Skip to main content
Log in

Model independent evolution of transverse momentum dependent distribution functions (TMDs) at NNLL

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

Abstract

We discuss the evolution of the eight leading-twist transverse momentum dependent parton distribution functions, which turns out to be universal and spin independent. By using the highest order perturbatively calculable ingredients at our disposal, we perform the resummation of the large logarithms that appear in the evolution kernel of transverse momentum distributions up to next-to-next-to-leading logarithms (NNLL), thus obtaining an expression for the kernel with highly reduced model dependence. Our results can also be obtained using the standard CSS approach when a particular choice of the b prescription is used. In this sense, and while restricted to the perturbative domain of applicability, we consider our results as a “prediction” of the correct value of b max which is very close to 1.5 GeV−1. We explore under which kinematical conditions the effects of the non-perturbative region are negligible, and hence the evolution of transverse momentum distributions can be applied in a model independent way. The application of the kernel is illustrated by considering the unpolarized transverse momentum dependent parton distribution function and the Sivers function.

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

Access this article

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

Instant access to the full article PDF.

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

Similar content being viewed by others

Notes

  1. All quantities with tilde in this paper refer to quantities calculated in impact parameter space.

  2. Since the evolution kernel is the same for \(\tilde{F}_{n}\) and \(\tilde {F}_{{\bar{n}}}\), we have dropped out the \(n,{\bar{n}}\) labels.

  3. In the notation of [29], our d n (0) corresponds to their \(d_{n}^{q}/2\).

References

  1. A. Accardi, J.L. Albacete, M. Anselmino, N. Armesto, E.C. Aschenauer et al., Electron Ion Collider: the next QCD frontier—understanding the glue that binds us all. arXiv:1212.1701 [nucl-ex] (2012)

  2. D.W. Sivers, Single spin production asymmetries from the hard scattering of point-like constituents. Phys. Rev. D 41, 8 (1990)

    Article  ADS  Google Scholar 

  3. J.C. Collins, Fragmentation of transversely polarized quarks probed in transverse momentum distributions. Nucl. Phys. B 396, 161–182 (1993). arXiv:hep-ph/9208213 [hep-ph]

    Article  ADS  Google Scholar 

  4. V. Barone, F. Bradamante, A. Martin, Transverse-spin and transverse-momentum effects in high-energy processes. Prog. Part. Nucl. Phys. 65, 267–333 (2010). arXiv:1011.0909 [hep-ph]

    Article  ADS  Google Scholar 

  5. S. Mert Aybat, T.C. Rogers, TMD parton distribution and fragmentation functions with QCD evolution. Phys. Rev. D 83, 114042 (2011). arXiv:1101.5057 [hep-ph]

    Article  ADS  Google Scholar 

  6. S. Mert Aybat, J.C. Collins, J.-W. Qiu, T.C. Rogers, The QCD evolution of the Sivers function. Phys. Rev. D 85, 034043 (2012). arXiv:1110.6428 [hep-ph]

    Article  ADS  Google Scholar 

  7. S. Mert Aybat, A. Prokudin, T.C. Rogers, Calculation of TMD evolution for transverse single spin asymmetry measurements. Phys. Rev. Lett. 108, 242003 (2012). arXiv:1112.4423 [hep-ph]

    Article  ADS  Google Scholar 

  8. Z.-B. Kang, B.-W. Xiao, F. Yuan, QCD resummation for single spin asymmetries. Phys. Rev. Lett. 107, 152002 (2011). arXiv:1106.0266 [hep-ph]

    Article  ADS  Google Scholar 

  9. M. Anselmino, M. Boglione, S. Melis, A strategy towards the extraction of the Sivers function with TMD evolution. Phys. Rev. D 86, 014028 (2012). arXiv:1204.1239 [hep-ph]

    Article  ADS  Google Scholar 

  10. P. Sun, B.-W. Xiao, F. Yuan, Gluon distribution functions and Higgs boson production at moderate transverse momentum. Phys. Rev. D 84, 094005 (2011). arXiv:1109.1354 [hep-ph]

    Article  ADS  Google Scholar 

  11. B.U. Musch, P. Hägler, M. Engelhardt, J.W. Negele, A. Schäfer, Sivers and Boer–Mulders observables from lattice QCD. Phys. Rev. D 85, 094510 (2012). arXiv:1111.4249 [hep-lat]

    Article  ADS  Google Scholar 

  12. M.G. Echevarria, A. Idilbi, I. Scimemi, Soft and collinear factorization and transverse momentum dependent parton distribution functions. arXiv:1211.1947 [hep-ph] (2012)

  13. M.G. Echevarria, A. Idilbi, I. Scimemi, Factorization theorem for Drell–Yan at low q T and transverse momentum distributions on-the-light-cone. J. High Energy Phys. 1207, 002 (2012). arXiv:1111.4996 [hep-ph]

    Article  ADS  Google Scholar 

  14. M.G. Echevarria, A. Idilbi, I. Scimemi, Definition and evolution of transverse momentum distributions. Int. J. Mod. Phys. Conf. Ser. 20, 92–108 (2012). arXiv:1209.3892 [hep-ph]

    Article  Google Scholar 

  15. J. Collins, Foundations of Perturbative QCD. Cambridge Monographs on Particle Physics, Nuclear Physics and Cosmology, vol. 32 (2011)

    Book  Google Scholar 

  16. J.C. Collins, T.C. Rogers, Equality of two definitions for transverse momentum dependent parton distribution functions. Phys. Rev. D 87, 034018 (2013). arXiv:1210.2100 [hep-ph]

    Article  ADS  Google Scholar 

  17. J.C. Collins, D.E. Soper, Parton distribution and decay functions. Nucl. Phys. B 194, 445 (1982)

    Article  ADS  Google Scholar 

  18. J.C. Collins, D.E. Soper, Back-to-back jets in QCD. Nucl. Phys. B 193, 381 (1981)

    Article  ADS  Google Scholar 

  19. J.C. Collins, D.E. Soper, G.F. Sterman, Transverse momentum distribution in Drell–Yan pair and W and Z boson production. Nucl. Phys. B 250, 199 (1985)

    Article  ADS  Google Scholar 

  20. A.V. Konychev, P.M. Nadolsky, Universality of the Collins–Soper–Sterman nonperturbative function in gauge boson production. Phys. Lett. B 633, 710–714 (2006). arXiv:hep-ph/0506225 [hep-ph]

    Article  ADS  Google Scholar 

  21. A.V. Manohar, I.W. Stewart, The zero-bin and mode factorization in quantum field theory. Phys. Rev. D 76, 074002 (2007). arXiv:hep-ph/0605001 [hep-ph]

    Article  ADS  Google Scholar 

  22. P.J. Mulders, R.D. Tangerman, The complete tree level result up to order 1/Q for polarized deep inelastic leptoproduction. Nucl. Phys. B 461, 197–237 (1996). arXiv:hep-ph/9510301 [hep-ph]

    Article  ADS  Google Scholar 

  23. D. Boer, P.J. Mulders, Time reversal odd distribution functions in leptoproduction. Phys. Rev. D 7, 5780–5786 (1998). arXiv:hep-ph/9711485 [hep-ph]

    Article  ADS  Google Scholar 

  24. A. Idilbi, I. Scimemi, Singular and regular gauges in soft collinear effective theory: the introduction of the new Wilson line T. Phys. Lett. B 695, 463–468 (2011). arXiv:1009.2776 [hep-ph]

    Article  ADS  Google Scholar 

  25. M. Garcia-Echevarria, A. Idilbi, I. Scimemi, SCET, light-cone gauge and the T-Wilson lines. Phys. Rev. D 84, 011502 (2011). arXiv:1104.0686 [hep-ph]

    Article  ADS  Google Scholar 

  26. A. Idilbi, X.-d. Ji, F. Yuan, Resummation of threshold logarithms in effective field theory for DIS, Drell–Yan and Higgs production. Nucl. Phys. B 753, 42–68 (2006). arXiv:hep-ph/0605068 [hep-ph]

    Article  ADS  Google Scholar 

  27. S. Moch, J.A.M. Vermaseren, A. Vogt, The quark form-factor at higher orders. J. High Energy Phys. 0508, 049 (2005). arXiv:hep-ph/0507039 [hep-ph]

    Article  ADS  Google Scholar 

  28. S. Moch, J.A.M. Vermaseren, A. Vogt, The three loop splitting functions in QCD: the nonsinglet case. Nucl. Phys. B 688, 101–134 (2004). arXiv:hep-ph/0403192 [hep-ph]

    Article  ADS  MATH  MathSciNet  Google Scholar 

  29. T. Becher, M. Neubert, Drell–Yan production at small q T , transverse parton distributions and the collinear anomaly. Eur. Phys. J. C 71, 1665 (2011). arXiv:1007.4005 [hep-ph]

    Article  ADS  Google Scholar 

  30. A. Grozin, Decoupling in QED and QCD. Int. J. Mod. Phys. A 28, 1350015 (2013). arXiv:1212.5144 [hep-ph]

    Article  ADS  MathSciNet  Google Scholar 

  31. S.A. Larin, T. van Ritbergen, J.A.M. Vermaseren, The large quark mass expansion of Γ(Z 0→hadrons) and Γ(τ ν τ +hadrons) in the order \(\alpha_{s}^{3}\). Nucl. Phys. B 438, 278–306 (1995). arXiv:hep-ph/9411260 [hep-ph]

    Article  ADS  Google Scholar 

  32. K.G. Chetyrkin, B.A. Kniehl, M. Steinhauser, Decoupling relations to \(\mathcal{O}(\alpha_{s}^{3})\) and their connection to low-energy theorems. Nucl. Phys. B 510, 61–87 (1998). arXiv:hep-ph/9708255 [hep-ph]

    ADS  Google Scholar 

  33. K.G. Chetyrkin, J.H. Kühn, M. Steinhauser, RunDec: a Mathematica package for running and decoupling of the strong coupling and quark masses. Comput. Phys. Commun. 133, 43–65 (2000). arXiv:hep-ph/0004189 [hep-ph]

    Article  ADS  MATH  Google Scholar 

  34. K.G. Chetyrkin, J.H. Kühn, C. Sturm, QCD decoupling at four loops. Nucl. Phys. B 744, 121–135 (2006). arXiv:hep-ph/0512060 [hep-ph]

    Article  ADS  Google Scholar 

  35. Y. Schroder, M. Steinhauser, Four-loop decoupling relations for the strong coupling. J. High Energy Phys. 0601, 051 (2006). arXiv:hep-ph/0512058 [hep-ph]

    Article  ADS  MathSciNet  Google Scholar 

  36. M. Anselmino, M. Boglione, U. D’Alesio, A. Kotzinian, F. Murgia et al., The role of Cahn and Sivers effects in deep inelastic scattering. Phys. Rev. D 71, 074006 (2005). arXiv:hep-ph/0501196 [hep-ph]

    Article  ADS  Google Scholar 

  37. P. Schweitzer, T. Teckentrup, A. Metz, Intrinsic transverse parton momenta in deeply inelastic reactions. Phys. Rev. D 81, 094019 (2010). arXiv:1003.2190 [hep-ph]

    Article  ADS  Google Scholar 

  38. J.C. Collins, A.V. Efremov, K. Goeke, S. Menzel, A. Metz et al., Sivers effect in semi-inclusive deeply inelastic scattering. Phys. Rev. D 73, 014021 (2006). arXiv:hep-ph/0509076 [hep-ph]

    Article  ADS  Google Scholar 

  39. M. Anselmino, M. Boglione, U. D’Alesio, A. Kotzinian, S. Melis et al., Sivers effect for pion and kaon production in semi-inclusive deep inelastic scattering. Eur. Phys. J. A 39, 89–100 (2009). arXiv:0805.2677 [hep-ph]

    Article  ADS  Google Scholar 

  40. A.D. Martin, W.J. Stirling, R.S. Thorne, G. Watt, Parton distributions for the LHC. Eur. Phys. J. C 63, 189–285 (2009). arXiv:0901.0002 [hep-ph]

    Article  ADS  Google Scholar 

  41. P.J. Mulders, J. Rodrigues, Transverse momentum dependence in gluon distribution and fragmentation functions. Phys. Rev. D 63, 094021 (2001). arXiv:hep-ph/0009343 [hep-ph]

    Article  ADS  Google Scholar 

  42. F. Landry, R. Brock, P.M. Nadolsky, C.P. Yuan, Tevatron run-1 Z boson data and Collins–Soper–Sterman resummation formalism. Phys. Rev. D 67, 073016 (2003). arXiv:hep-ph/0212159 [hep-ph]

    Article  ADS  Google Scholar 

Download references

Acknowledgements

This work is supported by the Spanish MEC, FPA2011-27853-CO2-02. We would like to thank Umberto D’Alesio, V. Braun, M. Diehl, A. Metz and J. Zhou for useful discussions. M.G.E. is supported by the Ph.D. funding program of the Basque Country Government. A.I. and A.S. are supported by BMBF (06RY9191). I.S. was partly supported by the Ramón y Cajal Program.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Miguel G. Echevarría.

Appendices

Appendix A: CSS approach to the evolution of TMDs

In various works following Collins’ approach to TMDs [6, 9, 15], large L logarithms in the D term of the evolution kernel in Eq. (7) were resummed using the CSS approach [19], which, for the sake of completeness and comparison, we explain below.

First, the D term is resummed using its RG-evolution in Eq. (8),

$$\begin{aligned} D (b;Q_i ) &= D (b;\mu_b ) + \int _{\mu_b}^{Q_i}\frac{d\bar{\mu}}{\bar{\mu}} \varGamma_{\mathrm{cusp}} , \end{aligned}$$
(A.1)

where large L logarithms in the D term on the right hand side are canceled by choosing \(\mu_{b} = 2e^{-\gamma_{E}}/b\). Thus, they are resummed by evolving D from μ b to Q i . However, since we need to Fourier transform back to momentum space, at some value of b the effective coupling α s (μ b ) will hit the Landau pole. In Fig. 7(a) we can see the evolution kernel \(\tilde {R}(b;Q_{i}=\sqrt{2.4}~\mathrm{GeV},Q_{f}=5~\mathrm{GeV})\) where we have used Eq. (A.1) to resum the L logarithms in D and the appearance of the Landau pole is manifest. The breakdown of the perturbative series is driven by the running coupling α s (μ b ), when μ b is sufficiently small.

Fig. 7
figure 7

Evolution kernel from \(Q_{i}=\sqrt{2.4}~\mathrm{GeV}\) up to Q f =5 GeV using RG-evolution in Eq. (A.1) to resum the D term. The resummation accuracy is given in Table 2. (a) With \(\mu_{b} = 2e^{-\gamma_{E}}/b\) the Landau pole appears clearly. (b) With \(\mu_{b^{*}} = 2e^{-\gamma_{E}}/{b^{*}}\) and b max=1.5 GeV−1 to avoid hitting the Landau pole. (c) Adding the BLNY non-perturbative model to recover the information at large b

In order to avoid this issue, CSS did not actually introduce a sharp cut-off but a smoothed one defined as \(b^{*}=b/\sqrt{1+(b/b_{\max})^{2}}\). Obviously b cannot exceed b max and the effective coupling \(\alpha_{s}(\mu_{b^{*}})\) does not hit the Landau pole. As is shown in Fig. 7(b), the kernel saturates and does not present any uncontrolled behavior. It is also worth noticing that comparing Fig. 7(a) with Fig. 7(b), we see that the implementation of the b prescription has some appreciable effect in the perturbative region, which now depends on this parameter.

The lost information due to the cut-off is recovered by adding a non-perturbative model that has to be extracted from experimental data of a measured cross section. This model not only gives the proper information in the non-perturbative region, but also restores the correct shape of the kernel within the perturbative domain, which was affected by the b prescription. When implementing, for example, the Brock–Landry–Nadolsky–Yuan (BLNY )model the evolution kernel can be written as

$$\begin{aligned} & \tilde{R}^{\mathrm{CSS}}(b;Q_i,Q_f) \\ &\quad = \exp \biggl\{ \int_{Q_i}^{Q_f} \frac{d\bar{\mu}}{\bar{\mu}} \gamma_F \biggl(\alpha _s(\bar{\mu}),\ln \frac{Q_f^2}{\bar{\mu}^2} \biggr) \biggr\} \\ &\qquad{}\times \biggl(\frac{Q_f^2}{Q_i^2} \biggr)^ {- [D (b^*;Q_i )+\frac{1}{4}g_2 b^2 ]} , \end{aligned}$$
(A.2)

where D(b ;Q i ) is resummed using Eq. (A.1). In this model g 2=0.68 GeV2 for b max=0.5 GeV−1 [42] and g 2=0.184 GeV2 for b max=1.5 GeV−1 [20]. From the theoretical point of view these two choices are legitimate and they can be used to define the model dependence of the final result. However, considering b max as a fitting parameter the choice of b max=1.5 GeV−1 should be preferred according to Ref. [20]. Figure 7(c) shows the complete evolution kernel with the CSS approach, Eq. (A.2), while implementing the BLNY model.

Appendix B: Derivation of D R up to NNNLL

Below we provide the details of the derivation of D R within the CSS formalism, i.e., solving Eq. (24). Using the perturbative expansion of Γ cusp(α s ) and β(α s ) one can write,

$$\begin{aligned} &\int_{\mu_b}^{Q_i} d(\ln \mu) \varGamma_{\mathrm{cusp}} \\ &\quad = \int_{\alpha_s(\mu_b)}^{\alpha_s(Q_i)} d\alpha\; \frac{\varGamma _{\mathrm{cusp}}(\alpha)}{\beta(\alpha)} \\ &\quad =\int_{\alpha_s(\mu_b)}^{\alpha_s(Q_i)} d\alpha\; \biggl\{ \frac{-\varGamma_0}{2 \alpha\beta_0}+ \frac{\varGamma_0\beta_1-\varGamma_1 \beta_0}{8 \pi\beta_0^2}+ \frac{\alpha (-\beta_0^2 \varGamma_2+\beta_0 \beta_1 \varGamma _1+\beta_0 \beta_2 \varGamma_0-\beta_1^2 \varGamma_0 )}{32 \pi^2 \beta_0^3} \\ &\qquad {} + \frac{\alpha^2 ( -\beta_0^3 \varGamma_3+\beta_0^2 \beta_1 \varGamma_2+\beta_0^2 \beta_2\varGamma_1+\beta_0^2 \beta _3\varGamma_0-\beta_0 \beta_1^2 \varGamma_1-2 \beta_0 \beta_1\beta_2 \varGamma_0+\beta_1^3 \varGamma_0 )}{128 \pi^3\beta_0^4} \biggr\} \\ &\quad = \frac{-\varGamma_0}{2 \beta_0}\ln \frac{\alpha_s(Q_i)}{\alpha _s(\mu_b)}+ \bigl[ \alpha_s(Q_i)-\alpha_s(\mu_b) \bigr]\frac{\varGamma_0\beta _1-\varGamma _1 \beta_0}{8 \pi\beta_0^2}+ \bigl[\alpha_s^2(Q_i)- \alpha_s^2(\mu_b) \bigr]\frac{ (-\beta _0^2 \varGamma_2+\beta_0 \beta_1 \varGamma_1+\beta_0 \beta_2 \varGamma _0-\beta_1^2 \varGamma_0 )}{64 \pi^2 \beta_0^3} \\ &\qquad {}+ \bigl[\alpha_s^3(Q_i)- \alpha_s^3(\mu_b) \bigr]\frac{ ( -\beta _0^3 \varGamma_3+\beta_0^2 \beta_1 \varGamma_2+\beta_0^2 \beta_2\varGamma_1+\beta_0^2 \beta _3\varGamma_0-\beta_0 \beta_1^2 \varGamma_1-2 \beta_0 \beta_1\beta_2 \varGamma_0+\beta_1^3 \varGamma_0 )}{384 \pi^3\beta_0^4} . \end{aligned}$$
(B.1)

Then in the equation above we use the running of the strong coupling to expand α s (μ b ) in terms of α s (Q i ),

$$\begin{aligned} \alpha_s(\mu_b)&= \alpha_s(Q_i) \frac{1}{1-X} - \alpha_s^2(Q_i) \frac{\beta_1 \ln (1-X)}{4\pi(1-X)^2 \beta_0}- \alpha_s^3(Q_i) \frac{ (-X \beta_0 \beta_2+ \beta_1^2 (X-\ln ^2(1-X)+ \ln (1-X)) )}{16 \pi^2 (1-X)^3 \beta_0^2} \\ &\quad {} -\alpha_s^4(Q_i) \frac{ (\beta_1^3 (X^2+2 \ln ^3(1-X)-5 \ln ^2(1-X)-4 X \ln (1-X) ) )}{128 \pi^3 (X-1)^4 \beta_0^3} \\ &\quad {}-\alpha_s^4(Q_i)\frac{ (+2\beta_0\beta_1\beta_2 ((2 X+1) \ln (1-X)-(X-1) X)+(X-2) X \beta_0^2 \beta_3 )}{128 \pi^3 (X-1)^4\beta_0^3} \end{aligned}$$
(B.2)

and implement it up to the appropriate order in α s (Q i ). In order to finally get D R at NNLL one should consider also the term D(b;μ b ) in Eq. (24), which at second order does not vanish due to the presence of the finite d 2(0) term,

$$\begin{aligned} D^{(2)}(b;\mu_b)= d_2(0) \biggl( \frac{\alpha_s(\mu_b)}{4\pi} \biggr)^2=d_2(0) \frac {a^2}{(1-X)^2} , \end{aligned}$$
(B.3)

with a=α s (Q i )/4π. Inserting this result in Eq. (B.1) and the expansion in Eq. (B.2) up to order \(\alpha_{s}^{2}(Q_{i})\), one gets our Eq. (16).

Finally, for completeness and future reference, we provide also D R at NNNLL,

$$\begin{aligned} D^{R(3)}&= \frac{a^3}{(1-X)^3} \biggl(d_3(0)-2d_2(0) \frac{\beta _1}{\beta_0}\ln (1-X)+D^{R(3)}_\varGamma \biggr) , \\ D^{R(3)}_\varGamma&=-\frac{1}{12\beta_0^4} \bigl[\beta_0^2 \bigl(2 \varGamma_2 \beta_1 \bigl(X \bigl(X^2-3 X+3 \bigr) \\ &\quad {}+3 \ln (1-X) \bigr)+X^2 \bigl(2 \varGamma_1 (X-3) \beta_2 \\ &\quad {}+\varGamma_0 (2 X-3)\beta_3 \bigr) \bigr) \\ &\quad {} -2 \beta_0\beta_1 \bigl(\varGamma_1 \beta_1 \bigl((X-3) X^2+3 \ln ^2(1-X) \bigr) \\ &\quad {}+\varGamma _0 X \beta_2 \bigl(X (2 X-3)-3 \ln (1-X)\bigr) \bigr) \\ &\quad {}-2 \varGamma_3 X \bigl(X^2-3 X+3 \bigr) \beta_0^3 \\ &\quad {} +\varGamma_0 \beta_1^3 \bigl(X^2 (2 X-3)+2 \ln ^3(1-X) \\ &\quad -3 \ln ^2(1-X)-6 X \ln (1-X) \bigr) \bigr] . \end{aligned}$$
(B.4)

Appendix C: Evolution of the hard matching coefficient

The evolution of the hard matching coefficient C V , where H=|C V |2, is given by

$$\begin{aligned} \begin{array}{l} \displaystyle\frac{d}{d\ln \mu}\ln C_V\bigl(Q^2/ \mu^2\bigr) = \gamma_{C_V} \biggl(\alpha_s( \mu),\ln \displaystyle\frac{Q^2}{\mu^2} \biggr) , \\ \gamma_{C_V} = \varGamma_{\mathrm{cusp}}(\alpha_s)\ln \displaystyle\frac{Q^2}{\mu^2} + \gamma^V(\alpha_s) , \end{array} \end{aligned}$$
(C.1)

where the cusp term is related to the evolution of the Sudakov double logarithms and the remaining term with the evolution of single logarithms. The exact solution of this equation is

$$\begin{aligned} & C_V\bigl(Q^2/\mu_f^2 \bigr) \\ &\quad = C_V\bigl(Q^2/\mu_i^2 \bigr) \exp \biggl[ \int_{\mu_i}^{\mu_f} \frac{d\bar{\mu}}{\bar{\mu}} \gamma_{C_V} \biggl(\alpha_s(\bar{\mu}), \ln \frac{Q^2}{\bar{\mu}^2} \biggr) \biggr] \\ &\quad = C_V\bigl(Q^2/\mu_i^2\bigr) \exp \biggl[ \int_{\alpha_s(\mu_i)}^{\alpha_s(\mu_f)} \frac {d\bar{\alpha}_s}{\beta(\bar{\alpha}_s)} \gamma_{C_V} (\bar{\alpha}_s ) \biggr] , \end{aligned}$$
(C.2)

where we have used d/dlnμ=β(α s )d/ s , where β(α s )= s /dlnμ is the QCD β-function.

Below we give the expressions for the anomalous dimensions and the QCD β-function, in the \(\overline{{\rm MS}}\) renormalization scheme. We use the following expansions:

$$\begin{aligned} &\varGamma_{\mathrm{cusp}} = \sum_{n=1}^{\infty} \varGamma_{n-1} \biggl( \frac{\alpha_s}{4\pi } \biggr)^n , \\ &\gamma^V_{n-1} = \sum_{n=1}^{\infty} \gamma^V_{n} \biggl( \frac{\alpha _s}{4\pi} \biggr)^n , \\ &\beta= -2\alpha_s\sum _{n=1}^\infty\beta_{n-1} \biggl( \frac {\alpha_s}{4\pi} \biggr)^n . \end{aligned}$$
(C.3)

The coefficients for the cusp anomalous dimension Γ cusp are

$$\begin{aligned} &\varGamma_0 = 4 C_F , \\ &\varGamma_1 = 4 C_F \biggl[ \biggl( \frac{67}{9} - \frac{\pi^2}{3} \biggr) C_A - \frac{20}{9}T_F n_f \biggr] , \\ &\varGamma_2 = 4 C_F \biggl[ C_A^2 \biggl( \frac{245}{6} - \frac{134\pi^2}{27} + \frac{11\pi^4}{45} + \frac{22}{3}\zeta_3 \biggr)\\ &\hphantom{\varGamma_2 =} + C_A T_F n_f \biggl( - \frac{418}{27} + \frac{40\pi^2}{27} - \frac{56}{3}\zeta_3 \biggr) \\ &\hphantom{\varGamma_2 =} {}+ C_F T_F n_f \biggl( - \frac{55}{3} + 16\zeta_3 \biggr) - \frac{16}{27} T_F^2 n_f^2 \biggr] . \end{aligned}$$
(C.4)

The anomalous dimension γ V can be determined up to 3-loop order from the partial 3-loop expression for the on-shell quark form factor in QCD. We have

$$\begin{aligned} &\gamma_0^V = -6 C_F , \\ &\gamma_1^V = C_F^2 \bigl( -3 + 4\pi^2 - 48\zeta_3 \bigr) \\ &\hphantom{\gamma_1^V =} {}+C_F C_A \biggl( - \frac{961}{27} - \frac{11\pi^2}{3} + 52\zeta_3 \biggr) \\ &\hphantom{\gamma_1^V =} {}+ C_F T_F n_f \biggl( \frac{260}{27} + \frac{4\pi^2}{3} \biggr) , \\ & \gamma_2^V = C_F^3 \biggl( -29 - 6\pi^2 - \frac{16\pi^4}{5} - 136\zeta_3 \\ &\hphantom{\gamma_2^V =} {}+ \frac{32\pi^2}{3}\zeta_3+ 480\zeta_5 \biggr) \\ &\hphantom{\gamma_2^V =} {}+ C_F^2 C_A \biggl( - \frac{151}{2} + \frac{410\pi^2}{9} + \frac{494\pi^4}{135} - \frac{1688}{3}\zeta_3\\ &\hphantom{\gamma_2^V =} {}- \frac{16\pi^2}{3}\zeta_3 - 240\zeta_5 \biggr) \\ &\hphantom{\gamma_2^V =} {}+ C_F C_A^2 \biggl( - \frac{139345}{1458} - \frac{7163\pi^2}{243} - \frac{83\pi^4}{45} + \frac{7052}{9}\zeta_3 \\ &\hphantom{\gamma_2^V =} {}- \frac{88\pi^2}{9}\zeta_3 - 272\zeta_5 \biggr) \\ &\hphantom{\gamma_2^V =} {}+ C_F^2 T_F n_f \biggl( \frac{5906}{27} - \frac{52\pi^2}{9} - \frac{56\pi^4}{27} + \frac{1024}{9}\zeta_3 \biggr) \\ &\hphantom{\gamma_2^V =} {}+ C_F C_A T_F n_f \biggl( - \frac{34636}{729} + \frac{5188\pi^2}{243} + \frac{44\pi^4}{45} \\ &\hphantom{\gamma_2^V =} {} - \frac{3856}{27} \zeta_3 \biggr) + C_F T_F^2 n_f^2 \biggl( \frac{19336}{729} - \frac{80\pi^2}{27} - \frac{64}{27} \zeta_3 \biggr) . \end{aligned}$$
(C.5)

Finally, the coefficients for the QCD β-function are

$$\begin{aligned} &\beta_0 = \frac{11}{3}C_A - \frac{4}{3}T_F n_f , \\ &\beta_1 = \frac{34}{3}C_A^2 - \frac{20}{3}C_A T_F n_f - 4 C_F T_F n_f , \\ &\beta_2 = \frac{2857}{54}C_A^3 + \biggl( 2 C_F^2 - \frac{205}{9}C_F C_A - \frac{1415}{27}C_A^2 \biggr) T_F n_f\\ &\hphantom{\beta_2 =} {}+ \biggl( \frac{44}{9}C_F + \frac{158}{27}C_A \biggr) T_F^2 n_f^2 , \\ &\beta_3 = \frac{149753}{6} + 3564\zeta_3 - \biggl( \frac{1078361}{162} + \frac{6508}{27}\zeta_3 \biggr) n_f \\ &\hphantom{\beta_3 =} {}+ \biggl( \frac{50065}{162} + \frac{6472}{81}\zeta_3 \biggr) n_f^2 + \frac{1093}{729}n_f^3 , \end{aligned}$$
(C.6)

where for β 3 we have used N c =3 and \(T_{F}=\frac{1}{2}\).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Echevarría, M.G., Idilbi, A., Schäfer, A. et al. Model independent evolution of transverse momentum dependent distribution functions (TMDs) at NNLL. Eur. Phys. J. C 73, 2636 (2013). https://doi.org/10.1140/epjc/s10052-013-2636-y

Download citation

  • Received:

  • Revised:

  • Published:

  • DOI: https://doi.org/10.1140/epjc/s10052-013-2636-y

Keywords

Navigation