The Voronoi summation formula

Jordan Bell
September 16, 2015

1 Mellin transform

The Mellin transform of f:(0,) is defined by

(f)(s)=0xs-1f(x)𝑑x.

For example, sΓ(s) is the Mellin transform of xe-x.

Suppose that f continuous on (0,), that there is some α such that f(x)=O(x-α) as x0, and that for any n1, f(x)xn0 as x. Then [4, p. 107, Proposition 9.7.7] (f)(s) is holomorphic on (s)>α, and for σ>α and x>0,

f(x)=12πi(s)=σx-s(f)(s)𝑑s.

(The Mellin inversion formula.)

2 Generalized Poisson summation formula

Cohen [4, pp. 177–182, §10.2.5] presents a “generalized Poisson summation formula” which yields both the Poisson summation formula and the Voronoi summation formula.

We denote by 𝒮() the Fréchet space of Schwartz functions .

Theorem 1.

Let a be arithmetic function and define

L(a,s)=n=1a(n)n-s,(s)>1.

Suppose that L(a,s) has an analytic continuation to whose only possible pole is at s=1. Suppose also that there are A,a1,,ag>0 such that for

γ(s)=Asj=1gΓ(ajs),

L(a,s) satisfies the functional equation

γ(s)L(a,s)=γ(1-s)L(a,1-s).

Let f𝒮() and define for x>0,

K(x)=12πi(s)=32γ(s)γ(1-s)x-s𝑑s,g(x)=0f(y)K(xy)𝑑y.

Then,

n=1a(n)f(n)=f(0)L(a,0)+Ress=1(f)(s)L(a,s)+n=1a(n)g(n).
Proof.

Since f is a Schwartz function, (f) is holomorphic on (s)>0. Futhermore, for (s)>0, integrating by parts,

(f)(s)=0xs-1f(x)𝑑x=f(x)xss|0-0f(x)xss𝑑x=-1s(f)(s+1).

It follows that (f) has an analytic continuation to possibly with poles at 0,-1,-2,-3,. Write F=(f). By the Mellin inversion formula we get

n=1a(n)f(n) =n=1a(n)12πi(s)=32n-sF(s)𝑑s
=12πi(s)=32F(s)n=1ann-sds
=12πi(s)=32F(s)L(a,s)𝑑s.

The only possible pole of L(a,s) is at s=1. From

(f)(s)=-1s(f)(s+1),

the only possible pole of F(s) in the half-plane (s)>-1 is at s=0, and the residue of F(s)L(a,s) at s=0 is

-(f)(1)=-0f(x)𝑑x=-(f()-f(0))=f(0),

so the residue of F(s)L(a,s) at s=0 is

f(0)L(a,0).

Therefore, by the residue theorem, taking as given that F(s)L(a,s)0 uniformly in -12(s)32 as |(s)|,

n=1a(n)f(n) =f(0)L(a,0)+Ress=1F(s)L(a,s)+12πi(s)=-12F(s)L(a,s)𝑑s.

Define

G(s)=F(1-s)γ(s)γ(1-s).

Using the functional equation for L(a,s),

12πi(s)=-12F(s)L(a,s)𝑑s =12πi(s)=-12F(s)γ(1-s)γ(s)L(a,1-s)𝑑s
=12πi(s)=32F(1-s)γ(s)γ(1-s)L(a,s)𝑑s
=12πi(s)=32G(s)L(a,s)𝑑s.

Furthermore, define

J(x)=12πi(s)=3211-sγ(s)γ(1-s)x1-s𝑑s,

which satisfies

J(x)=K(x).

We have

12πi(s)=32x-sG(s)𝑑s =12πi(s)=32x-sF(1-s)γ(s)γ(1-s)𝑑s
=12πi(s)=32x-s(-11-s(f)(2-s))γ(s)γ(1-s)𝑑s
=12πi(s)=32x-s(-11-s0y1-sf(y)𝑑y)γ(s)γ(1-s)𝑑s
=-1x0f(y)12πi(s)=3211-sγ(s)γ(1-s)(xy)1-s𝑑s
=-1x0f(y)J(xy)𝑑y
=-1xf(y)J(xy)|0+1x0f(y)J(xy)x𝑑y
=0+0f(y)J(xy)𝑑y
=0f(y)K(xy)𝑑y
=g(x).

Therefore,

n=1a(n)g(n) =n=1a(n)12πi(s)=32n-sG(s)𝑑s
=12πi(s)=32G(s)n=1a(n)n-sds
=12πi(s)=32G(s)L(a,s)𝑑s.

Thus we have

n=1a(n)f(n)=f(0)L(a,0)+Ress=1F(s)L(a,s)+n=1a(n)g(n)

Take a(n)=1 for all n. Then,

L(a,s)=n=1n-s=ζ(s).

The Riemann zeta function satisfies the functional equation

π-s/2Γ(s2)ζ(s)=π-(1-s)/2Γ(1-s2)ζ(1-s).

So with

γ(s)=π-s/2Γ(s2),

we have

γ(s)ζ(s)=γ(1-s)ζ(1-s).

Using

Γ(1-z)Γ(z)=πsinπz

and

Γ(z)Γ(z+12)=21-2zπΓ(2z),

we have

Γ(1-s2) =Γ(1-s+12)
=πsinπ(s+1)2Γ(s+12)
=πsinπ(s+1)2Γ(s2+12)
=πΓ(s2)sinπ(s+1)221-sπΓ(s),

and so

π-s/2Γ(s2)π-(1-s)/2Γ(1-s2) =π-s+12Γ(s2)sinπ(s+1)221-sπΓ(s)πΓ(s2)
=sinπ(s+1)22(2π)-sΓ(s)
=cosπs22(2π)-sΓ(s).

Therefore

K(x)=12πi(s)=32γ(s)γ(1-s)x-s𝑑s=12πi(s)=32cosπs22(2π)-sΓ(s)x-s𝑑s

But, taking as known

0cos(2πx)xs-1𝑑x=(2π)-scosπs2Γ(s),

it follows that

K(x)=2cos2πx.

Thus Theorem 1 tells us that for f𝒮(),

n=1f(n)=f(0)ζ(0)+Ress=1(f)(s)ζ(s)+2n=10f(y)cos(2πny)𝑑y,

i.e.,

n=1f(n)=-12f(0)+0f(x)𝑑x+2n=10f(y)cos(2πny)𝑑y.

If f: is even, this is the Poisson summation formula.

Take a(n)=d(n) for all n. Then,

L(d,s)=n=1d(n)n-s=ζ2(s).

For

γ(s)=π-sΓ(s2)2,

it follows from the functional equation for the Riemann zeta function that L(d,s) satisfies the functional equation

γ(s)L(d,s)=γ(1-s)L(d,1-s).

We worked out above that

π-s/2Γ(s2)π-(1-s)/2Γ(1-s2)=cosπs22(2π)-sΓ(s),

whence

γ(s)γ(1-s) =(2π)-2s4cos2πs2Γ(s)2
=(2π)-2s(2+2cosπs)Γ(s)2.

Taking as given two identities for Bessel functions

0xs-1K0(4πx1/2)𝑑x=12(2π)-2sΓ(s)2

and

0xs-1Y0(4πx1/2)𝑑x=-1π(2π)-2scosπsΓ(s)2,

it follows that

K(x)=4K0(4πx1/2)-2πY0(4πx1/2).

Thus Theorem 1 tells us that for f𝒮(),

n=1d(n)f(n) =f(0)ζ2(0)+Ress=1(f)(s)ζ2(s)
+n=1d(n)0f(y)(4K0(4π(ny)1/2)-2πY0(4π(ny)1/2))𝑑y.

Using

ζ2(s)=1(s-1)2+2γs-1+O(1),s1,

and

xs-1=1+(s-1)logx+O(|s-1|2),

we have

Ress=1(f)(s)ζ2(s)=2γ+logx,

and so

n=1d(n)f(n) =14f(0)+0f(x)(2γ+logx)𝑑x
+n=1d(n)0f(y)(4K0(4π(ny)1/2)-2πY0(4π(ny)1/2))𝑑y.

3 Bernoulli numbers

The Bernoulli polynomials are defined by

tetxet-1=m=0Bm(x)tmm!.

The Bernoulli numbers are defined by Bm=Bm(0).

We denote by [x] the greatest integer x, and we define {x}=x-[x], namely, the fractional part of x. We define Pm(x)=Bm({x}), the Bernoulli functions.

4 Wigert

The following result is proved by Wigert [18]. Our proof follows Titchmarsh [13, p. 163, Theorem 7.15]. Cf. Landau [10].

Theorem 2.

For λ<12π and N1,

n=1d(n)e-nz=γz-logzz+14-n=0N-1B2n+22(2n+2)!(2n+2)z2n+1+O(|z|2N)

as z0 in any angle |argz|λ.

Proof.

For σ>1, s=σ+it,

ζ2(s)=n=1d(n)ns.

Using this, for z>0 we have

12πi2-i2+iΓ(s)ζ2(s)z-s𝑑s =n=1d(n)12πi2-i2+iΓ(s)(nz)-s𝑑s
=n=1d(n)e-nz. (1)

Define F(s)=Γ(s)ζs(s)z-s. F has poles at 1,0, and the negative odd integers. (At each negative even integer, Γ has a first order pole but ζ2 has a second order zero.) First we determine the residue of F at 1. We use the asymptotic formula

ζ(s)=1s-1+γ+O(|s-1|),s1,

the asymptotic formula

Γ(s)=1-γ(s-1)+O(|s-1|2),s1,

and the asymptotic formula

z-s=1z-logzz(s-1)+O(|s-1|2),s1,

to obtain

Γ(s)ζs(s)z-s =(1-γ(s-1)+O(|s-1|2))(1(s-1)2+2γs-1+O(|s-1|2))
(1z-logzz(s-1)+O(|s-1|2))
=1z(s-1)2-γz(s-1)+2γz(s-1)-logzz(s-1)+O(1)
=1z(s-1)2+γz(s-1)-logzz(s-1)+O(1).

Hence the residue of F at 1 is

γz-logzz.

Now we determine the residue of F at 0. The residue of Γ at 0 is 1, and hence the residue of F at 0 is

1ζ2(0)z0=ζ2(0)=(-12)2=14.

Finally, for n0 we determine the residue of F at -(2n+1). The residue of Γ at -(2n+1) is (-1)2n+1(2n+1)!, hence the residue of F at -(2n+1) is

(-1)2n+1(2n+1)!ζ2(2n+1)z2n+1=-B2n+22(2n+2)!(2n+2)z2n+1

using

ζ(-m)=-Bm+1m+1,m1.

Let M>0, and let C be the rectangular path starting at 2-iM, then going to 2+iM, then going to -2N+iM, then going to -2N-iM, and then ending at 2-iM. By the residue theorem,

CF(s)𝑑s=2πi(γz-logzz+14+n=0N-1-B2n+22(2n+2)!(2n+2)z2n+1). (2)

Denote the right-hand sideof (2) by 2πiR. We have

CF(s)𝑑s=2-iM2+iMF(s)𝑑s+2+iM-2N+iMF(s)𝑑s+-2N+iM-2N-iMF(s)𝑑s+-2N-iM2-iMF(s)𝑑s.

We shall show that the second and fourth integrals tend to 0 as M. For s=σ+it with -2Nσ2, Stirling’s formula [14, p. 151] tells us that

|Γ(s)|2πe-π2|t||t|σ-12,|t|.

As well [13, p. 95], there is some K>0 such that in the half-plane σ-2N,

ζ(s)=O(|t|K).

Also,

z-s =e-slogz
=e-(σ+it)(log|z|+iargz)
=e-σlog|z|+targz-i(σargz+tlog|z|),

and so for |argz|λ,

|z-s|=e-σlog|z|+targze-σlog|z|+λ|t|=|z|-σeλ|t|.

Therefore

|2+iM-2N+iMF(s)𝑑s|(2+2N)sup-2Nσ2|F(σ+iM)|=O(e-π2MMσ-12M2K|z|-σeλM),

and because λ<π2 this tends to 0 as M. Likewise,

|-2N-iM2-iMF(s)𝑑s|0

as M. It follows that

2-i2+iF(s)𝑑s+-2N+i-2N-iF(s)𝑑s=2πiR.

Hence,

2-i2+iF(s)𝑑s=2πiR+-2N-i-2N+iF(s)𝑑s.

We bound the integral on the right-hand side. We have

-2N-i-2N+iF(s)𝑑s=σ=-2N,|t|1F(s)𝑑s+σ=-2N,|t|>1F(s)𝑑s.

The first integral satisfies

|σ=-2N,|t|1F(s)𝑑s|σ=-2N,|t|1|Γ(s)ζ2(s)||z|-σeλ|t|𝑑s=|z|2NO(1)=O(|z|2N),

because Γ(s)ζ2(s) is continuous on the path of integration. The second integral satisfies

|σ=-2N,|t|>1F(s)𝑑s| σ=-2N,|t|>1e-π2|t||t|σ-12|t|K|z|-σeλ|t|𝑑s
=|z|2Nσ=-2N,|t|>1e-π2|t||t|-2N-12|t|Keλ|t|𝑑t
=|z|2NO(1)
=O(|z|2N),

because λ<π2. This establishes

12πi2-i2+iF(s)𝑑s=R+O(|z|2N).

Using (1) and (2), this becomes

n=1d(n)e-nz=γz-logzz+14-n=0N-1B2n+22(2n+2)!(2n+2)z2n+1+O(|z|-2N),

completing the proof. ∎

For example, as B2=16,B4=-130,B6=142, the above theorem tells us that

n=1d(n)e-nz=γz-logzz+14-z144-z386400-z57620480+O(|z|6).

5 Other works on the Voronoi summation formula

Voronoi’s papers on the Voronoi summation formula are [15] and [17] and [16].

Iwaniec and Kowalski [9, Chaper 4]

Stein and Shakarchi [12, p. 392, Theorem 8.11].

Ivic [8, pp. 83ff., Chapter 3] and [7]

Miller and Schmid [11]

Hejhal [6]

Flajolet, Gourdon and Dumas [5]

Bettin and Conrey [1]

Chandrasekharan and Narasimhan [2]

Chandrasekharan [3, Chapter VIII]

References

  • [1] S. Bettin and B. Conrey (2013) Period functions and cotangent sums. Algebra Number Theory 7 (1), pp. 215–242. Cited by: §5.
  • [2] K. Chandrasekharan and R. Narasimhan (1961) Hecke’s functional equation and arithmetical identities. Ann. of Math. (2) 74, pp. 1–23. Cited by: §5.
  • [3] K. Chandrasekharan (1970) Arithmetical functions. Die Grundlehren der mathematischen Wissenschaften, Vol. 167, Springer. Cited by: §5.
  • [4] H. Cohen (2007) Number theory, volume II: analytic and modern tools. Graduate Texts in Mathematics, Vol. 240, Springer. Cited by: §1, §2.
  • [5] P. Flajolet, X. Gourdon, and P. Dumas (1995) Mellin transforms and asymptotics: harmonic sums. Theoret. Comput. Sci. 144 (1-2), pp. 3–58. Cited by: §5.
  • [6] D. A. Hejhal (1979) A note on the Voronoĭ summation formula. Monatsh. Math. 87 (1), pp. 1–14. Cited by: §5.
  • [7] A. Ivić (1998) The Voronoi identity via the Laplace transform. Ramanujan J. 2 (1-2), pp. 39–45. Cited by: §5.
  • [8] A. Ivić (2003) The Riemann zeta-function: theory and applications. Dover Publications. Cited by: §5.
  • [9] H. Iwaniec and E. Kowalski (2004) Analytic number theory. American Mathematical Society Colloquium Publications, Vol. 53, American Mathematical Society, Providence, RI. Cited by: §5.
  • [10] E. Landau (1918) Über die Wigertsche asymptotische Funktionalgleichung für die Lambertsche Reihe. Archiv der Mathematik und Physik, 3. Reihe 27, pp. 144–146. Note: Collected Works, volume 7, pp. 135–137 Cited by: §4.
  • [11] S. D. Miller and W. Schmid (2004) Summation formulas, from Poisson and Voronoi to the present. In Noncommutative Harmonic Analysis: In Honor of Jacques Carmona, P. Delorme and M. Vergne (Eds.), Progress in Mathematics, Vol. 220, pp. 419–440. Cited by: §5.
  • [12] E. M. Stein and R. Shakarchi (2011) Functional analysis. Princeton Lectures in Analysis, Vol. IV, Princeton University Press. Cited by: §5.
  • [13] E. C. Titchmarsh (1986) The theory of the Riemann zeta-function. second edition, Clarendon Press, Oxford. Cited by: §4, §4.
  • [14] E. C. Titchmarsh (2002) The theory of functions. second edition, Oxford University Press. Cited by: §4.
  • [15] G. Voronoï (1903) Sur un problème du calcul des fonctions asymptotiques. J. Reine Angew. Math. 126 (4), pp. 241–282. Cited by: §5.
  • [16] G. Voronoï (1904) Sur une fonction transcendante et ses applications à la sommation de quelques séries, seconde partie. Annales Scientifiques de l’École Normale Supérieure, troisième série 21, pp. 459–533. Cited by: §5.
  • [17] G. Voronoï (1904) Sur une fonction transcendante et ses applications à la sommation de quelques séries. Annales Scientifiques de l’École Normale Supérieure, troisième série 21, pp. 207–267. Cited by: §5.
  • [18] S. Wigert (1916) Sur la série de Lambert et son application à la théorie des nombres. Acta Math. 41, pp. 197–218. Cited by: §4.