Skip to main content

Polylogarithms and Multizeta Values in Massless Feynman Amplitudes

  • Conference paper
  • First Online:
Lie Theory and Its Applications in Physics

Part of the book series: Springer Proceedings in Mathematics & Statistics ((PROMS,volume 111))

Abstract

The last two decades have seen a remarkable development of analytic methods in the study of Feynman amplitudes in perturbative quantum field theory. The present lecture offers a physicists’ oriented survey of Francis Brown’s work on singlevalued multiple polylogarithms, the associated multizeta periods and their application to Schnetz’s graphical functions and to x-space renormalization. To keep the discussion concrete we restrict attention to explicit examples of primitively divergent graphs in a massless scalar QFT.

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 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover 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.

    Iterated integrals were introduced in the mid 1950s and developed essentially single-handedly for over 20 years by Chen (1923–1987) [12] before gaining recognition in both mathematics and QFT—see [5].

References

  1. Bloch, S.L.: Applications of the dilogarithm function in algebraic K-theory and algebraic geometry. In: Proceedings of the Internat. Symp. on Alg. Geometry, Kinokuniya, Tokyo (1978)

    Google Scholar 

  2. Broadhurst, D.J.: Summation of an infinite series of ladder diagrams. Phys. Lett. B307, 132–139 (1993)

    Article  MathSciNet  Google Scholar 

  3. Broadhurst, D.J., Kreimer, D.: Associatioon of multiple zeta values with positive knots via Feynman diagrams up to 9 loops. Phys. Lett. B393, 403–412 (1997). arXiv:hep-th/9609128

    Google Scholar 

  4. Brown, F.: Polylogarithmes multiples uniformes en une variable. C. R. Acad. Sci. Paris Ser. I 338, 522–532 (2004)

    Article  Google Scholar 

  5. Brown, F.: Iterated Integrals in Quantum Field Theory. Villa de Leyva, Columbia Notes (2009)

    Google Scholar 

  6. Brown, F.: Mixed Tate motives over Z. Ann. Math. 175(1), 949–976 (2012). arXiv:1102.1312 [math.AG]; On the decomposition of motivic multiple zeta values. arXiv:1102.1310 [math.NT]

    Google Scholar 

  7. Brown, F.: Depth-graded motivic multiple zeta values (2013). arXiv:1301.3053 [math.NT]

    Google Scholar 

  8. Brown, F.: Single-valued periods and multiple zeta values (2013). arXiv:1309.5309 [math.NT]

    Google Scholar 

  9. Brown, F., Schnetz, O.: Proof of the zig-zag conjetcure (2012). arXiv:1208.1890 [math.NT]

    Google Scholar 

  10. Cartier, P.: Fonctions polylogarithmes, nombres polyzêtas et groupes pro-unipotents, Séminaire Bourbaki 53ème année n. 885, Astérisque 282, 137–173 (2002)

    MathSciNet  Google Scholar 

  11. Cartier, P.: On the double zeta values. Adv. Stud. Pure Math. 63, 91–119 (2012)

    MathSciNet  Google Scholar 

  12. Chen, K.T.: Iterated path integrals. Bull. Am. Math. Soc. 83, 831–879 (1977)

    Article  MATH  Google Scholar 

  13. Chetyrkin, K.G., Kataev, A.I., Tkachov, F.V.: New approach to evaluation of multiloop Feynman integrals: the Gegenbauer polynomial x-space technique. Nucl. Phys. B174, 345–377 (1980)

    Article  MathSciNet  Google Scholar 

  14. Deligne, P.: Multizêtas, d’après Francis Brown, Séminaire Bourbaki 64ème année n.1048 Astérisque 352 (2013)

    Google Scholar 

  15. Del Duca, V., Dixon, L.J., Duhr, C., Pennington, J.: The BFKL equation, Mueller-Navelet jets and single-valued harmonic polylogarithms. arXiv:1309.6647 [hep-ph]

    Google Scholar 

  16. Drummond, J.M., Henn, J.G., Korchemsky, P., Sokatchev, E.: Dual superconformal symmetry of scattering amplitudes in N = 4 super-Yang–Mills theory. Nucl. Phys. B828, 317–374 (2010). arXiv: 0807.1095 [hep-th]; Drummond, J.M., Henn, J., Smirnov, V.A., Sokatchev, E.: Magic identities for conformal four-point integrals. J. High Energy Phys. 0701, 064 (2007). hep-th/0607160v3

    Google Scholar 

  17. Drummond, J., Duhr, C., Eden, B., Heslop, P., Pennington, J., Smirnov, V.A.: Leading singularities and off shell conformal integrals. J. High Energy Phys. 1308, 133 (2013). arXiv:1303.6909v2 [hep-th]

    Google Scholar 

  18. Duhr, C., Gangl, H., Rhodes, J.R.: From polygons and symbols to polylogarithmic functions, 75 pp. (2011). arXiv:1110.0458 [math-ph]

    Google Scholar 

  19. Dyson, F.J.: Missed opportunities. Bull. Am. Math. Soc. 78(5), 635–652 (1972)

    Article  MATH  MathSciNet  Google Scholar 

  20. Gaiotto, D., Maldacena, J., Sever, A., Vieira, P.: Pulling the straps of polygons. J. High Energy Phys. 1112, 011 (2011). arXiv:1102.0062 [hep-th]

    Google Scholar 

  21. Goncharov, A., Spradlin, M., Vergu, C., Volovich, A.: Classical polylogarithms for amplitudes and Wilson loops. Phys. Rev. Lett. 105, 151605 (2010). arXiv:1006.5703v2 [hep-th]

    Google Scholar 

  22. Kinoshita, T.: Fine structure constant, electron anomalous magnetic moment, and quantum electrodynamics (2010). www.riken.jp/lab-www/theory/colloquium.pdf

  23. Lalin, M.N.: Polylogarithms and hyperbolic volumes, Fribourg (2007)

    Google Scholar 

  24. Milnor, J.W.: Hyperbolic geometry: the first 150 years. Bull. Am. Math. Soc. 6, 1 (1982)

    Article  MathSciNet  Google Scholar 

  25. Nikolov, N.M., Stora, R., Todorov, I.: Euclidean configuration space renormalization, residues and dilation anomaly. In: Dobrev, V. (ed.) Proceedings of the Varna Workshop Lie Theory and Its Applications in Physics (LT9), pp. 127–147. Springer, Tokyo/Heidelberg (2013). CERN-TH-PH/2012-076

    Google Scholar 

  26. Nikolov, N.M., Stora, R., Todorov, I.: Renormalization of massless Feynman amplitudes in configuration space. Rev. Math. Phys. 26(4), 143002 (2014). arXiv:1307.6854 [hep-th]

    Google Scholar 

  27. Remiddi, E., Vermaseren, J.A.M.: Harmonic polylogarithms. Int. J. Mod. Phys. A15, 725–754 (2000). arXiv:hep-ph/9905237

    Google Scholar 

  28. Schnetz, O.: Quantum periods: a census of ϕ 4-transcendentals. Commun. Number Theory Phys. 4(1), 1–46 (2010). arXiv:0801.2856v2 [hep-th]

    Google Scholar 

  29. Schnetz, O.: (2014). http://www2.mathematik.hu-berlin.de/~kreimer/tools/

  30. Schnetz, O.: Graphical functions and single-valued multiple polylogarithms (2013). arXiv:1302.6445 [math.NT]

    Google Scholar 

  31. Styer, D.: Calculation of the anomalous magnetic moment of the electron (2012). www.oberlin.edu/physics/dstyer/StrangeQM/Moment.pdf

  32. Todorov, I.: Studying quantum field theory. Bulg. J. Phys. 42, 93–114 (2013). arXiv:1311.7258 [math-ph], IHES/P/13/38

    Google Scholar 

  33. Ussyukina, N.I., Davydychev, A.I.: An approach to the evaluation of three- and four-point ladder diagrams. Phys. Lett. B298, 363–370 (1993); Exact results for three- and four-point ladder diagrams with an arbitrary number of rungs. Phys. Lett. B305, 136–143 (1993)

    Google Scholar 

  34. Waldschmidt, M.: Lectures on multiple zeta values IMSC 2011, Updated September 29 (2012)

    Google Scholar 

  35. Zagier, D.: The dilogarithm function. In: Cartier, P., Julia, B., Moussa, P., Vanhove, P. (eds.) Frontiers in Number Theory, Physics and Geometry II, pp. 3–65. Springer, Berlin (2006)

    Google Scholar 

Download references

Acknowledgements

It is a pleasure to thank Francis Brown, Pierre Cartier and Oliver Schnetz for enlightening discussions and pertinent remarks. The author thanks IHES for hospitality and support during the course of this work and Cécile Gourgues for her expert and efficient help.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ivan Todorov .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Computation of the Integral (5)

Using conformal invariance we can send the variable x 1 to infinity, x 4 to zero, x 2—to a unit 4-vector e and set

$$\displaystyle{ x_{3} = Z\,,\quad \mathrm{where}\quad Z^{2} = z\,\bar{z}\,,\ 2Ze = z +\bar{ z} }$$
(54)

so that the cross ratios (6) assume the form

$$\displaystyle{ u = Z^{2} = z\,\bar{z}\,,\quad v = (Z - e)^{2} = (z - 1)(\bar{z} - 1) }$$
(55)

in accord with (7). Then we can write, introducing spherical coordinates x = r ω, \(Z =\vert z\vert \,\omega _{z}\),

$$\displaystyle\begin{array}{rcl} \mathcal{F}(u,v)& =& F(z) = \frac{1} {\pi ^{2}} \int \frac{d^{4}\,x} {x^{2}(x - e)^{2}(x - Z)^{2}} \\ & =& \frac{1} {\pi ^{2}} \int ^{\infty }r\,dr\int _{ \mathbb{S}^{3}} \frac{d^{3}\,\omega } {(r^{2} - 2r\,e \cdot \omega +1)(r^{2} +\vert z\vert ^{2} - 2r\,\vert z\vert \,\omega \,\omega _{z})}\,.{}\end{array}$$
(56)

Assuming \(\vert z\vert < 1\) we can split the radial integral F into three terms \(F = F_{1} + F_{2} + F_{3}\) corresponding to the domains \(r <\vert z\vert\), \(\vert z\vert < r < 1\) and r > 1, respectively. In the first one we can write

$$\displaystyle\begin{array}{rcl} (r^{2} - 2r\,e \cdot \omega +1)^{-1}& =& \sum _{ n=0}^{\infty }r^{n}\,C_{ n}^{1}(\omega e)\,, \\ (r^{2} +\vert z\vert ^{2} - 2r\vert z\vert \,\omega \,\omega _{ z})^{-1}& =& \frac{1} {\vert z\vert ^{2}}\sum _{m=0}^{\infty }\left (\frac{r} {\vert z\vert }\right )^{\!\!m}C_{ m}^{1}(\omega \,\omega _{ z})\ (\mathrm{for}\ r <\vert z\vert < 1){}\end{array}$$
(57)

where the hyperspherical (Gegenbauer) polynomials C n 1 can be written as

$$\displaystyle{ C_{n}^{1}(\cos \theta ) = \frac{\sin (n + 1)\,\theta } {\sin \theta } \,. }$$
(58)

Using further the orthogonality relation

$$\displaystyle{ \int _{\mathbb{S}^{3}}C_{n}^{1}(\omega \cdot e)\,C_{ m}^{1}(\omega \,w_{ z})\,\frac{d^{3}\omega } {\pi ^{2}} = \frac{2\,\delta _{mn}} {n + 1}\,C_{n}^{1}(\omega _{ z}\,e) }$$
(59)

where, according to (55)

$$\displaystyle{ \omega _{z} \cdot e = \frac{z +\bar{ z}} {2\,\vert z\vert } \,. }$$
(60)

Inserting in F 1 and using (58) (or (8)) and (11) we find

$$\displaystyle{ F_{1}(z) =\int _{ 0}^{\vert z\vert }\frac{r\,dr} {\vert z\vert ^{2}} \ \sum _{n=0}^{\infty }\ \frac{2} {n + 1}\,\frac{r^{2n}} {\vert z\vert ^{n}} \,C_{n}^{1}\left (\frac{z +\bar{ z}} {2\vert z\vert } \right ) = \frac{Li_{2}(z) - Li_{2}(\bar{z})} {z -\bar{ z}} \,. }$$
(61)

The same result is obtained for F 3(z):

$$\displaystyle{ F_{3}(z) =\int _{ 1}^{\infty }\frac{dr} {r^{3}} \ \sum _{n=0}^{\infty }\ \frac{2} {n + 1}\, \frac{\vert z\vert ^{n}} {r^{2n}}\,C_{n}^{1}\left (\frac{z +\bar{ z}} {2\vert z\vert } \right ) = \frac{Li_{2}(z) - Li_{2}(\bar{z})} {z -\bar{ z}} = F_{1}(z)\,. }$$
(62)

Finally,

$$\displaystyle{ F_{2}(z) = 2\int _{\vert z\vert }^{1}\frac{dr} {r} \ \frac{Li_{1}(z) - Li_{1}(\bar{z})} {z -\bar{ z}} =\ln z\,\bar{z}\,\frac{\ln (1 - z) -\ln (1 -\bar{ z})} {z -\bar{ z}} \,; }$$
(63)

this, together with (61), (62) completes the proof of (9) (10) for \(\vert z\vert < 1\). The same expression can be obtained in a similar fashion for \(\vert z\vert > 1\); alternatively, it can be deduced from the result for \(\vert z\vert < 1\) using the symmetry of F(z) implied by (14). The result can also be established by verifying that it is single valued and satisfies the first equation (39) (in view of the uniqueness of SVMP,Theorem 3.1; cf. [30]).

Appendix 2: Identities Among MZV

Equation (22) which relates the MZV ζ w (labeled by words in the two letters {0, 1}) with \(\zeta (n_{1},\ldots,n_{r})\), \(n_{i} = 1,2,\ldots\) becomes particularly simple for words of depth one,

$$\displaystyle{ \zeta _{0^{n_{0}}10^{n_{1}-1}} = (-1)^{n_{0}+1}\left (\begin{array}{*{10}c} n_{0} + n_{1} - 1 \\ n_{1} - 1 \end{array} \right )\zeta (n_{0}+n_{1})\,. }$$
(64)

This allows to write the depth one contribution to the generating function Z (29) in terms of multiple commutators:

$$\displaystyle{ \sum _{n=1}^{\infty }\zeta (n+1)\mathop{\underbrace{[\ldots [}}\limits _{ n}e_{0},e_{1}],e_{0}],\ldots,e_{0}] =\!\sum _{ n=1}^{\infty }\zeta (n+1)\!\sum _{ k=0}^{n}(-1)^{k+1}\left (\begin{array}{*{10}c} n\\ k \end{array} \right )e_{ 0}^{k}\,e_{ 1}\,e_{0}^{n-k}, }$$
(65)

which is another way to write down (64). It is more interesting—and more difficult— to deduce the relations among ζ w for words of higher depth. We shall write down all such relations for depth two and weight \(\vert w\vert \leq 5\). Note that the number d n of linearly independent MZV of a given weight n can be read off the generating function conjectured by Don Zagier

$$\displaystyle{ \frac{1} {1 - t^{2} - t^{3}} =\sum _{ n=0}^{\infty }d_{ n}\,t^{n}\quad d_{ 2} = d_{3} = d_{4} = 1\quad d_{5} = d_{6} = 2,\ldots }$$
(66)

(and proven for the motivic analog of MZV by Brown [6]; in general, d n provide an upper bound of the independent MZV).

The Euler’s relation (27) is a special case of either of the following more general relations which only involve proper (convergent) zeta series:

$$\displaystyle{ \zeta (\mathop{\underbrace{1,\ldots,1}}\limits _{n-2},2) =\zeta (n)\,, }$$
(67a)
$$\displaystyle{ \sum _{{ s_{i}\geq 1;s_{k}\geq 2 \atop \varSigma s_{i}=n} }\zeta (s_{1},\ldots,s_{k}) =\zeta (n)\,. }$$
(67b)

The “improper” (regularized) zeta value ζ(n, 1) is determined from the stuffle relation:

$$\displaystyle{ 0 =\zeta (1)\,\zeta (n) =\zeta (1,n) +\zeta (n,1) +\zeta (n + 1)\,. }$$
(68)

In particular, for n = 2, we find

$$\displaystyle{ \zeta (2,1) = -\zeta (3) -\zeta (1,2) = -2\,\zeta (3)\,. }$$
(69)

From Euler’s formula

$$\displaystyle{ \zeta (2) = \frac{\pi ^{2}} {6} }$$
(70)

(a special case of (28)) and from the shuffle and stuffle relations one deduces that all zeta values of weight four are rational multiples of π 4 (in accord with the Zagier conjecture (66)). In particular, the relations for , \(10 \times 10\) and (67) for n = 4,

$$\displaystyle{\zeta (2)^{2} =\zeta _{ 10\times 10} = 2\,\zeta (2,2) +\zeta (4)\,;\quad \zeta (1,3) +\zeta (2,2) =\zeta (4)\,,}$$

allow to express all weight four words of length not exceeding two as integer multiples of \(\zeta (1,3)\):

$$\displaystyle{\zeta (4) = 4\,\zeta (1,3)(=\zeta (1,1,2))\,,\ \zeta (2,2) = 3\,\zeta (1,3)\,,\ \zeta (2)^{2} = 10\,\zeta (1,3)}$$
$$\displaystyle{ \Rightarrow \zeta (1,3) = \frac{\pi ^{4}} {360}\,. }$$
(71)

Proceeding in a similar fashion with the two products of the words 10 and 100 we find

$$\displaystyle{\zeta (2)\,\zeta (3) = 3\,\zeta _{10100} + 6\,\zeta _{11000} +\zeta _{10010} = 3\,\zeta (2,3) + 6\,\zeta (1,4) +\zeta (3,2)\,,}$$
$$\displaystyle{\zeta (2)\,\zeta (3) =\zeta (2,3) +\zeta (3,2) +\zeta (5)\,;\quad \zeta (1,4) +\zeta (2,3) +\zeta (3,2) =\zeta (5)\,.}$$

These three equations determine a two-dimensional space of zeta values of weight five (in accord with (66)). Selecting as a basis ζ(1, 4) and ζ(2, 3) we express the remaining convergent ζ-values of weight 5 in terms of this basis with positive integer coefficients

$$\displaystyle{\zeta (1,1,3) =\zeta (1,4)\,,\ \zeta (1,2,2) =\zeta (2,3)\,,}$$
$$\displaystyle{\zeta (5) = 2\,\zeta (2,3) + 6\,\zeta (1,4)\,,\ \zeta (3,2) =\zeta (2,1,2) =\zeta (2,3) + 5\,\zeta (1,4)\,,}$$
$$\displaystyle{ \zeta (2)\,\zeta (3) = 4\,\zeta (2,3) + 11\,\zeta (1,4) }$$
(72)

(while \(\zeta (4,1) = -\zeta (1,4) -\zeta (5) = -7\,\zeta (1,4) - 2\,\zeta (2,3)\)).

For the study of single valued MZV it is more natural to use the basis \((\zeta (5),\zeta (2)\,\zeta (3))\) instead. Then we find

$$\displaystyle{(\zeta (1,1,3) =)\,\zeta (1,4) = 2\,\zeta (5) -\zeta (2)\,\zeta (3)\,,}$$
$$\displaystyle{(\zeta (1,2,2) =)\,\zeta (2,3) = 3\,\zeta (2)\,\zeta (3) -\frac{11} {2} \,\zeta (5)}$$
$$\displaystyle{ (\zeta (2,1,2) =)\,\zeta (3,2) = \frac{9} {2}\,\zeta (5) - 2\,\zeta (2)\,\zeta (3)\,;\ \zeta (4,1) =\zeta (2)\,\zeta (3) - 3\,\zeta (5)\,. }$$
(73)

Brown [6] has demonstrated that a basis for “motivic” MZV for all weights is given by \(\zeta (s_{1},\ldots,s_{k})\), with s i  ∈ { 2, 3}.

From the iterated integral representation of MZV it follows that the generating function (29) satisfies:

$$\displaystyle{ Z_{e_{0}e_{1}}^{-1} = Z_{ e_{1}e_{0}} =\tilde{ Z}_{-e_{0},-e_{1}}\,. }$$
(74)

(The first equation incorporates, in particular, (67a).)

Appendix 3: Monodromy at z = 1: Single Valued MZV

The representation (33) can be obtained from (32) by noticing that the substitution z → 1 − z corresponds to the exchange \(e_{0} \leftrightarrow e_{1}\) and that the path from 0 to z can be viewed as a composition of two paths: from 0 to 1 and from 1 to z. For 0 < z < 1 one should just set \(h_{1}(z) = h_{0}(1 - z)\). Equation (30) follows from (32) (33) and the relations

$$\displaystyle{ \mathcal{M}_{0}\ln z =\ln z + 2\pi i\,,\quad \mathcal{M}_{1}\ln (1 - z) =\ln (1 - z) + 2\pi i\,. }$$
(75)

Applying (30) one should take into account the relation (74)

$$\displaystyle{ Z_{e_{0},e_{1}}^{-1} = Z_{ e_{1},e_{0}} =\tilde{ Z}_{-e_{0},-e_{1}} }$$
(76)

where the tilde indicates that each word is replaced by its opposite. We leave it to the reader to verify that the first few terms in the expansion of (30) reproduce (75) and give

$$\displaystyle{\mathcal{M}_{1}\,L_{01}(z) = L_{01}(z)(=\ln z\ln (1 - z) + Li_{2}(z))}$$
$$\displaystyle{ \mathcal{M}_{1}\,L_{10}(z)(= \mathcal{M}_{1}(-Li_{2}(z))) = L_{10}(z) + 2\pi i\ln z\,. }$$
(77)

We now proceed to the evaluation of the element e 1 defined by Eq. (47). To this end we introduce the Lie algebra valued function

$$\displaystyle{ F(e_{0},e_{1}) = Z_{e_{0}e_{1}}e_{1}Z_{e_{0}e_{1}}^{-1} - e_{ 1} =\zeta (2)[[e_{0},e_{1}],e_{1}] +\zeta (3)[[[e_{0},e_{1}],e_{1}],e_{0} + e_{1}]+\ldots }$$
(78)

Equation (47) can then be solved recursively, writing \(e_{1}^{{\prime}} =\mathop{\lim }\limits_{ k \rightarrow \infty }\,e_{1}^{(k)}\) with

$$\displaystyle{ e_{1}^{(0)} = e_{ 1}\,,\ e_{1}^{(k+1)} = e_{ 1} + F(e_{0},e_{1}) + F_{0}(-e_{0},-e_{1}^{(k)})\,. }$$
(79)

The weight three term with ζ(2) cancels out and one finds

$$\displaystyle{ e_{1}^{{\prime}} = e_{ 1} + 2\,\zeta (3)\,[[[e_{0},e_{1}],e_{1}],e_{0} + e_{1}] +\zeta (5)(\ldots )+\ldots }$$
(80)

where, according to Schnetz [30], the ζ(5) contribution consists of eight bracket words of weight six. (The ζ(3) contribution will be sufficient to the application that follows.)

The SVMPs in the right hand side of (51) are obtained from those in g 3(z) by adding a letter 0 in front and at the end of each labeling word. Evaluating the regularized limit at z = 1 (and noting that for L 11(z) it is zero) while \(\bar{L}_{01}(1) = -\bar{L}_{10}(1) =\zeta (2)\) we find that for each (5-letter) word-label w in (51) we obtain the following counterpart of (50)

$$\displaystyle{P_{w}(1) = P_{w}^{0}(1) + 2\,\zeta (2)\,\zeta (3)\,\langle w,w_{ 23}\rangle }$$
$$\displaystyle{ w_{23}:= \left [e_{0},[[[e_{0},e_{1}],e_{1}],e_{0} + e_{1}]\right ]\,. }$$
(81)

We shall see that the role of the second term in the right hand side of (81) is to cancel the product ζ(2) ζ(3) in P w 0(1), in accord with the observation that ζ SV(2) = 0.

Indeed the depth one contributions are proportional to ζ(5):

$$\displaystyle{P_{0^{3}10}(1)(= P_{0^{3}10}^{0}(1)) = L_{ 0^{3}10}(1) + L_{010^{3}}(1) = 8\,\zeta (5)\,,}$$
$$\displaystyle{P_{0^{2}10^{2}}(1) = 2\,L_{0^{2}10^{2}}(1) = -12\,\zeta (5)}$$

and their difference reproduces (52). For depth two we find (after cancelling the products ζ(2) ζ(3)) a negative multiple of ζ(5):

$$\displaystyle\begin{array}{rcl} P_{0^{2}101}^{0}(1)& =& \zeta _{ 0^{2}101} +\zeta _{1010^{2}} +\zeta _{100}\,\zeta _{01} {}\\ & =& 3\,\zeta (4,1) + 2\,\zeta (3,2) + 2\,\zeta (2,3) -\zeta (2)\,\zeta (3) = 4\,\zeta (2)\,\zeta (3) - 11\,\zeta (5)\,, {}\\ \end{array}$$

where in the last step we used (73), \(\langle 0^{2}101,w_{23}\rangle = -2\) so that \(P_{0^{2}101}(1) = P_{0^{2}101}^{0} + 2\,\zeta (2)\,\zeta (3)\langle 0^{2}101,w_{23}\rangle = -11\,\zeta (5)\); similarly \(P_{01010}(1) = 4\,\zeta (5) = P_{0^{3}1^{2}}(1)\), \(P_{01^{2}0^{2}}(1) = -\zeta (5)\), so that

$$\displaystyle{ P_{0^{2}101}(1) - P_{01010}(1) + P_{01^{2}0^{2}}(1) - P_{0^{3}1^{2}}(1) = -20\,\zeta (5)\,. }$$
(82)

Finally, the depth three contribution is equal to that of depth one. Indeed we find, using [7],

$$\displaystyle{P_{0101^{2}}(1) =\zeta _{0101^{2}} +\zeta _{1^{2}010} +\zeta _{10}\,\zeta _{01^{2}} + 6\,\zeta (2)\,\zeta (3) = 11\,\zeta (5)\,,P_{01^{2}01}(1) = -9\,\zeta (5)}$$
$$\displaystyle{ \Rightarrow P_{0101^{2}}(1) - P_{01^{2}01}(1) = 20\,\zeta (5)\,. }$$
(83)

This completes the proof of (53). (The expressions (82) and (83) can be also extracted from the polylog- and polyzeta-procedures of [29].)

Rights and permissions

Reprints and permissions

Copyright information

© 2014 Springer Japan

About this paper

Cite this paper

Todorov, I. (2014). Polylogarithms and Multizeta Values in Massless Feynman Amplitudes. In: Dobrev, V. (eds) Lie Theory and Its Applications in Physics. Springer Proceedings in Mathematics & Statistics, vol 111. Springer, Tokyo. https://doi.org/10.1007/978-4-431-55285-7_10

Download citation

Publish with us

Policies and ethics