You are on page 1of 7

Integral representations

Any of the following integral representations furnishes the analytic continuation of the polylogarithm beyond the circle of convergence |z| = 1 of the defining power series. 1. The polylogarithm can be expressed in term of the integral of the BoseEinstein distribution:

This converges for Re(s) > 0 and all z except for z real and 1. The polylogarithm in this context is sometimes referred to as a Bose integral or a BoseEinstein integral. Similarly, the polylogarithm can be expressed in terms of the integral of the FermiDirac distribution:

This converges for Re(s) > 0 and all z except for z real and 1. The polylogarithm in this context is sometimes referred to as a Fermi integral or a FermiDirac integral (GSL 2010). These representations are readily verified by Taylor expansion of the integrand with respect to z and termwise integration. 2. A complementary integral representation applies to Re(s) < 0 and to all z except to z real and 0:

This integral follows from the general relation of the polylogarithm with the Hurwitz zeta function (see above) and a familiar integral representation of the latter. 3. The polylogarithm may be quite generally represented by a Hankel contour integral (Whittaker & Watson 1927, 12.22, 13.13), which extends the BoseEinstein representation to negative orders s. As long as the t = pole of the integrand does not lie on the non-negative real axis, and s 1, 2, 3, ..., we have:

where H represents the Hankel contour. The integrand has a cut along the real axis from zero to infinity, with the axis belonging to the lower half plane of t. The integration starts at + on the upper half plane (Im(t) > 0), circles the origin without enclosing any of the poles t = + 2ki, and terminates at + on the lower half plane (Im(t) < 0). For the case where is real and nonnegative, we can simply subtract the contribution of the enclosed t = pole:

where R is the residue of the pole:

4. When the AbelPlana formula is applied to the defining series of the polylogarithm, a Hermite-type integral representation results that is valid for all complex z and for all complex s:

where is the upper incomplete gamma-function. Note that all (but not part) of the ln(z) in this expression can be replaced by ln(1z). A related representation which also holds for all complex s,

avoids the use of the incomplete gamma function, but this integral fails for z on the positive real axis if Re(s) 0. This expression is found by writing 2s Lis(z) / (z) = (z2, s, 12) z (z2, s, 1), where is the Lerch transcendent, and applying the AbelPlana formula to the first series and a complementary formula that involves 1 / (e2t + 1) in place of 1 / (e2t 1) to the second series.

Series representations
1. As noted under integral representations above, the BoseEinstein integral representation of the polylogarithm may be extended to negative orders s by means of Hankel contour integration:

where H is the Hankel contour, s 1, 2, 3, ..., and the t = pole of the integrand does not lie on the non-negative real axis. The contour can be modified so that it encloses the poles of the integrand at t = 2ki, and the integral can be evaluated as the sum of the residues (Wood 1992, 12, 13; Gradshteyn & Ryzhik 1980, 9.553):

This will hold for Re(s) < 0 and all except where e = 1. For 0 < Im() 2 the sum can be split as:

where the two series can now be identified with the Hurwitz zeta function:

This relation, which has already been given under relationship to other functions above, holds for all complex s 0, 1, 2, 3, ... and was first derived in (Jonquire 1889, eq. 6). 2. In order to represent the polylogarithm as a power series about = 0, we write the series derived from the Hankel contour integral as:

When the binomial powers in the sum are expanded about = 0 and the order of summation is reversed, the sum over h can be expressed in closed form:

This result holds for || < 2 and, thanks to the analytic continuation provided by the zeta functions, for all s 1, 2, 3, ... . If the order is a positive integer, s = n, both the term with k = n 1 and the gamma function become infinite, although their sum does not. One obtains (Wood 1992, 9; Gradshteyn & Ryzhik 1980, 9.554):

where the sum over h vanishes if k = 0. So, for positive integer orders and for || < 2 we have the series:

where Hn denotes the nth harmonic number:

The problem terms now contain ln() which, when multiplied by n1, will tend to zero as 0, except for n = 1. This reflects the fact that Lis(z) exhibits a true logarithmic singularity at s = 1 and z = 1 since:

For s close, but not equal, to a positive integer, the divergent terms in the expansion about = 0 can be expected to cause computational difficulties (Wood 1992, 9). Note also that Erdlyi's corresponding expansion (Erdlyi et al. 1981, 1.11-15) in powers of ln(z) is not correct if one

assumes that the principal branches of the polylogarithm and the logarithm are used simultaneously, since ln(1z) is not uniformly equal to ln(z). For nonpositive integer values of s, the zeta function (s k) in the expansion about = 0 reduces to Bernoulli numbers: (n k) = B1+n+k / (1 + n + k). Numerical evaluation of Lin(z) by this series does not suffer from the cancellation effects that the finite rational expressions given under particular values above exhibit for large n. 3. By use of the identity

the BoseEinstein integral representation of the polylogarithm (see above) may be cast in the form:

Replacing the hyperbolic cotangent with a bilateral series,

then reversing the order of integral and sum, and finally identifying the summands with an integral representation of the upper incomplete gamma function, one obtains:

For both the bilateral series of this result and that for the hyperbolic cotangent, symmetric partial sums from kmax to kmax converge unconditionally as kmax . Provided the summation is performed symmetrically, this series for Lis(z) thus holds for all complex s as well as all complex z. 4. Introducing an explicit expression for the Stirling numbers of the second kind into the finite sum for the polylogarithm of nonpositive integer order (see above) one may write:

The infinite series obtained by simply extending the outer summation to (Guillera & Sondow 2008, Theorem 2.1):

turns out to converge to the polylogarithm for all complex s and for complex z with Re(z) < 12, as can be verified for |z(1z)| < 12 by reversing the order of summation and using:

For the other arguments with Re(z) < 12 the result follows by analytic continuation. This procedure is equivalent to applying Euler's transformation to the series in z that defines the polylogarithm.

Asymptotic expansions
For |z| 1, the polylogarithm can be expanded into asymptotic series in terms of ln(z):

where B2k are the Bernoulli numbers. Both versions hold for all s and for any arg(z). As usual, the summation should be terminated when the terms start growing in magnitude. For negative integer s, the expansions vanish entirely; for non-negative integer s, they break off after a finite number of terms. Wood (1992, 11) describes a method for obtaining these series from the BoseEinstein integral representation (note that his equation 11.2 for Lis(e) requires 2 < Im() 0).

Limiting behavior
The following limits result from the various representations of the polylogarithm (Wood 1992, 22):

Note that Wood's first limit for Re() has been corrected in accordance with his equation 11.3. The limit for Re(s) follows from the general relation of the polylogarithm with the Hurwitz zeta function (see above).

Dilogarithm
Main article: Spence's function The dilogarithm is just the polylogarithm of order s = 2. An alternate integral expression of the dilogarithm for arbitrary complex argument z is (Abramowitz & Stegun 1972, 27.7):

A source of confusion is that some computer algebra systems define the dilogarithm as dilog(z) = Li2(1z). In the case of real z 1 the first integral expression for the dilogarithm can be written as

from which expanding ln(t1) and integrating term by term we obtain

The Abel identity for the dilogarithm is given by (Abel 1881)

This is immediately seen to hold for either x = 0 or y = 0, and for general arguments is then easily verified by differentiation /x /y. For y = 1x the identity reduces to Euler's reflection formula

where Li2(1) = (2) = 16 2 has been used and x may take any complex value. In terms of the new variables u = x/(1y), v = y/(1x) the Abel identity reads

which corresponds to the pentagon identity given in (Rogers 1907). From the Abel identity for x = y = 1z and the square relationship we have Landen's identity

and applying the reflection formula to each dilogarithm we find the inversion formula

and for real z 1 also

Known closed-form evaluations of the dilogarithm at special arguments are collected in the table below. Arguments in the first column are related by reflection x 1x or inversion x 1x to either x = 0 or x = 1; arguments in the third column are all interrelated by these operations. Historical note: Maximon (2003) discusses the 17th to 19th century references. The reflection formula was already published by Landen in 1760, prior to its appearance in a 1768 book by Euler (Maximon 2003, 10); an equivalent to Abel's identity was already published by Spence in 1809, before Abel wrote his manuscript in 1826 (Zagier 1989, 2). The designation bilogarithmische Function was introduced by Carl Johan Danielsson Hill (professor in Lund, Sweden) in 1828 (Maximon 2003, 10). Don Zagier (1989) has remarked that the dilogarithm is the only mathematical function possessing a sense of humor. Special values of the dilogarithm

Here

denotes the golden ratio.

You might also like