

So, using (14.179),

$$\displaystyle{ \vert \varDelta \hat{\boldsymbol{q}}_{\perp }\vert \sim \vert \varDelta \hat{q}_{\mathsf{3}}\vert \,\frac{z^{2}} {z_{\star }^{2}} \sim \frac{\varDelta \ell_{\mathrm{stop}}} {\ell_{\mathrm{stop}}}\, \frac{z_{\star }^{2}} {z_{\mathrm{h}}^{2}}\,. }$$
(14.192)

The condition \(\vert \varDelta \hat{\boldsymbol{q}}_{\perp }\vert \ll 1\) is therefore the same as the previous condition (14.188) and so is also weaker than (14.180).

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Vaman, D. (2015). Beyond Supergravity in AdS-CFT: An Application to Jet Quenching. In: Papantonopoulos, E. (eds) Modifications of Einstein's Theory of Gravity at Large Distances. Lecture Notes in Physics, vol 892. Springer, Cham. https://doi.org/10.1007/978-3-319-10070-8_14

Download citation

Publish with us

Policies and ethics