Vous êtes sur la page 1sur 62

EFFECTIVE APPROXIMATION OF HEAT FLOW EVOLUTION OF THE

RIEMANN ξ FUNCTION, AND AN UPPER BOUND FOR THE DE


BRUIJN-NEWMAN CONSTANT

D.H.J. POLYMATH

Abstract. For each t ∈ R, define the entire function


Z ∞
2
Ht (z) B etu Φ(u) cos(zu) du
0
where Φ is the super-exponentially decaying function
X∞
Φ(u) B (2π2 n4 e9u − 3πn2 e5u ) exp(−πn2 e4u ).
n=1
This is essentially the heat flow evolution of the Riemann ξ function. From the work of de Bruijn
and Newman, there exists a finite constant Λ (the de Bruijn-Newman constant) such that the
zeroes of Ht are all real precisely when t ≥ Λ. The Riemann hypothesis is equivalent to the
assertion Λ ≤ 0; recently, Rodgers and Tao established the matching lower bound Λ ≥ 0. Ki,
Kim and Lee established the upper bound Λ < 12 .
In this paper we establish several effective estimates on Ht (x + iy) for t ≥ 0, including some
that are accurate for small or medium values of x. By combining these estimates with numerical
computations, we are able to obtain a new upper bound Λ ≤ 0.22 unconditionally, as well as
improvements conditional on further numerical verification of the Riemann hypothesis. We also
obtain some new estimates controlling the asymptotic behavior of zeroes of Ht (x + iy) as x → ∞.

1. Introduction
Let H0 : C → C denote the function
!
1 1 iz
(1) H0 (z) B ξ + ,
8 2 2
where ξ denotes the Riemann ξ function
s(s − 1) −s/2  s 
(2) ξ(s) B π Γ ζ(s)
2 2
(after removing all singularities) and ζ is the Riemann ζ function. Then H0 is an entire even
function with functional equation H0 (z) = H0 (z), and the Riemann hypothesis is equivalent to
the assertion that all the zeroes of H0 are real.
It is a classical fact (see [25, p. 255]) that H0 has the Fourier representation
Z ∞
H0 (z) = Φ(u) cos(zu) du
0
where Φ is the super-exponentially decaying function
X∞
(3) Φ(u) B (2π2 n4 e9u − 3πn2 e5u ) exp(−πn2 e4u ).
n=1
1
2 D.H.J. POLYMATH

The sum defining Φ(u) converges absolutely for negative u also. From Poisson summation one
can verify that Φ satisfies the functional equation Φ(u) = Φ(−u) (i.e., Φ is even).
De Bruijn [4] introduced (with somewhat different notation) the more general family of func-
tions Ht : C → C for t ∈ R, defined by the formula
Z ∞
2
(4) Ht (z) B etu Φ(u) cos(zu) du.
0

As noted in [8, p.114], one can view Ht as the evolution of H0 under the backwards heat equation
∂t Ht (z) = −∂zz Ht (z). As with H0 , each of the Ht are entire even functions with functional
2
equation Ht (z) = Ht (z); from the super-exponential decay of etu Φ(u) we see that the Ht are in
fact entire of order 1. It follows from the work of Pólya [18] that if Ht has purely real zeroes
for some t, then Ht0 has purely real zeroes for all t0 > t; de Bruijn showed that the zeroes of
Ht are purely real for t ≥ 1/2. Newman [13] strengthened this result by showing that there is
an absolute constant −∞ < Λ ≤ 1/2, now known as the De Bruijn-Newman constant, with the
property that Ht has purely real zeroes if and only if t ≥ Λ. The Riemann hypothesis is then
clearly equivalent to the upper bound Λ ≤ 0. Recently in [20] the complementary bound Λ ≥ 0
was established, answering a conjecture of Newman [13], and improving upon several previous
lower bounds for Λ [5, 14, 7, 6, 15, 21]. Furthermore, Ki, Kim, and Lee [9] sharpened the upper
bound Λ ≤ 1/2 of de Bruijn [4] slightly to Λ < 1/2.
In this paper we improve the upper bound:
Theorem 1.1 (New upper bound). We have Λ ≤ 0.22.
The proof of Theorem 1.1 combines numerical verification with some new asymptotics and
observations about the Ht which may be of independent interest. Firstly, by analyzing the dy-
namics of the zeroes of Ht , we establish in Section 3 the following criterion for obtaining upper
bounds on Λ:
Theorem 1.2 (Upper bound criterion). Suppose that t0 , X > 0 and 0 < y0 ≤ 1 obey the following
hypotheses:
(i) (Numerical verification of RH at initial time 0) There are no zeroes ζ(σ + iT ) = 0 with
1+y0
2 ≤ σ ≤ 1 and 0 ≤ T ≤ 2 .
X

(ii) (Asymptotic
q zero-free region at √final time t0 ) There are no zeroes Ht0 (x + iy) = 0 with
x ≥ X + 1 − y20 and y0 ≤ y ≤ 1 − 2t0 .
(iii) (Barrier
q at intermediate times) There are no zeroes Ht (x + iy) = 0 with X ≤ x ≤
q √
X + 1 − y0 , y0 + 2(t0 − t) ≤ y ≤ 1 − 2t, and 0 ≤ t ≤ t0 .
2 2

Then Λ ≤ t0 + 12 y20 .
Informally, hypothesis (i) implies that at time t = 0, there are no zeroes Ht (x + iy) = 0 with
large values of y to the left of the barrier region in (iii). The absence of zeroes in that barrier,
together with a continuity argument and an analysis of the time derivative of each zero, can then
be used to show that for later times 0 < t ≤ t0 , there continue to be no zeroes Ht (x + iy) = 0
with large values of y to the left of the barrier; see Figure 1. Hypothesis (ii) then gives the
complementary assertion to the right of the barrier, and one can use an existing theorem of de
Bruijn (Theorem 3.2) to conclude.
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 3

In practice, we have found it convenient numerically to replace the barrier region in Theorem
1.2 with the larger and simpler region
X ≤ x ≤ X + 1; y0 ≤ y ≤ 1; 0 ≤ t ≤ t0 .

Figure 1. A visualization of Theorem 1.2. If at time t0 one can show that there
are no zeroes Ht (x + iy) = 0 in the “canopy” 0 ≤ x < ∞, y0 ≤ y ≤ 1, then a
theorem of de Bruijn allows one to conclude the desired bound Λ ≤ t0 + 21 y20 .
The hypothesis (i) prevents zeroes hitting this canopy from an initial position
to the left of the barrier; the hypothesis (iii) prevents zeroes from lying in the
canopy to the right of the barrier; and the hypothesis (ii) prevents zeroes from
starting to the right of the barrier and reaching the canopy to the left of the
barrier.

We will obtain Theorem 1.1 by applying Theorem 1.2 with the specific numerical choices
t0 = 0.2, X = 6 × 1010 + 83952 − 0.5, and y0 = 0.2. The reason we choose X close to 6 × 1010
is that this is near the limit of known numerical verifications of the Riemann hypothesis such
as [17], which we need for the hypothesis (i) of the above theorem; the shift 83952 − 0.5 is
!−1
Q 1
in place to make the partial Euler product p≤11 1 − 1−iX large, which helps in keeping
p 2
the functions Ht (x + iy), Ht0 (x + iy) large in magnitude, which in turn is helpful for numerical
verifications of (ii) and (iii); see also Figure 11. The choices t0 = 0.2, y0 = 0.2 are then close to
the limit of our ability to numerically verify hypothesis (ii) for this choice of X. (The hypothesis
(iii) is also verified numerically, but can be done relatively quickly compared to (ii), and so does
not present the main bottleneck to further improvements to Theorem 1.1.) Further upper bounds
to Λ can be obtained if one assumes the Riemann hypothesis to hold up to larger heights than
that in [17]: see Section 11.
To verify (ii) and (iii), we need efficient approximations (of Riemann-Siegel type) for Ht (x +
iy) in the regime where t, y are bounded and x is large. For sake of numerically explicit constants,
we will focus attention on the region
1
(5) 0 < t ≤ ; 0 ≤ y ≤ 1; x ≥ 200,
2
4 D.H.J. POLYMATH

though the results here would also hold (with different explicit constants) if the numerical quan-
tities 21 , 1, 200 were replaced by other quantities.
A key difficulty here is that Ht (x + iy) decays exponentially fast in x (basically because of the
Gamma factor in (2)); see Figure 3. This means that any direct attempt to numerically establish
a zero-free region for Ht (x + iy) for large x would require enormous amounts of numerical
precision. To get around this, we will first renormalise the function Ht (x + iy) by dividing it
by a nowhere vanishing explicit function Bt (x + iy) (basically a variant of the aforementioned
Gamma factor) that removes this decay. To describe this function, we first introduce the function
M0 : C\(−∞, 1] → C\{0} defined by the formula
1 s(s − 1) −s/2 √
! !
s 1 s s
(6) M0 (s) B π 2π exp − Log − ,
8 2 2 2 2 2
where Log denotes the standard branch of the complex logarithm, with branch cut at the negative
axis and imaginary part  in (−π, π]. One may interpret M0 (s) as the Stirling approximation to the
1 s(s−1) −s/2
factor 8 2 π Γ 2s appearing in (1), (2); it decays exponentially as one moves to infinity
s → ±i∞ along the critical strip. We may form a holomorphic branch log M0 : C\(−∞, 1] → C
of the logarithm of M0 by the formula
√ !
s 2π s 1 s s
(7) log M0 (s) B Logs + Log(s − 1) − log π + log + − Log − ;
2 16 2 2 2 2
differentiating this, we see that the logarithmic derivative α : C\(−∞, 1] → C of this function,
defined by
M00
(8) α B (log M0 )0 =
M0
is given explicitly by the formula
1 1 1 1 s 1
α(s) = + − log π + log −
(9) s s−1 2 2 2 2s
1 1 1 s
= + + log .
2s s − 1 2 2π
For any time t ∈ R, we then define the deformation Mt : C\(−∞, 1] of M0 by the formula
t 
(10) Mt (s) B exp α(s)2 M0 (s)
4
for any t ≥ 0. In the region (5), we introduce the quantity
1 + y − ix
!
(11) Bt (x + iy) B Mt .
2
For fixed t ≥ 0 and y > 0, Bt (x + iy) is non-vanishing, and it is easy to verify the asymptotic
π
|Bt (x + iy)| = e−( 8 +o(1))x . As it turns out, Bt (x + iy) is an asymptotic approximation to Ht (x + iy)
in the region (5), in the sense that
Ht (x + iy)
(12) lim =1
x→∞ Bt (x + iy)
for any fixed t > 0 and y > 0; see Figure 2. (However, the convergence of (12) is not uniform as
t approaches zero.)
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 5

Figure 2. The quantity |Ht (x + iy)/Bt (x + iy) − 1| when y = t = 0.4 and log10 x ≤
12 (left) and when for 1000 ≤ x ≤ 1030 and various choices of (y, t) (right).

Figure 3. The quantities log |Ht (x + iy)| and log |Bt (x + iy)| when y = t = 0.4
and log10 x ≤ 7. Both quantities decay like − π8 x ≈ −0.393x.

In fact we have the following significantly more accurate approximation (of Riemann-Siegel
type) with effective error estimates. For any real number X, let O≤ (X) denote a quantity that is
bounded in magnitude by X. We also use x+ = max(x, 0) to denote the positive part of a real
number x.
6 D.H.J. POLYMATH

Theorem 1.3 (Effective Riemann-Siegel approximation to Ht (x + iy)). Let t, x, y lie in the region
(5). Then we have
Ht (x + iy)
= ft (x + iy) + O≤ eA + eB + eC,0

(13)
Bt (x + iy)
where
N N
X btn X
y bn
t
(14) ft (x + iy) B + γ n
n=1
n s∗ n=1
n s∗ +κ
t
(15) btn B exp( log2 n)
4
Mt 1−y+ix
 
2
(16) γ = γ(x + iy) B
Mt 1+y−ix
 
2
1 + y − ix t 1 + y − ix
!
(17) s∗ = s∗ (x + iy) B + α
2 2 2
1 − y + ix 1 + y + ix
! !!
t
(18) κ = κ(x + iy) B α −α
2 2 2
 r 
 x t 
(19) N B  + 
4π 16
and eA , eB , eC,0 are certain explicitly computable positive quantities1 depending on t and x + iy.
Furthermore, we have the following bounds:
 x −y/2
(20) |γ| ≤ e0.02y

1+y t 4y(1 + y)
!
x t
(21) Res∗ ≥ + log − 2 1 − 3y +
2 4 4π 2x x2 +
ty
(22) |κ| ≤
2(x − 6)
 t2
 16 log2 x 2 + 0.626 
N
  
X bt  
4πn
(23) eA + eB ≤ (1 + |γ|N |κ| ny ) Re(s
n exp   − 1
n=1
n ∗)   x − 6.66  
π
1.24 × (3y + 3−y ) 3| log 4π + i 2 | + 10.44
 x − 1+y x !
4 t 2 x
(24) eC,0 ≤ exp − log + +
4π 16 4π N − 0.125 x − 8.52
This theorem will be proven in Section 6; see Figures 4, 5 for a numerical illustration of the
approximation. The strategy is to express Ht as a convolution of H0 with a gaussian heat kernel,
then apply an effective Riemann-Siegel expansion to H0 to rewrite Ht as the sum of various
contour integrals; see Section 4 for details. One then uses the saddle point method to shift each
such contour to a location that is suitable for effective estimation.
From (13) and the triangle inequality, we have a numerically verifiable criterion to establish
non-vanishing of Ht at a given point:

1See (71)-(74) for the precise definition of these quantities.


UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 7

Figure 4. Comparison of | ft (x + iy)| and |Ht (x + iy)|/|Bt (x + iy)| for y = 0.9,


t = 0.1, 1000 ≤ x ≤ 1030, (left) and when 106 ≤ x ≤ 106 + 30 (right). The
approximation improves as x gets larger.

Figure 5. Comparison of | ft (x+iy)| and |Ht (x+iy)|/|Bt (x+iy)| for y = 0, t = 0.3,


1000 ≤ x ≤ 1030 (left), and when 106 ≤ x ≤ 106 + 30 (right). Again notice the
improving approximation with x.

Corollary 1.4 (Criterion for non-vanishing). Let t, x, y lie in the region (5), and let ft , eA , eB , eC,0
be as in Theorem 1.3. If one has the inequality

(25) | ft (x + iy)| > eA + eB + eC,0

then Ht (x + iy) , 0.
8 D.H.J. POLYMATH

Figure 6. The error upper bound |eA + eB + eC,0 | versus | ft − Ht /Bt | when y =
t = 0.2 and 2 ≤ log10 x ≤ 5.

Figure 7. The individual error upper bounds |eA |, |eB | and |eC,0 | with y = t = 0.2
and 2 ≤ log10 x ≤ 5. The eC,0 -term clearly dominates.

In the asymptotic limit x → ∞, one easily sees that


log2 x
!
eA + e B = O
x
 3+y  t 
eC,0 = O x− 4 exp − log2 x
 t 16
ft (x + iy) = 1 + O x − 4 log 2
,
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 9

thus giving the crude asymptotic (12) in the region (5) at least. In practice, the eC,0 term nu-
merically dominates the eA + eB term, although both errors will be quite small in the ranges of
x under consideration; in particular, for the ranges needed to verify conditions (ii) and (iii) of
Theorem 1.2, we can make eA + eB and eC,0 both significantly smaller than | ft (x + iy)|. In the
spirit of expanding the Riemann-Siegel approximation to higher order, we also obtain an even
more accurate explicit approximation in which a correction term − CBtt is added to ft , and the error
term eC,0 is replaced by a smaller quantity eC ; see (69) and Figure 8.

(x+iy)
Figure 8. Comparison of | ft (x+iy)|, | ft (x+iy)− CBtt (x+iy) | and |Ht (x+iy)|/|Bt (x+iy)|
for y = 0, t = 0.3, 1000 ≤ x ≤ 1012.

In addition to establishing upper bounds such as Theorem 1.1, one can use Theorem 1.3
and Corollary 1.4 (together with variants in slightly larger regions than (5), for instance if y is
allowed to be as large as 10) to obtain asymptotic control on the zeroes of Ht , refining previous
work of Ki, Kim, and Lee [9]. Indeed, in Section 10 we will establish
Theorem 1.5 (Distribution of zeroes of Ht ). Let 0 < t ≤ 1/2, let C > 0 be a sufficiently large
absolute constant, and let c > 0 be a sufficiently small absolute constant. For x ≥ 4π, define
x x x 11 t x
g(x, t) B log − + + log
4π 4π 4π 8 16 4π
and all n ≥ C, let xn be the unique real number greater than 4π such that
(26) g(xn , t) = n.
(This is well-defined since the g(x, t) is increasing in x for x ≥ 4π.)
(i) If x ≥ exp( Ct ) and Ht (x + iy) = 0, then y = 0, and
x = xn + O(x−ct )
for some n.
(ii) Conversely, for each n ≥ exp( Ct ) there is exactly one zero Ht in the disk {x + iy : |x + iy −
xn | ≤ logc xn } (and by part (i), this zero will be real and lie within O(x−ct ) of xn ).
10 D.H.J. POLYMATH

(iii) If X ≥ exp( Ct ), the number Nt (X) of zeroes with real part between 0 and X (counting
multiplicity) is
Nt (X) = g(X, t) + O(1).
(iv) For any X ≥ 0, one has
Nt (X + 1) − Nt (X) ≤ O(log(2 + X))
and
Nt (X) = g(X, t) + O(log(2 + X)).
Here and in the sequel we use X = O(Y) to denote the estimate |X| ≤ AY for some constant A
that is absolute (in particular, A is independent of t and C).
Roughly speaking, these estimates tell us that the zeroes of Ht behaves (on macroscopic
scales) like those of H0 in the region x = O(exp(O(1/t))), and are very evenly spaced (and on
xn
the real axis) outside of this range. The factor 16t log 4π in (26) indicates that as time t advances,
the zeroes (or at least those with large values of x) will tend to move towards the origin at a
speed of approximately π4 . Although we will not prove this here, the conclusions (i) and (iii)
suggest that one in fact has an asymptotic of the form
j k
Nt (X) = g(X, t) + O(X −ct )
when X ≥ exp(C/t); in particular (since the sawtooth function x − bxc has average value 21 ) one
would have the heuristic approximation
X X X 7 t X
Nt (X) ≈ log − + + log
4π 4π 4π 8 16 4π
after performing some averaging in X, thus recovering the familiar 87 term in the usual averaged
asymptotics for N0 (X).

Figure 9. The real zeroes xi of Ht will converge to integer values of g(xi , t)


when t (left) and/or x (right) increases.

The results in Theorem 1.5 refine previous results of Ki, Kim, and Lee [9, Theorems 1.3,
1.4], which gave similar results but with constants that depended on t in a non-uniform (and
ineffective) fashion, and error terms that were of shape o(1) rather than O(x−ct ) in the limit
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 11

x → ∞ (holding t fixed). The results may also be compared with those in [2], who (in our
notation) show that assuming RH, the zeroes of H0 are precisely the solutions xn to the equation
 2iϑ(x /2) ζ 0 ( 1−ix
 n 

1 )
2
=n

arg −e n
2π ζ( 2 )
0 1+ixn 

for integer n, where −ϑ(t) is the phase of ζ( 12 + it) and one chooses a branch of the argument so
that the left-hand side is − 12 when xn = 0.
Remark 1.6. One can draw an analogy between the various potential behaviours of zeroes of Ht
and the three classical states of matter. A “gaseous” state corresponds to the situation in which
the zeroes of Ht are strictly complex. A “liquid” state corresponds to a situation in which the
zeroes are real, but disordered (with highly unequal spacings between zeroes). A “solid” state
corresponds to a situation in which the zeroes are real and arranged roughly in an arithmetic
progression. Thus for instance the Riemann hypothesis and the GUE hypothesis assert (roughly
speaking) that the zeroes of H0 should exhibit liquid behaviour everywhere, while Theorem 1.5
asserts that the zeroes of Ht , t > 0 “solidify” in the region x ≥ exp(C/t). Below this region we
expect liquid behaviour. In general, as the parameter t increases, the zeroes appear to “cool”
down, transitioning from gaseous to liquid to solid type states; see [20] for some formalisations
of this intuition.
1.1. About this project. This paper is part of the Polymath project, which was launched by
Timothy Gowers in February 2009 as an experiment to see if research mathematics could be
conducted by a massive online collaboration. The current project (which was administered by
Terence Tao) is the fifteenth project in this series. Further information on the Polymath project
can be found on the web site michaelnielsen.org/polymath1. Information about this spe-
cific project may be found at
michaelnielsen.org/polymath1/index.php?title=De Bruijn-Newman constant
and a full list of participants and their grant acknowledgments may be found at
michaelnielsen.org/polymath1/index.php?title=Polymath15 grant acknowledgments

2. Notation
We use the standard branch Log of the logarithm to define the standard complex powers

zw B exp(wLogz), and in particular define the standard square root z B z1/2 = exp( 12 Logz).
We record the familiar gaussian identity
π
r
b2
Z   !
(27) exp −(au + bu + c) du =
2
exp −c
R a 4a
for any complex numbers a, b, c with Rea > 0.
When using order of magnitude notation such as O≤ (X), any expression of the form A = B
using this notation should be interpreted as the assertion that any quantity of the form A is also
of the form B, thus for instance O≤ (1) + O≤ (1) = O≤ (3). (In particular, the equality relation is
no longer symmetric with this notation.)
If F is a meromorphic function, we use F 0 to denote its derivative. We also use F ∗ to denote
the reflection F ∗ (s) := F(s) of F. Observe from analytic continuation that if F : Ω → C is
holomorphic on a connected open domain Ω ⊂ C containing an interval in R, and is real-valued
12 D.H.J. POLYMATH

on Ω ∩ R, then it is equal to its own reflection: F = F ∗ (since the holomorphic function F − F ∗


has an uncountable number of zeroes).

3. Dynamics of zeroes
In this section we control the dynamics of the zeroes of Ht in order to establish Theorem 1.2.
As Ht is even with functional equation Ht = Ht∗ , the zeroes are symmetric around the origin and
the real axis; from (4) and the positivity of Φ, we also see that Ht (iy) > 0 for all y ∈ R, so there
are no zeroes on the imaginary axis. From the super-exponential decay of Φ and (4) we see that
the entire function Ht is of order 1; by Jensen’s formula, this implies that the number of zeroes
in a large disk D(0, R) is at most O(R1+o(1) ) as R → ∞.
We begin with the analysis of the dynamics of a single zero of Ht :
Proposition 3.1 (Dynamics of a single zero). Let t0 ∈ R, and let (zk (t0 ))k∈Z\{0} be an enumeration
of the zeroes of Ht0 in C (counting multiplicity), with the symmetry condition z−k (t0 ) = −zk (t0 ).
(i) If j ∈ Z\{0} is such that z j (t0 ) is a simple zero of Ht0 , then there exists a neighbourhood
U of z j (t0 ), a neighbourhood I of t0 in R, and a smooth map z j : I → U such that for
every t ∈ I, z j (t) is the unique zero of Ht in U. Furthermore one has the equation
0
∂z j X 1
(28) (t0 ) = 2
∂t z (t
k, j j 0
) − zk (t0 )

where the sum is over those k ∈ Z\{0} with k , j, and the prime means that the k and −k
terms are summed together (except for the k = − j term, which is summed separately) in
order to make the sum convergent.
(ii) If j ∈ Z\{0} is such that z j (t0 ) is a repeated zero of Ht0 of order m ≥ 2, then there is a
neighbourhood U of z j (t0 ) such that for t sufficiently close to t0 , there are precisely m
zeroes of Ht in U, and they take the form

z j (t0 ) + 2(t − t0 )1/2 x j + O(|t − t0 |)
for j = 1, . . . , m as t → t0 , where x1 < · · · < xm are the roots of the mth Hermite
polynomial
z2 d m z2
! !
(29) Hem (z) B (−1)m exp exp −
2 dzm 2
X m! z m−2l
(30) = (−1)l l
0≤l≤m/2
l!(m − 2l)! 2
and the implied constant in the O() notation can depend on t0 , j, and m.
The differential equation (28) was previously derived in [8, Lemma 2.4] in the case t > Λ (in
which all zeroes are real and simple); however, in our applications we also need to consider the
regime t < Λ in which the zeroes are permitted to be complex or repeated. The roots x1 , . . . , xm
can be given explicitly for small values of m as
x1 = −1; x2 = +1
when m = 2, √ √
x1 = − 3; x2 = 0; x3 = + 3
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 13

when m = 3, and
√ √ √ √
q q q q
x1 = − 3 + 6; x2 = − 3 − 6; x3 = 3 − 6; x4 = 3+ 6
when m = 4. From (29) and iterating Rolle’s
 2 theorem we see that all the roots x1 , . . . , xm of
2 − z dz + m Hem (z) = 0 and the Picard uniqueness
d d
Hem are real; from the Hermite equation dz
theorem for ODE we see that the zeroes are all simple.

Proof. First suppose we are in the situation of (i). As z j (t0 ) is simple, ∂z Ht is non-zero at z j (t0 );
since Ht (z) is a smooth function of both t and z, we conclude from the implicit function theorem
that there is a unique solution z j (t) ∈ U to the equation
Ht (z j (t)) = 0
with z j (t) in a sufficiently small neighbourhood U of z j (t0 ), if t is in a sufficiently small neigh-
bourhood I of t0 ; furthermore, z j (t) depends smoothly on t, and agrees with z j (t0 ) when t = t0 .
Differentiating the above equation at t0 , we obtain
∂Ht ∂z j
|t=t0 (z j (t0 )) + (t0 )Ht00 (z j (t0 )) = 0,
∂t ∂t
where the primes denote differentiation in the z variable. On the other hand, from (4) and differ-
entiation under the integral sign (which can be justified using the rapid decrease of Φ) we have
the backwards heat equation
∂Ht
(31) = −Ht00
∂t
for all t ≥ 0. Inserting this into the previous equation, we conclude that
∂z j H 00
(32) (t0 ) = t0 (z j (t0 )),
∂t Ht
noting that the denominator Ht0 (z j (t0 )) vanishes by the hypothesis that the zero at z j (t0 ) is simple.
Henceforth we omit the dependence on t0 for brevity. From Taylor expansion of Ht , Ht0 , and Ht00
around the simple zero z j we see that
Ht00 Ht0
!
1
(33) (z j ) = 2 lim (z) − .
Ht0 z→z j Ht z − zj
On the other hand, as Ht is even, non-zero at the origin, and entire of order 1, we see from the
Hadamard factorization theorem that
0 !
Y z
Ht (z) = Ht (0) 1− ,
k
zk
where the prime indicates that the k and −k factors are multiplied together. The product is locally
uniformly convergent, so we may take logarithmic derivatives and conclude that
0
Ht0 X 1
(z) = .
Ht k
z − zk
Inserting this into (32), (33) and using the dominated convergence theorem (again using the
growth in the number of zeroes to justify the interchange of summation and limits), we obtain
the claim (i).
14 D.H.J. POLYMATH

Now we prove (ii). We abbreviate z j (t0 ) as z j . By Taylor expansion we have


∂2k Ht0
(z) = m(m − 1) . . . (m − 2k + 1)am (z − z j )m−2k + O(|z − z j |max(m−2k+1,0) )
∂z2k
as z → z j for any fixed integer k ≥ 0 and some non-zero complex number am = am (z j , t0 ) (with
the implied constant in the O() notation allowed to depend on k, z j , t0 ); applying the backwards
heat equation (31) we thus have
∂k Ht
|t=t0 (z) = (−1)k m(m − 1) . . . (m − 2k + 1)am (z − z j )m−2k + O(|z − z j |max(m−2k+1,0) ).
∂tk
Performing Taylor expansion in time and using (30), we conclude that in the regime z − z j =
O(|t − t0 |1/2 ), one has the bound
√ z − zj
!  !
Ht (z) = am ((t − t0 ) ) Hem 2
1/2 m
+ O |t − t0 | 1/2
(t − t0 )1/2
as t → t0 , using (say) the standard branch of the square root. By the inverse function theorem
(and the simple nature of the zeroes of Hem ), we conclude that for t sufficiently close but not
equal to to t0 , we have m zeroes of Ht of the form
z j + (t − t0 )1/2 x j + O(|t − t0 |).
By Rouche’s theorem, if U is a sufficiently small neighborhood of z j then these are the only
zeroes of Ht in U for t sufficiently close to t0 . The claim follows. 
Next, we recall the following bound of de Bruijn:
Theorem 3.2. Suppose that t0 ∈ R and y0 > 0 is such that there are no zeroes Ht0 (x + iy) = 0
with x ∈ R and y > y0 . Then for any t > t0 , there are no zeroes Ht (x + iy) = 0 with x ∈ R and
y > max(y20 − 2(t − t0 ), 0)1/2 . In particular one has Λ ≤ t0 + 21 y20 .
Proof. See [4, Theorem 13]. 
We are now ready to prove Theorem 1.2. The main step is to establish
Proposition 3.3 (Zero-free region criterion). Suppose that t0 , X > 0 and 0 < y0 ≤ 1 obey the
following hypotheses:
q
(i) There are no zeroes H0 (x + iy) = 0 with 0 ≤ x ≤ X and y20 + 2t0 ≤ y ≤ 1.
q √
(ii) There are no zeroes Ht0 (x + iy) = 0 with x ≥ X + 1 − y20 and y0 ≤ y ≤ 1 − 2t0 .
q q
(iii) There are no zeroes Ht (x + iy) = 0 with X ≤ x ≤ X + 1 − y20 , y20 + 2(t0 − t) ≤ y ≤

1 − 2t, and 0 ≤ t ≤ t0 .
Then there are no zeroes Ht0 (x + iy) = 0 with x ∈ R and y ≥ y0 .
Proof. It is well known that the Riemann ξ function has no zeroes outside of the strip {0 ≤
Re(s) ≤ 1}, hence there are no zeroes H0 (x + iy)√= 0 with y > 1. By √ Theorem 3.2, we may thus
remove the upper bound constraints y ≤ 1, y ≤ 1 − 2t0 , and y ≤ 1 − 2t from (i), (ii), and (iii)
respectively.
By hypotheses (ii), (iii) and the symmetry properties of Ht , it suffices to show that for every
0 ≤ t ≤ t0 , there are no zeroes Ht (x + iy) = 0 with 0 ≤ x ≤ X and y ≥ Y(t), where Y(t) B
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 15
q
y20 + 2(t0 − t). By hypothesis (i), this is true at time t = 0. Suppose the claim failed for some
time 0 < t ≤ t0 . Let t1 ∈ (0, t0 ] be the minimal time in which this occurred (such a time exists
because Ht varies continuously in t, and there are no zeroes Ht (x + iy) = 0 with (say) y > 1).
From Rouche’s theorem (or Proposition 3.1) we conclude that there is a zero Ht1 (x + iy) = 0
with x + iy on the boundary of the region {x + iy : 0 ≤ x ≤ X, y ≥ Y(t1 )}. The right side x = X
of this boundary is ruled out by hypothesis (ii), and (as mentioned at the start of the section) the
left side x = 0 is ruled out by (4) and the positivity of Φ. Thus by the symmetry properties of
Ht1 we must have
Ht1 (x + iY(t1 )) = 0
for some 0 < x < X.
Suppose first that Ht1 has a repeated zero at x + iy0 . Using Proposition 3.1(ii) and observing
(from the symmetry of Hem ) that at least one of the roots x1 , . . . , xm is positive, we then see
that for t < t1 sufficiently close to t1 , Ht has a zero in the region {x + iy : 0 ≤ x ≤ X, y ≥ Y(t)},
contradicting the minimality of t1 . Thus the zero x+iY(t1 ) of Ht1 must be simple. In particular, by
Proposition 3.1(i) we can write x+iY(t1 ) = z j (t1 ) for some smooth function z j in a neighbourhood
of t1 obeying (28), such that z j (t) is a zero of Ht for all t close to t1 . We will prove that
∂ ∂
(34) Im z j (t1 ) < Y(t1 ),
∂t ∂t
which implies that there is a zero of Ht in the region {x + iy : 0 ≤ x ≤ X, y ≥ Y(t)} for t < t1
sufficiently close to t1 , giving the required contradiction.
The right-hand side of (34) is
∂ 1
Y(t1 ) = − .
∂t Y(t1 )
By Proposition 3.1(i), the left-hand side of (34) is
0
X Y(t1 ) − yk
2
k, j
(x − xk )2 + (Y(t1 ) − yk )2

where we write zk = xk + iyk . Clearly any zero xk + iyk with imaginary part yk in [−Y(t1 ), Y(t1 )]
gives a non-positive contribution to this sum, the contribution of the zero x − iY(t1 ) is − Y(t11 ) ,
the contribution of the zero −x + iY(t1 ) vanishes, and the contribution of −x − iY(t1 ) is negative.
Grouping the remaining zeroes with their complex conjugates, it then suffices to show that
Y(t1 ) − yk Y(t1 ) + yk
+ ≤0
(x − xk ) + (y j − yk )
2 2 (x − xk )2 + (Y(t1 ) + yk )2
whenever yk > Y(t1 ). Cross-multiplying and canceling like terms, this inequality eventually
simplifies to
y2k ≤ (x − xk )2 + Y(t1 )2 .
p
But from the hypothesis (iii) and the assumption yk > Y(t1 ), we have |xk | ≥ X + 1 − Y(t1 )2 , so
(x − xk )2 ≥ 1 − Y(t1 )2 . On the other hand from Theorem 3.2 one has yk < 1, giving the required
contradiction. 
By combining Proposition 3.3 with Theorem 3.2, we obtain Theorem 1.2, noting from (1),
(2) that condition (i) of Proposition 3.3 is equivalent to condition (i) of Theorem 3.2.
16 D.H.J. POLYMATH

4. Applying the fundamental solution for the heat equation


As discussed in the introduction, we will establish Theorem 1.3 by writing Ht in terms of H0
using the fundamental solution to the heat equation. Namely, for any t > 0, we have from (27)
that

Z
tu2 1 2
e = e±2 tvu √ e−v dv
R π
for any complex u and either choice of sign ±. Multiplying by e±izu and averaging, we conclude
that
√   1
Z
tu2 2

e cos(zu) = cos z − 2i tv u √ e−v dv
R π
for any complex z, u. Multiplying by Φ(u) and using Fubini’s theorem, we conclude the heat
kernel representation

Z
1 2
Ht (z) = H0 (z − 2i tv) √ e−v dv
R π
for any complex z. Using (1), we thus have
1 1 + iz √
Z !
1 2
(35) Ht (z) = ξ + tv √ e−v dv.
R 8 2 π
Remark 4.1. We have found numerically that the formula (35) gives a fast and accurate means
to compute Ht (z) when z is of moderate size, e.g., if z = x+iy with |x| ≤ 106 and |y| ≤ 1. However,
we will not need to directly compute the right-hand side of (35) for our application to bounding
Λ, as we will only need to control Ht (x + iy) for large values of x, and we will shortly develop
tractable approximations of Riemann-Siegel type that are more suitable for this regime.
We now combine this formula with expansions of the Riemann ξ-function. From [25, (2.10.6)]
we have the Riemann-Siegel formula
1
(36) ξ(s) = R0,0 (s) + R∗0,0 (1 − s)
8
for any complex s that is not an integer (in order to avoid the poles of the Gamma function),
where R0,0 (s) is the contour integral
2
w−s eiπw
Z
1 s(s − 1) −s/2  s 
R0,0 (s) := π Γ dw
8 2 2 0.1 eπiw − e−πiw
with 0 . 1 any infinite line oriented in the direction e5πi/4 that crosses the interval [0, 1]. From
the residue theorem (and the gaussian decrease of eiπw along the eπi/4 and e5πi/4 directions) we
2

may expand
N
X
R0,0 (s) = r0,n (s) + R0,N (s)
n=1
for any non-negative integer N, where r0,n , R0,N are the meromorphic functions
1 s(s − 1) −s/2  s  −s
(37) r0,n (s) := π Γ n ,
8 2 2
2
w−s eiπw
Z
1 s(s − 1) −s/2  s 
(38) R0,N (s) := π Γ dw
8 2 2 N.N+1 eπiw − e−πiw
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 17

and N . N + 1 denotes any infinite line oriented in the direction e5πi/4 that crosses the interval
[N, N + 1]. For any z √
that is not purely imaginary,
√ we see from Stirling’s approximation that the
functions r0,n ( 2 + tv) and R0,N ( 2 + tv) grow slower than gaussian as v → ±∞ (indeed
1+iz 1+iz

they grow like exp(O(|v| log |v|)), where the implied constants depend on t, z). From this and
(35), (36) we conclude that
N N
1 + iz 1 + iz
! X ! ! !
X 1 − iz 1 − iz
(39) Ht (z) = rt,n + ∗
rt,n + Rt,N + Rt,N

n=1
2 n=1
2 2 2
for any t > 0, any z that is not purely imaginary, and any non-negative integer N, where
rt,n (s), Rt,N (s) are defined for non-real s by the formulae
√  1
Z  2
rt,n (s) := r0,n s + tv √ e−v dv
R π
√  1 −v2
Z 
Rt,N (s) := R0,N s + tv √ e dv;
R π
these can be thought of as the evolutions of r0,n , R0,N respectively under the forward heat equa-
tion.
The functions r0,n (s), R0,N (s) grow slower than gaussian as long as the imaginary part of s is
bounded√ and bounded away from zero. As a consequence, we may shift contours (replacing v
by v + 2t αn ) and write
 t Z  √   √ t  1 2
(40) rt,n (s) = exp − α2n exp − tvαn r0,n s + tv + αn √ e−v dv
4 R 2 π
for any complex number αn with Im(s), Im(s + 2t αn ) having the same sign. Similarly we may
write
 t Z  √   √ t  1 2
(41) Rt,N (s) = exp − β2N exp − tvβN R0,N s + tv + βN √ e−v dv
4 R 2 π
for any complex number βN with Im(s), Im(s + 2t βN ) having the same sign. In the spirit of the
saddle point method, we will select the parameters αn , βN later in the paper in order to make the
integrands in (40), (41) close to stationary in phase at v = 0, in order to obtain good estimates
and approximations for these terms.

5. Elementary estimates
In order to explicitly estimate various error terms arising in the proof of Theorem 1.3, we will
need the following elementary estimates:
Lemma 5.1 (Elementary estimates). Let x > 0.
(i) If a > 0 and b ≥ 0 are such that x > b/a, then
a ! !
b a
O≤ + O≤ 2 = O≤ .
x x x − b/a

More generally, if a > 0 and b, c ≥ 0 are such that x > b/a, c/a, then
a ! !
b c a
O≤ + O≤ 2 + O≤ 3 = O≤ √ .
x x x x − max(b/a, c/a)
18 D.H.J. POLYMATH

(ii) If x > 1, then !! !


1 1
log 1 + O≤ = O≤ .
x x−1
or equivalently ! !!
1 1
1 + O≤ = exp O≤ .
x x−1
(iii) If x > 1/2, then !! !
1 1
exp O≤ = 1 + O≤ .
x x − 0.5
(iv) We have
exp (O≤ (x)) = 1 + O≤ (e x − 1).
(v) If z is a complex number with |Im(z)| ≥ 1 or Rez ≥ 1, then

! !!
1 1
Γ(z) = 2π exp z − log z − z + O≤ .
2 12(|z| − 0.33)
(vi) If a, b > 0, y ≥ 0 and x ≥ x0 ≥ exp(a/b) and x0 > c ≥ 0, then
loga |x + iy| loga |x0 + iy|
≤ .
(x − c)b (x0 − c)b
Proof. Claim (i) follows from the geometric series formula
a a at at2
= + 2 + 3 + ...
x−t x x x
whenever 0 ≤ t < x.
For Claim (ii), we use the Taylor expansion of the logarithm to note that
!! !
1 1 1 1
log 1 + O≤ = O≤ + 2 + 3 + ...
x x 2x 3x
which on comparison with the geometric series formula
1 1 1 1
= + 2 + 3 + ...
x−1 x x x
gives the claim. Similarly for Claim (iii), we may compare the Taylor expansion
!! !
1 1 1 1
exp O≤ = 1 + O≤ + + + ...
x x 2!x2 3!x3
with the geometric series formula
1 1 1 1
= + 2 + 2 3 + ...
x − 0.5 x 2x 2 x
k
and note that k! ≥ 2 for all k ≥ 2.
Claim (iv) follows from the trivial identity e x = 1 + (e x − 1) and the elementary inequality
e ≥ 1 − (e x − 1). For Claim (v), we may use the functional equation Γ = Γ∗ to assume that
−x

Im(z) ≥ 0. From the work of Boyd [3, (1.13), (3.1), (3.14), (3.15)] we have the effective Stirling
approximation

! ! !
1 1
Γ(z) = 2π exp z − log z − z 1 + + R2 (z)
2 12z
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 19

where the remainder R2 (z) obeys the bound


√ C2 Γ(2)
|R2 (z)| ≤ (2 2 + 1)
(2π)3 |z|2
for Re(z) ≥ 0 and
√ C2 Γ(2)
|R2 (z)| ≤ (2 2 + 1)
(2π)3 |z|2 |1 − e2πiz |
for Re(z) ≤ 0, where C2 is the constant
π2
!
1 1
C2 := (1 + ζ(2)) = 1+ .
2 2 6

In the latter case, we have Im(z) ≥ 1 by hypothesis, and hence |1 − e2πiz | ≥ 1 − e−2π . We conclude
that in all ranges of z of interest, we have
√ C2 Γ(2) 0.0205
|R2 (z)| ≤ (2 2 + 1) ≤
(2π) |z|2 (1 − e−2π )
3 |z|2
and hence by Claim (i)

! ! !!
1 1
Γ(z) = 2π exp z − log z − z 1 + O≤
2 12(|z| − 0.246)
and the claim then follows by Claim (ii).
loga |x+iy|
For Claim (vi), it suffices to show that the function x 7→ (x−c)b
is non-increasing for x ≥
2
log(1+ y2 )
exp(a/b). Since log |x + iy| = (log x)(1 + x
2 log x ) and the second factor is monotone decreasing
a
log x
in x, it suffices to show that x 7→ (x−c)b
is non-increasing in this region. Taking logarithms
a b b b
and differentiating, we wish to show that x log x − x−c ≤ 0. But this is clear since x−c ≥ x and
log x ≥ a/b. 

6. Estimates for large x


We can now begin the proof of Theorem 1.3. The strategy is to use the expansion (39), which
turns out to be an effective approximation in the region (5), since we will be able to ensure that
√ √
quantities such as s + tv + 2t αn or s + tv + 2t βN , with s = 1+i(x+iy)
2 , stay away from the real
axis where the poles of Γ are located (and also where the error terms in the Riemann-Siegel
approximation deteriorate).
Accordingly, we will need effective estimates on the functions rt,n , Rt,N appearing in Section
4. We will treat these two functions separately.

6.1. Estimation of rt,n . We recall the function α(s) defined in (8). From differentiating (9) we
see that
1 1 1
(42) α0 (s) = − 2
− 2
+
2s (s − 1) 2s
20 D.H.J. POLYMATH

whenever s ∈ C\(−∞, 1]. If Im(s) > 3, we conclude in particular the useful bound
! ! !
1 1 1
α (s) = O≤
0
+ O≤ + O≤
2Im (s)2 Im (s)2 2Im (s)
(43) !
1
= O≤
2Im (s) − 6
thanks to Lemma 5.1(i).
We also recall the function Mt and the coefficients btn from (10), (15) respectively. It turns out
we have a good approximation
btn
rt,n (σ + iT ) ≈ Mt (σ + iT ) .
nσ+iT + 2 α(σ+iT )
t

More precisely, we have


Proposition 6.1 (Estimate for rt,n ). Let σ be real, let T > 10, let n be a positive integer, and let
0 < t ≤ 1/2. Then
btn
rt,n (σ + iT ) = Mt (σ + iT ) 1 + O≤ (εt,n (σ + iT ))

nσ+iT + 2 α(σ+iT )
t

where
 t2
 |α(σ + iT ) − log n|2 + + 16 
t

(44) εt,n (σ + iT ) B exp  8 4
 − 1.
T − 3.33
Proof. From (37), (6) and Lemma 5.1(v) one has
!!
1
r0,n (s) = M0 (s)n−s exp O≤
6(|s| − 0.66)
whenever Im(s) > 2. Let αn denote the quantity
(45) αn B α(σ + iT ) − log n;
this is the logarithmic derivative of M(s)n−s at s = σ + iT . By (9) and the hypothesis T ≥ 10,
the imaginary part of αn may be lower bounded by
1 1
(46) Im(αn ) ≥ − − ≥ −0.15;
2T T
in particular, σ + iT and σ + iT + 2t αn have imaginary parts of the same sign. We can now apply
(40) to obtain
 t Z  √   √ t 
rt,n (σ + iT ) = exp − α2n exp − tvαn M0 σ + iT + tv + αn ×
4 2
  R  
√ t  1  1 −v2
× exp − σ + iT + tv + αn log n + O≤ 
 
√  √ e dv.
2 6(|σ + iT + tv + 2t αn | − 0.66) π

From (46) we see that σ + iT + tv + 2t αn has imaginary part at least T − 0.08. Thus
  ! !
1 1 1
 = O≤ = O≤ .
 
O≤   √ 
6(|σ + iT + tv + 2t αn | − 0.66) 6(T − 0.74) 6(T − 3.08)
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 21

From (43) we have


!
1
α (s) = O≤ 0
2(T − 3.08)

for all s on the line segment between σ + iT and σ + iT + tv + 2t αn . Applying Taylor’s theorem
with remainder to the branch of the logarithm log M0 defined in (7), we conclude that
 √
 | tv + 2t αn |2 
 
√ t √ t
M0 (σ + iT + tv + αn ) = M0 (σ + iT ) exp α(σ + iT )( tv + αn ) + O≤   .

 
2 2 4(T − 3.08) 

Combining these estimates, writing α(σ + iT ) = αn + log n, estimating | tv + 2t αn |2 by 2tv2 +
t2 2
2 |αn | , and simplifying, we conclude that
t 
rt,n (s) = M0 (σ + iT ) exp α2n − (σ + iT ) log n ×
4
  v2 + t2 |αn |2 + 1  1
Z  t 
2
× exp O≤  2 8 6 
 √ e−v dv.
R T − 3.08 π
Using (45), (10), (15) we see that
t  btn
M0 (σ + iT ) exp α2n − (σ + iT ) log n = Mt (σ + iT )
nσ+iT + 2 α(σ+iT )
t
4
and so it suffices to show that
2  t2
  2 v2 + t8 |αn |2 +  |αn |2 + t +
Z  t 1 
  
1

6 1 −v2  
exp O≤   √ e dv = 1 + O exp  8
 4 6
 − 1 .

R T − 3.08 π T − 3.33
2
Since √1 e−v dv integrates to one, it suffices by Lemma 5.1(iv) to show that
π

 tv2 + t2 |αn |2 +


 t2
 |αn |2 + t +
Z  
1

1
8 6 1 −v2
exp   √ e dv ≤ exp  8
 4 6
 .

R 2(T − 3.08) π T − 3.33
t2
Since 1
T −3.08 < T −3.33 , we can remove the 8 |αn | + 6
1 2 1 terms from both sides and reduce to showing
that
tv2
Z ! !
1 −v2 t
(47) exp √ e dv ≤ exp .
R 2(T − 3.08) π 4(T − 3.33)
Using (27), the left-hand side may be calculated exactly as
!−1/2
t
1− .
2(T − 3.08)
Applying Lemma 5.1(ii) and using the hypotheses t ≤ 1/2, T ≥ 10, one has
!!
t t
1− = exp O≤
2(T − 3.08) 2(T − 3.33)
and the claim follows. 
22 D.H.J. POLYMATH

6.2. Estimation of Rt,N . We begin with the following estimates of Arias de Reyna [1] on the
R −s eiπw2
term N.N+1 ewπiw −e −πiw appearing in (38):

Proposition 6.2. Let σ be real and T 0 > 0, and define the quantities
(48) s B σ + iT 0
r
T0
(49) aB

(50) N B bac
(51) p B 1 − 2(a − N)
T0 T0 T0 π
!!
(52) U B exp −i log − − .
2 2π 2 8
Let K be a positive integer. Then we have the expansion
2
 K 
Z
w−s eiπw X Ck (p, σ) 
πiw − e−πiw
= (−1) N−1
Ua −σ 

k
+ RS K (s) 
N.N+1 e a
 
k=0

where C0 (p, σ) = C0 (p) is independent of σ and is given explicitly by the formula


p2 3 √
eπi( 2 + 8 ) − i 2 cos πp
2
(53) C0 (p) B
2 cos(πp)
(removing the singularities at p = ±1/2), while for k ≥ 1 the Ck (p, σ) are complex numbers
obeying the bounds
√ σ
2 9 Γ(k/2)
(54) |Ck (p, σ)| ≤
2π 2k
for σ > 0 and
1
2 2 −σ Γ(k/2)
(55) |Ck (p, σ)| ≤
2π 2π((3 − 2 log 2)π)k/2
for σ ≤ 0, while the error term RS K (s) is a complex number obeying the bounds
1 Γ((K + 1)/2)
(56) |RS K (s)| ≤ 23σ/2
7 (a/1.1)K+1
for σ ≥ 0, and
!d−σe
1 9 Γ((K + 1)/2)
(57) |RS K (s)| ≤
2 10 (a/1.1)K+1
if σ < 0 and K + σ ≥ 2.
Proof. This follows from [1, Theorems 3.1, 4.1, 4.2] combined with [1, (3.2), (5.2)]. The
dependence of Ck (p, σ), k ≥ 1 on σ and the dependence of RS K (s) on s is suppressed in
[1], but can be discerned from the definitions of these quantities (and the related quantities
g(τ, z), Pk (z) = Pk (z, σ), RgK (τ, z)) in [1, (3.9), (3.10), (3.7), (3.6)]. 
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 23

Note that p ranges in the interval [−1, 1]. One can show that
1
(58) |C0 (p)| ≤
2
for all p ∈ [−1, 1]; this follows for instance from the n = 0 case of [1, Theorem 6.1]. See also
Figure 10.

Figure 10. Plot of |C0 (p)| for −1 ≤ p ≤ 1.

Informally, the above proposition (and (38), (6)) yield the approximation
1 s(s − 1) −s/2  s 
R0,N (s) ≈ π Γ (−1)N−1 Ua−σC0 (p)
8 2 2
≈ (−1)N−1 U M0 (s)a−σC0 (p).

If one writes s = σ + iT , then by using the approximation α(s) ≈ 1


2
iT
log 2π for the log-derivative
of M0 , one can then obtain the approximate formula
R0,N (s) ≈ (−1)N−1 Ueπiσ/4 M0 (iT )C p (p).
In fact we have the more general approximation
tπ2
!
πiσ/4
Rt,N (s) ≈ (−1) N−1
Ue exp M0 (iT 0 )C p (p)
64
πt
where T 0 B T + 8. More precisely, we have
Proposition 6.3 (Estimate for Rt,N ). Let 0 ≤ σ ≤ 1, let T ≥ 100, and let 0 < t ≤ 1/2. Set
πt
T0 B T +
8
24 D.H.J. POLYMATH

and then define a, N, p, U, C0 (p) using (49), (50), (52), (53). Then
tπ2
!
πiσ/4
Rt,N (σ + iT ) = (−1) Ue
N−1
exp M0 (iT 0 ) (C0 (p) + O≤ (ε̃(σ + iT )))
64
where
0.397 × 9σ
! !
5 3.49
(59) ε̃(σ + iT ) B + exp .
a − 0.125 3(T 0 − 3.33) T 0 − 3.33
Proof. We apply (41) with βN := πi/4 to obtain

tπ2 √
!Z !
tvπi 1 2
Rt,N (σ + iT ) = exp exp − R0,N (σ + iT 0 + tv) √ e−v dv.
64 R 4 π
From (38) we have
K √ 
√ 1 s (s − 1) s  √  v
C (p, σ + tv)
π−sv /2 Γ
v v v
X k

R0,N (σ+iT 0 + tv) = (−1)N−1 Ua−σ− tv  +

RS (s )
 
k Kv v
8 2 2 k=0
a 

for any positive√integer Kv that we permit to depend (in a measurable fashion) on v, where
sv B σ + iT 0 + tv. From (6) and Lemma 5.1(v) we thus have
K √ 
√ √  v
σ +
!!
1 X C k (p, tv) 
R0,N (σ+iT 0 + tv) = M0 (sv ) exp O≤ N−1 −σ− tv 
+  .

(−1) Ua RS (s )

 Kv v
12(T 0 − 0.33) a k

 
k=0

From (43) and Taylor expansion of the logarithm log M0 defined in (7), we have

√ (σ + tv)2
!!
M0 (sv ) = M0 (iT ) exp α(iT )(σ + tv) + O≤
0 0
.
4(T 0 − 0.33)
From (9), (49) one has
iT 0
! ! !
1 1 1 iπ 3
α(iT ) = O≤
0
+ O≤ 0 + Log = log a + + O≤
2T 0 T 2 2π 4 2T 0
3 6
and hence (bounding 2T 0 by 4(T 0 −0.33) )
√ √ !
√ √ πiσ tvπi 6|σ + tv|
α(iT 0 )(σ + tv) = (σ + tv) log a + + + O≤ .
4 4 4(T 0 − 0.33)
We conclude that
√ √ √
  (σ + tv)2 + 6|σ + tv| + 13 
  

!
tvπi
exp − R0,N (σ + iT + tv) = M0 (iT ) exp O≤ 
0 0  ×
4 4(T 0 − 0.33)
K √ 
v
C (p, σ + tv)
× (−1)N−1 Ueπiσ/4 
X k

k
+ RS Kv (s v )  .
k=0
a 
√ √
Bounding 6|σ + tv| ≤ 3(σ + tv)2 + 3, we have
√ √ √
(σ + tv)2 + 6|σ + tv| + 13 (σ + tv)2 + 56
≤ .
4(T 0 − 0.33) T 0 − 0.33
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 25

Putting all this together, we obtain


tπ2
!
Rt,N (σ + iT ) = (−1)N−1 Ueπiσ/4 exp M0 (iT 0 )×
64
√ √
  (σ + tv)2 + 65  X
    Kv 
σ +
Z
C k (p, tv)  1
 √ e−v2 dv.
× exp O≤  0
 
k
+ RS (s )
Kv v   π
R T − 0.33 
k=0
a
2
We separate the k = 0 term from the rest. By Lemma 5.1(iv) and the fact that √1π e−v integrates
to one, we can write the above expression as
tπ2
!
πiσ/4
(60) Rt,N (σ + iT ) = (−1) Ue N−1
exp M0 (iT 0 ) (C0 (p)(1 + O≤ ()) + O≤ (δ))
64
where √
Z   (σ + tv)2 + 56 
  
 :=  − 1 √1 e−v2 dv
 
exp 
R T 0 − 0.33 π
and √ √
 (σ + tv)2 + 56  X
   Kv 
σ +
Z
|C k (p, tv)|  1
 √ e−v2 dv.
δ := exp  0
 
k
+ |RS Kv (s v )|  π
R T − 0.33 
k=1
a
√ 2
Bounding (σ + tv) ≤ 2σ2 + 2tv2 and using (27) we obtain
 2σ2 + 65 
  !−1/2
2t
 ≤ exp  0  1 − − 1.
T − 0.33  T 0 − 0.33
Applying Lemma 5.1(ii) and using the hypotheses t ≤ 1/2, T ≥ 100, one has
!!
2t 2t
1− 0 = exp O≤ 0
T − 0.33 T − 3.33
and hence
 2σ + t + 56 
 2 
 ≤ exp  0  − 1.
T − 3.33 
With t ≤ 1/2 and 0 ≤ σ ≤ 1, one has 2σ2 + t + 56 ≤ 10
3.
By the mean value theorem we then have
!
10 10
(61) ≤ exp .
3(T 0 − 3.33) 3(T 0 − 3.33)

Now we work on δ. Making the change of variables u B σ + tv, we have
 2 5   K̃u 
Z  u + 6  X |Ck (p, u)|  1
 √ e−(u−σ)2 /t du,
δ= exp  0   + |RS K̃u (u + iT 0 
)|
T − 0.33 k=1 ak
πt
 
R

where K̃u is a positive integer parameter that can depend arbitrarily on u (as long as it is measur-
able, of course).
0
We choose K̃u to equal 1 when u ≥ 0 and max(b−σc + 3, b Tπ c) when u < 0, so that Proposition
6.2 applies. The expression
K̃u
X |Ck (p, u)|
+ |RS K̃u (u + iT 0 )|
k=1
ak
26 D.H.J. POLYMATH

is then bounded by
√ u
2 9 Γ(1/2) 1 3u/2 Γ(1) 0.200 × 9u 0.173 × 23u/2
(62) + 2 ≤ +
2π 2a 7 (a/1.1)2 a a2
for u ≥ 0 and
X 2 12 −u Γ(k/2) 1 Γ((K̃u + 1)/2)
(63) + (9/10)d−ue
2π 2π((3 − 2 log 2)π)k/2 ak 2
1≤k≤K̃u
(a/1.1)K̃u +1

for u < 0. One can calculate that


1
22 1 1
≤ 0.036 ≤
2π 2π 2
and
1
≤ 0.445 ≤ 1.1
((3 − 2 log 2)π)1/2
and hence we can bound (63) by
X Γ(k/2) 1 −u X Γ(k/2)
0.0362−u (0.445)k k
2 .
T0
a 2 T0
(a/1.1)k
1≤k≤ π π ≤k≤−u+4

For u ≥ 0, we can estimate (62) by


0.2 × 9u
!
1 0.865
0.2 × 9 u
+ ≤
a a2 a − 0.865
T0
thanks to Lemma 5.1(i). For u < 0, we observe that if k ≤ 2a2 = π then
Γ(k + 2/2) k Γ(k/2) Γ(k/2)
k+2
= 2 ≤
a 2a ak ak
and hence by the geometric series formula
X Γ(k/2) (0.445)2 Γ(2/2) 0.247
(0.445)k ≤ ≤
T0
ak 1 − (0.445)2 a a2
2≤k≤ π ,k even

and similarly
X Γ(k/2) (0.445)3 Γ(3/2) 0.098
(0.445)k ≤ 2 3

0 a k 1 − (0.445) (a/1.1) a3
3≤k≤ Tπ ,k odd

and hence we can bound (63) by



−u 0.445 π Γ(k/2)
!
0.247 0.098 1 X
0.0362 + 2
+ 3
+ 2−u .
a a a 2 T0
(a/1.1)k
π ≤k≤−u+4

By Lemma 5.1(i) we have



0.445 π 0.247 0.098
!
0.029
0.036 + + ≤
a a2 a3 a − 0.353
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 27

and thus we can bound (63) by


0.029 × 2−u 1 −u X Γ(k/2)
+ 2 (1.1)k .
a − 0.353 2 T0
ak
π ≤k≤−u+4

Putting this together, we conclude that


K̃u
X |Ck (p, u)| 0.2 × 9u 0.029 × 2−u 2−u X Γ(k/2)
+ |RS K̃u (u + iT 0 )| ≤ + + (1.1)k
k=1
ak a − 0.865 a − 0.353) 2 T0
ak
π ≤k≤−u+4

for all u (positive or negative). We conclude that δ ≤ δ1 + δ2 + δ3 , where


 u + 6  0.2 × 9u 1 −(u−σ)2 /t
Z  2 5 
δ1 B exp  0  √ e du
R T − 0.33  a − 0.865 πt
 u + 6  0.029 × 2−u 1 −(u−σ)2 /t
Z  2 5 
δ2 B exp  0  √ e du
R T − 0.33  a − 1.25 πt
 u + 6  2−u
 2 5 
Γ(k/2) 1 −(u−σ)2 /t
Z X
(64) δ3 B exp  0  (1.1)k √ e du.
R T − 0.33 2 T0
ak πt
π ≤k≤−u+4

For δ1 , we translate u by σ to obtain


0.2 × 9σ  u + 2σu + σ +
5
Z  2 2

 1
+ 2u log 3 √ e−u /t du
6 2
δ1 = exp 
a − 0.865 R T 0 − 0.33 πt
and hence by (27)
 σ + 6 σ
0.2 × 9σ t(log 3 + T 0 −0.33
 2 5
)2  
 −1/2
t
(65) δ1 = exp  0
 + t
 1 − .
a − 0.865 T − 0.33 1 − T 0 −0.33 T 0 − 0.33
One can write
1 t t
(66) t =1+ ≤1+ 0
1− T 0 −0.33
T0 − 0.33 − t T − 0.83
while by Lemma 5.1(ii) we have
t   t    t 
(67) 1− 0 = exp O≤ 0 = exp O≤ 0 .
T − 0.33 T − 0.33 − t T − 0.83
We conclude that
0.2 × 9σ 5 + 3t + 6σ2  σ 2  t !
δ1 ≤ exp + t log 3 + 0 1+ 0 .
a − 0.865 6(T 0 − 0.83) T − 0.33 T − 0.83
From Lemma 5.1(i) and the hypothesis 0 ≤ σ ≤ 1, we have
 
 σ 2 2σ/ log 3
log 3 + 0 ≤ (log 3) 1 + 0
2
 
σ 

T − 0.33 T − 0.33 − 2 log 3 
!
2σ/ log 3
≤ (log2 3) 1 + 0
T − 0.83
28 D.H.J. POLYMATH

and therefore by a further application of Lemma 5.1(i)


+t
 2σ

 σ 2  t  
log 3

log 3 + 0 1+ 0 ≤ log2 3 1 + 2σt/ log 3 

T − 0.33 T − 0.83 T 0 − 0.83 − 2σ/ 
log 3+t

+ t 
 
log 3

≤ log2 3 1 + 0 
T − 0.83 − t 

log 3 + t 

 2σ 

≤ log 3 1 + 0
2   
T − 1.33 
and thus
0.2 × 9σ exp(t log2 3) 5 + 3t + 6σ2 + 12tσ log 3 + 6t2 log2 3
!
δ1 ≤ exp .
a − 0.865 6(T 0 − 1.33)
By repeating the proof of (65), we have
  √ σ
2 
0.029 × 2 −σ  σ2 + 5
6
t − log 2 + 0
T −0.33 
 
t −1/2
δ2 = exp  0 + 1 − .
 
a − 0.353  T − 0.33 t T 0 − 0.33
1 − T 0 −0.33


√ σ

We can bound (− log 2+ T 0 −0.33 )
2 by log2
2. Using (66), (67) we thus have

0.029 × 2−σ exp(t log2 2) 5 + 3t + 6σ2
!
δ2 ≤ exp .
a − 0.353 6(T 0 − 1.33)
With t ≤ 1/2 and 0 ≤ σ ≤ 1 one has
0.2 exp(t log2 3) ≤ 0.366

0.029 exp(t log2 2) ≤ 0.031
5 + 3t + 6σ2 5 + 3t + 6σ2 + 12tσ log 3 + 6t2 log2 3
≤ ≤ 3.49
6 6
and hence
0.366 × 9σ
!
3.49
δ1 ≤ exp 0
a − 0.865 T − 1.33
and
0.031 × 2−σ
!
3.49
δ2 ≤ exp 0 .
a − 0.353 T − 1.33
Now we turn to δ3 , which will end up being extremely small compared to δ1 or δ2 . By (64)
and the Fubini-Tonelli theorem, we have
 u + 6
 2 5 
k Γ(k/2) (u − σ)2
Z 4−k
1 X
δ3 = √

(1.1) exp  0 − − u log 2 du.
2 πt T 0 a k
−∞ T − 0.33 t
k≥ 2.2π

T0
Since u ≤ 4 − k, k ≥ 2.2π , and T 0 ≥ T ≥ 100, we have k ≥ 14 and u ≤ −10; since σ ≥ 0, we may
u2 + 65 u2
thus lower bound (u − σ)2 /t by u2 /t. Since t ≤ 1/2, we can upper bound T 0 −0.33 − t by (say)
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 29

2
− u2t , thus

k Γ(k/2)
Z 4−k
1 X u2
δ3 ≤ √ (1.1) e− 2t −u log 2 du.
2 πt T 0 ak
−∞
k≥ 2.2π

u2 (k−4)u
We can bound e− 2t ≤ e 2t , in the range of integration and thus
Z 4−k
u2 1 (k−4)2 1
2t +(k−4) log 2
2 +(k−4) log 2
e− 2t −u log 2 du ≤ k−4
e− ≤ k−4
e−(k−4) ;
−∞ 2t − log 2 2t − log 2

bounding
k−4 k − 4 − 2t log 2 k − 6
− log 2 = ≥
2t 2t 2t
we conclude that
√ X
t Γ(k/2) −(k−4)2 +(k−4) log 2
δ3 ≤ √ (1.1)k e .
π T0 (k − 6)ak
k≥ 2.2π

2 +(k−4) log 2
For k ≥ 14 one can easily verify that (1.1)k Γ(k/2)e−(k−4) ≤ 10−30 ; discarding the √
t
π
1
and k−6 factors we thus have

X 10−30 2 × 10−30
δ3 ≤ ≤
k≥14
ak a14

(say). Since
0.031 × 2−σ 2 × 10−30 0.031 × 2−σ
+ ≤
a − 0.353 a14 a − 0.865
we thus have

0.366 × 9σ + 0.031 × 2−σ


!
3.49
δ ≤ δ1 + δ2 + δ3 ≤ exp 0 .
a − 0.865 T − 1.33

Inserting this and (61), (58) into (60), and crudely bounding 2−σ by 9σ , we obtain the claim. 

6.3. Combining the estimates. Combining Propositions 6.1, 6.3 with (39) and the triangle
inequality (and noting that M0 = M0∗ , Mt = Mt∗ and α = α∗ , and that U has magnitude 1), we
conclude the following “A + B − C approximation to Ht ”:

Corollary 6.4 (A + B − C approximation). Let t, x, y obey (5). Set


x πt
(68) T0 B +
2 8
30 D.H.J. POLYMATH

and then define a, N, p, U, C0 (p) using (49), (50), (52), (53). Define the quantities
1 + y − ix
s+ = s+ (x + iy) B
2
1 − y + ix
s− = s− (x + iy) B
2
N
X btn
At (x + iy) B Mt (s− )
s− + 2t α(s− )
n=1 n
N
X btn
Bt (x + iy) B Mt (s+ )
n s+ + 2 α(s+ )
t
n=1
tπ2
!
Ct (x + iy) B 2e −πiy/8 N
(−1) exp Re(M0 (iT 0 )C0 (p)Ueπi/8 )
64
where M0 , btn were defined in (10), (15). Then
Ht (x + iy) = At (x + iy) + Bt (x + iy) − Ct (x + iy) + O≤ (E A (x + iy) + E B (x + iy) + EC (x + iy))
where
N
X btn
E A (x + iy) B |Mt (s− )| 1−y t
εt,n (s− )
n=1 n 2 + 2 Reα(s− )

N
X btn
E B (x + iy) B |Mt (s+ )| 1+y t
εt,n (s+ )
n=1 n 2 + 2 Reα(s+ )

tπ2
!
EC (x + iy) B exp |M0 (iT 0 )| (ε̃(s− ) + ε̃(s+ ))
64
and εt,n , ε̃ were defined in (44), (59).
In our applications, we will just use the cruder “A + B” approximation that is immediate from
the above corollary and (58):
Corollary 6.5 (A + B approximation). With the notation and hypotheses as in Corollary 6.4, we
have
Ht (x + iy) = At (x + iy) + Bt (x + iy) + O≤ (E A (x + iy) + E B (x + iy) + EC,0 (x + iy))
where
tπ2
!
EC,0 (x + iy) B exp |M0 (iT 0 )| (1 + ε̃(s− ) + ε̃(s+ )) .
64
We can now prove Theorem 1.3. Dividing by the expression Bt from (11), and using (14), we
conclude that
Ht (x + iy) Ct (x + iy)
(69) = ft (x + iy) − + O≤ (eA + eB + eC )
Bt (x + iy) Bt (x + iy)
and
Ht (x + iy)
= ft (x + iy) + O≤ eA + eB + eC,0

(70)
Bt (x + iy)
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 31

where
N
X btn
(71) eA B eA (x + iy) B |γ| ny εt,n (s− )
n=1
nRe(s)+Re(κ)
N
X btn
(72) eB B eB (x + iy) B ε (s )
Re(s) t,n +
n=1
n
 2
exp tπ 0
64 |M0 (iT )|
(73) eC B eC (x + iy) B (ε̃(s− ) + ε̃(s+ )) .
|Mt (s+ )|
 2
exp tπ 0
64 |M0 (iT )|
(74) eC,0 B eC,0 (x + iy) B (1 + ε̃(s− ) + ε̃(s+ )) ,
|Mt (s+ )|
and where γ, s∗ , κ were defined in (16), (17), (18). Note also from (68), (49), (50) that N is given
by (19).
To conclude the proof of Theorem 1.3 it thus suffices to obtain the following estimates.
Proposition 6.6 (Estimates). Let the notation and hypotheses be as above.
(i) One has
 −y/2
0.02y x
|γ| ≤ e

(ii) One has
1+y t x (1 − 3y + 4y(1+y)
x2
)+ t
Res∗ ≥ + log − .
2 4 4π 2x2
(iii) One has !
ty
κ = O≤ .
2(x − 6)
(iv) One has
 t2
 16 log2 x 2 + 0.626 
N
  
X b t  
n 4πn  − 1 .
eA ≤ |γ|N |κ| ny Re(s exp 
x
n=1
n ∗)  
2 − 6.66
 

(v) One has


 t2
 log2 x 2 + 0.626 
N
  
X btn 
exp  16 4πn

 − 1 .
eB ≤
n=1
nRe(s∗ )   x − 6.66  

(vi) One has


 x − 1+y x
3| log 4π + i π2 | + 3.58 1.24 × (3y + 3−y )
! !
4 t 2 x 6.92
eC ≤ exp − log + + .
4π 16 4π x − 8.52 N − 0.125 x − 6.66
and
 x − 1+y x
3| log 4π + i π2 | + 3.58 1.24 × (3y + 3−y )
! !
4 t 2 x 6.92
eC,0 ≤ exp − log + 1+ + .
4π 16 4π x − 8.52 N − 0.125 x − 6.66
Note that to obtain the bound (24) from Proposition 6.6(vi) we may simply use the inequality
1 + u ≤ exp(u) for any u ∈ R, and then bound x−6.66
1 1
≤ x−8.52 .
32 D.H.J. POLYMATH
   1+y+ix 
Proof. From the mean value theorem (and noting that Mt = Mt∗ , so that Mt 1+y−ix 2
= Mt
2
),
we have 
d ix 
log |γ| = −y log Mt σ +
dσ 2
for some 1−y 1+y
2 ≤ σ ≤ 2 . From (8), (10) we have

d ix  t  ix  0  ix   ix 
log Mt σ + = Re α σ + α σ+ +α σ+ .
dσ 2 2 2 2 2
From (43) one has
!
ix 
 1
(75) α σ+ 0
= O≤
2 x−6
and from Taylor expansion we also have
ix  ix   σ 
α(σ + ) = α + O≤ ;
2 2 x−6
from (9) one has
π
! ! !
 ix  1 1 1 ix 1 x 2
α = O≤ + O≤ + Log = log + i + O≤
2 x x 2 4π 2 4π 4 x
and hence
π 2+σ
!
ix 1 x
(76) α(σ + ) = log + i + O≤ .
2 2 4π 4 x−6
Inserting these bounds, we conclude that
π 2+σ
!! !!!
1 x t
log |γ| = −yRe log + i + O≤ 1 + O≤ .
2 4π 4 x−6 2(x − 6)
Expanding this out, we have
t(2+σ) 
 2 + σ + 4t log 4πx
+ tπ8 + 2(x−6)
 
 1 x 
log |γ| = −y  log + O≤   .
2 4π x−6
In the region (5), which implies that 0 ≤ σ ≤ 1, we have
tπ t(2 + σ)
2+σ+ + ≤ 3.21
8 2(x − 6)
and thus
y x
t x
log 4π + 3.21
log |γ| ≤ − log +y4 .
2 4π x−6
log x
The function x 7→ x−64π is decreasing for x ≥ 200 thanks to Lemma 5.1(vi), hence
x
log 4πt
+ 3.21 t
4π + 3.21
log 200
y4 ≤y4 ≤ 0.02y.
x−6 200 − 6
Claim (i) follows. We remark that one can improve the e0.02y factor here by Taylor expanding α
to second order rather than first order, but we will not need to do so here.
To prove claim (ii), it suffices by (48) to show that
1 x (1 − 3y)+ 4y(1 + y)
Reα(s+ ) ≥ log − − .
2 4π x2 x4
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 33

By (9) one has


(1 + y)2 + x2
p
1+y 2(1 − y) 1
Reα(s+ ) = − + log .
(1 + y) + x
2 2 (1 − y) + x
2 2 2 4π
(1 + y)2 + x2 ≥ x and calculate
p
We bound
1+y 2(1 − y) 1 − 3y 4y(1 + y)
− =− −
(1 + y) + x
2 2 (1 − y) + x
2 2 (1 + y) + x
2 2 ((1 + y) + x2 )((1 + y)2 + x2 )
2
4y(1+y)
1 − 3y + x2
≥− .
(1 + y) + x2
2

Lower bounding the numerator by its nonnegative part and then lower bounding (1 + y)2 + x2 by
x2 , we obtain the claim.
Claim (iii) is immediate from (75) and the fundamental theorem of calculus. Now we turn to
(iv), (v). From (76) one has
1 ± y + ix π
! !
1 x 3
α − log n = log + i + O≤
2 2 4πn2 4 x−6
for either choice of sign ±. In particular, we have
π
2 + i2|
2 x
1 ± y + ix π2

3| log 4πn
! !
1 2 x 9
α = + + + .

(77) − log n log O ≤
2 4 4πn2 16 x−6 (x − 6)2

For any 1 ≤ n ≤ N, we have


πt
x + 16
1≤n ≤N ≤a =
2 2 2
;

x 2
in the region (5), the right-hand side is certainly bounded by ( 4π ) , so that
4π x x
≤ 2

x 4πn 4π
and hence
log x + i π ≤ log x + i π .

4πn2 2 4π 2
x
| log 4π +i π2 |
In the region (5) we have x ≥ 200, we see from Lemma 5.1(vi) (after squaring) that x−6 is
decreasing in x. Thus
π π
π2 3| log 4πn2 + i 2 | π2 3| log 4π + i 2 |
x 200
9 9
+ + 2
≤ + +
16 x−6 (x − 6) 16 200 − 6 (200 − 6)2
≤ 0.667.
Similarly, in (5) we also have
t2 t 1
× 0.667 + + ≤ 0.313.
8 4 6
We conclude from (44) that
 t2
 32 log2 x 2 + 0.313 

1 ± y + ix
!
4πn
εt,n ≤ exp   − 1.
2 T − 3.33 

Inserting this bound into (71), (72), we obtain claims (iv), (v).
34 D.H.J. POLYMATH

Now we establish (vi). From (10) we have


 2
exp tπ 0
64 |M0 (iT )| tπ2 t
! 0
2 |M0 (iT )|
= exp − Reα(s+ ) .
|Mt (s+ )| 64 4 |M0 (s+ )|
πit
Note that 1+y+ix
2 = iT 0 + 1+y 0 1
2 − 8 . From (43) we see that |α (s)| ≤ x−6 for any s on the line
segment between iT 0 and 1+y+ix2 . From Taylor’s theorem with remainder applied to a branch of
log M0 , and noting that |M0 (s+ )| = |M0 ( 1+y+ix
2 )|, we conclude that
 1+y πit 
2 + 8 | 

2 
|M0 (iT 0 )| + πit  | −
! !
 1 y
= exp Re − + α(iT ) + O≤ 
0
 .

|M0 (s+ )| 2 8 2(x − 6) 
For 0 ≤ y ≤ 1 and 0 < t ≤ 1
2 we have
1+y πit 2
|− 2 + 8 |
≤ 0.52
2
and from (9) one has
iT 0 1 T 0 iπ
! !
1 1 1 3
α(iT ) = O≤ 0
0
+ O ≤ 0
+ Log = log + + O≤ ( 0 )
2T T 2 2π 2 2π 4 2T
and hence
 3| − 1+y + πit
  
|M0 (iT 0 )|  1 + y 0 2
T tπ 2 8 | 0.52 
= exp −
 log − + O≤  +  .
|M0 (s+ )| 4 2π 32 2T 0 x − 6 
1+y πit
Bounding 1
2T 0 ≤ 1
x−6 and | − 2 + 8 |
≤ 1.02, this becomes
!− 1+y
|M0 (iT 0 )| T0 4 tπ2
!!
3.58
= exp − + O≤
|M0 (s+ )| 2π 32 x−6
and hence
tπ2 !− 1+y
 
|M0 (iT 0 )| !2
 !
exp 64 T0 4  tπ2 t 1 + y + ix 3.58
= + O≤  .

exp − − Reα
|Mt ( 1+y+ix
2 )|
2π 64 4 2 x−6 
By repeating the proof of (77) we have
1 ± y + ix 2 1 π2 3| log 4πx
+ i π2 |
!
2 x 9
Reα( ) = log − + O≤ + .
2 4 4π 16 x−6 (x − 6)2
As before, in the region (5) we have
π
x
3| log 4πn 2 + i2| 9
x
3| log 4π + i π2 | 9
+ 2
≤ +
x−6 (x − 6) x−6 (x − 6)2
and thus
 2 !− 1+y
exp tπ 0
64 |M0 (iT )| T0 4 3| log 4π x
+ i π2 | + 3.58
!!
t 2 x 9
= exp − log + O≤ +
|Mt (s+ )| 2π 16 4π x−6 (x − 6)2
1+y
x
+ i π2 | + 3.58
!−
T0 4 3| log 4π
!!
t x
= exp − log2 + O≤
2π 16 4π x − 8.52
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 35

x
thanks to Lemma 5.1(i). Finally, since T 0 ≥ ≥ 100 in (5), one has
2
!
3.49
exp 0 ≤ 1.037
T − 3.33
and hence by (59)
1 ± y + ix 1.24 × 3±y 1.73
ε̃( )≤ + 0 .
2 a − 0.125 T − 3.33
Hence
1.24 × (3y + 3−y ) 3.46
ε̃(s+ ) + ε̃(s− ) ≤ + 0
a − 0.125 T − 3.33
giving the claim (substituting T 0 = x/2 and a ≥ N). 

7. Bounding Dirichlet series


In view of Corollary 1.4, it is of interest to obtain lower bounds for the quantity ft (x + iy)
defined in (14) for x, y, t in the region (5). By the triangle inequality (and the trivial identity
|z| = |z|), we obtain the lower bound
 N
N

 X bt X b t 
| ft (x + iy)| ≥  n
− |γ| ny n  .

s
n=1 n
∗ n=1 n s ∗ +κ
+

It is thus of interest to obtain lower bounds for differences
 N
N

 X βn X α n

(78) ∆ B 


n s n s 
n=1 n=1 +

of magnitudes of Dirichlet series for various coefficients βn , αn . Our tools for this will be as
follows.
Lemma 7.1. Let N be a natural number, let s = σ + iT be a complex number for some real σ, T ,
and let αn , βn be complex numbers for n = 1, . . . , N with β1 = 1. Let ∆ denote the quantity (78).
(i) (Triangle inequality) We have
N
X |αn | + |βn |
∆ ≥ 1 − α1 − .
n=2

(ii) (Refined triangle inequality) If the αn , βn are all real and 0 ≤ α1 < 1, then we have
N max(|β − α |, 1−α1 |β + α |)
X n n 1+α n n
∆ ≥ 1 − α1 − 1
.
n=2

(iii) (Dirichlet mollifier) If λ1 , . . . , λD are complex numbers, not all zero, then
∆˜
∆ ≥ PD |λ |
d
d=1 dσ
where  DN DN 
 X β̃n X α̃n 
∆˜ B  −

s

s 
n n
n=1 n=1 
+
36 D.H.J. POLYMATH

and α̃n , β̃n are the Dirichlet convolutions of αn , βn with the λd :


X
α̃n B λd αn/d
1≤d≤D:d|n
X
β̃n B λd βn/d .
1≤d≤D:d|n

Proof. The claim (i) is immediate from the triangle inequality.


Now we prove (ii). We may assume that the right-hand side is positive, as the claim is trivial
otherwise. By a continuity argument (replacing βn , αn for n ≥ 2 by tβn , tαn with t increasing con-
tinuously from zero to one, noting that this only increases the right-hand side of the inequality)
it suffices to verify the claim when ∆ is positive. In this case, we may write

N
X βn − eiθ αn
∆ =
n=1 ns
for some phase θ. By the triangle inequality, we then have
N
X |βn − eiθ αn |
∆ ≥ |1 − e α1 | −

.
n=2

We factor out |1 − eiθ α1 |, which is at least 1 − α1 , to obtain the lower bound


 N

 X |β − e iθ α |/|1 − eiθ α | 
n n 1
∆ ≥ (1 − α1 ) 1 −  .

n=2

By the cosine rule, we have
 2 β2n + α2n − 2αn βn cos θ
|βn − eiθ αn |/|1 − eiθ α1 | = .
1 + α21 − 2α1 cos θ
This is a fractional linear function of cos θ with no poles in the range [−1, 1] of cos θ. Thus this
function is monotone on this range and attains its maximum at either cos θ = +1 or cos θ = −1.
We conclude that
|βn − eiθ an | |βn − αn | |βn + αn |
≤ max( , )
|1 − eiθ α1 | 1 − α1 1 + α1
and the claim follows.
For claim (iii), we recall the well-known relationship
DN
 D  N 
X α̃n X λd  X αn 
s
=  s
  s

n=1
n d=1
d n=1
n
DN
 D  N 
X β̃n   X λd  
 X βn 
=    
n=1
n s  d=1 d s   n=1 n s 
between Dirichlet convolution and Dirichlet series, which implies that

D
X λ d

∆˜ = ∆.
d=1 d s
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 37

Since ∆,
˜ ∆ are non-negative and

D D
X λd X |λd |
≤ ,
d=1 d s d=1 dσ

the claim follows. 


Returning to the estimation of ft (x + iy), we conclude from Lemma 7.1(i) with s replaced by
s∗ , βn replaced by btn , and αn replaced by |γ|ny−κ btn that
N N
X btn X btn
(79) | ft (x + iy)| ≥ 2 − − |γ| ,
n=1
nσ n=1
nσ−y−|κ|
where σ B Re s∗ . This rather crude bound will suffice when x is very large, particularly when
combined with the estimates in Proposition 6.6. For smaller values of x, we would like to use
parts (ii) and (iii) of Lemma 7.1. A technical difficulty arises because the quantity |λ|ny−κ btn
quantity need not be real, so that Lemma 7.1(ii) is not directly available. However, by writing
n−κ = 1 + O≤ (n|κ| − 1)
we see from the triangle inequality that
 N
N

N
 X btn X |γ|btn ny  X btn (n|κ| − 1)
| ft (x + iy)| ≥  − − |γ| .
nσ−y
 
n=1 n s∗ n=1 n s∗ 

+ n=1
Assuming for now that |γ| < 1 (which in practice will follow from Proposition 6.6(i)), we can
then apply Lemma 7.1(iii) follows by Lemma 7.1(ii) to conclude that
1−α̃
max(|β̃n −α̃n |, 1+α̃1 |β̃n +α̃n |)
1 − α̃1 − N
PN
n=2 nσ
1 X bt (n|κ| − 1)
(80) | ft (x + iy)| ≥ − |γ| n
PD |λd | nσ−y
d=1 dσ n=1
for any real numbers λ1 , . . . , λD with λ1 = 1, where
X
α̃n B λd btn/d |γ|ny
1≤d≤D:d|n
X
β̃n B λd btn/d .
1≤d≤D:d|n
PD λd
In practice, it has proven convenient to use this estimate with Dirichlet mollifiers d=1 d s that
are partial Euler products of the form
D
λd Y  btp 
X  
(81) s
= 1 − s 
d=1
d p≤P
p
for some small prime P, where the product is over primes p up to P. For instance, if P = 3,
then we would take D = 6, λ1 = 1, λ2 = −bt2 , λ3 = −bt3 , λ6 = bt2 bt3 , and all other λd vanishing.
This choice achieves a substantial amount of cancellation in the β̃n coefficients, which we have
found to make the lower bound in (80) favorable. (For instance, it makes β̃ p vanish for all primes
p ≤ P.) In the literature one also sees other choices of mollifier than this Euler product used,
for instance to control the extreme values of Dirichlet polynomials; however our numerical
experimentations with alternative mollifiers to (81) turned out to give inferior results for our
application.
38 D.H.J. POLYMATH

The above calculations also suggest that the quantity ft (x + iy) oscillates in inverse proportion
to the Euler product (81). Thus, to minimise the relative size of the right-hand side of (25) with
respect to the left-hand side, it would therefore seem to be advantageous to try to work as much
x
as possible in regions where this product is small, which heuristically corresponds to 4π log p
s
being close to an integer for p ≤ P (so that p has argument close to zero). In our subsequent
arguments, we will erect a barrier for x in the vicinity of 6 × 1010 (this number being largely
dictated by the numerical verifications of the Riemann hypothesis in the literature); we will in
fact place the barrier at
X B 6 × 1010 + 83952 − 0.5
X
noting that the fractional parts { 4π log p} for p ≤ 11 are somewhat close to zero:
X
log 2} = 0.0275 . . .
{

X
{ log 3} = 0.0437 . . .

X
{ log 5} = 0.0640 . . .

X
{ log 7} = 0.0774 . . .

X
{ log 11} = 0.0954 . . .

We found this shift by the following somewhat ad hoc procedure. We first introduced the quan-
tity

Y 1
,
eulerprod(x, pn ) B 1
p≤pn 1 − p1−ix/2

which is the exponent corresponding to y = 1 (where the minimum value of | ft (x + iy)| in the
barrier region is expected to occur). We numerically located candidate integers 1 ≤ q ≤ 105 for
which the quantity
min |eulerprod(x, 29)|
x−6×1010 −q∈{−0.5,0,0.5}

exceeded a threshold (we chose 4), to obtain seven candidates for q: 1046, 22402, 24198, 52806,
77752, 83952, and 99108. Among these candidates, we selected the value of q which maximised
the quantity
min | f0 (x + i)|,
x−6×1010 −q∈{−0.5,0,0.5}

namely q = 83952 (this quantity being ≈ 4.32 for this value of q).

8. Estimating a sum
In order to use the bound (79) for very large values of N, the following estimate will be used.
Lemma 8.1. Let N ≥ N0 ≥ 1 be natural numbers, and let σ, t > 0 be such that
t
σ > log N.
2
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 39

Figure 11. Improving approximation of epn (x) w.r.t. | f0 (x+iy)| as n is increased,


where epn (x) = eulerprod(x, pn ), shown near X = 6 × 1010 + 83951.5, which
was chosen as a barrier location.

Then
N N0
X btn X btn N
σ
≤ σ
+ max(N01−σ btN0 , N 1−σ btN ) log .
n=1
n n=1
n N0
Proof. From the identity
 
t
btn exp 4 (log N − log n)2 − 4t (log N)2
=
nσ nσ− 2 log N
t

btn
we see that the summands nσ are decreasing for 1 ≤ n ≤ N, hence by the integral test one has
N N0 Z N t
X btn X btn ba
(82) σ
≤ σ
+ σ
da.
n=1
n n=1
n N0 a

Making the change of variables a = eu , the right-hand side becomes


N0
X btn t
σ
exp((1 − σ)u + u2 ) du.
n=1
n 4
The expression (1 − σ)u + 4t u2 is convex in u, and is thus bounded by the maximum of its values
at the endpoints u = log N0 , log N; thus
t
exp((1 − σ)u + u2 ) ≤ N01−σ btN0 , N 1−σ btN .
4
The claim follows. 
Remark 8.2. The right-hand side of (82) can be evaluated exactly as

2 log N − σ + 1 2 log N0 − σ + 1
N0 t t
btn π −(σ − 1)2
X ! !!
+ √ exp( ) erfi √ − erfi √ .
n=1
nσ t t t t
40 D.H.J. POLYMATH

PN btn
In practice, this upper bound for n=1 nσ is slightly more accurate than the one in Lemma 8.1,
and is a good approximation even for relatively small values of N0 (e.g., N0 = 100). However,
the cruder bound above suffices for the numerical values of parameters needed to establish the
bound Λ ≤ 0.22.
8.1. Fast evaluation of multiple values of ft (s). Fix t ≥ 0. For the verification of the barrier
criterion (Theorem 1.2(iii)) using Corollary 1.4, we will need to evaluate the quantity ft (s) to
reasonable accuracy for a large number of values of s in the vicinity of a fixed complex number
X + iy. From (14) we have
N N
X nb btn X na btn
(83) ft (s) = 1+y−iX
+ γ(s) 1−y+iX
,
n=1 n 2 n=1 n 2

where btn is given by (15), γ(s) is given by (16), N is given by (19) and
1 + y − iX
b = b(s) B − s∗
2
and
1 − y − iX
a = a(s) B − s∗ − κ
2
with s∗ , κ defined by (17), (18). In practice the exponents a, b will be rather small, and N will be
fixed (in our main verification we will in fact have N = 69098).
A naive computation of ft (s) for M values of s would take time O(N M), which turns out to
be somewhat impractical for for the ranges of N, M we will need; indeed, for our main theorem,
the total number of pairs (t, s) at which we need to perform the evaluation is 785052 (spread out
over 152 values of t), and direct computation of all this data required 78.5 hours of computer
time, which was still feasible at this order of magnitude of X but would not scale to significantly
higher magnitudes. However, one can significantly speed up the computation (to about 0.025
hours) to extremely high accuracy by using Taylor series expansion to factorise the sums in (83)
into combinations of sums that do not depend on s and thus can be computed in advance.
We turn to the details. To make the Taylor series converge2 faster, we recenter the sum in n,
writing
XN b(N+1)/2c
X
F(n) = F(n0 + h)
n=1 h=−bN/2c+1
for any function F, where n0 B bN/2. We thus have
ft (s) = B(b) + γ(s)A(n0 , a)
where
b(N+1)/2c
X (n0 + h)b btn0 +h
B(b) B 1+y−iX
h=−bN/2c+1 (n0 + h) 2

and
b(N+1)/2c
X (n0 + h)a btn0 +h
A(a) B 1−y+iX
.
h=−bN/2c+1 (n0 + h) 2

2One can obtain even faster speedups here by splitting the summation range PN
n=1 into shorter intervals and using
a Taylor expansion for each interval, although ultimately we did not need to exploit this.
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 41

We discuss the fast computation of B(b) for multiple values of b; the discussion for A(a) is
analogous. We can write the numerator (n0 + h)b btn0 +h as
t
log2 (n0 + h));
exp(b log(n0 + h) +
4
writing log(n0 + h) = log n0 + log(1 + nh0 ), this becomes
b+ 4t log n0 t h t h
n0 exp( log2 (1 + )) exp((b + log n0 ) log(1 + )).
4 n0 2 n0
By Taylor expanding3 the exponentials, we can write this as
∞ ( t log2 (1 + h i
h (b + 2 log n0 ) j
∞ X t
b+ t log n0 n0 ))
X 4
n0 4 log j (1 + )
i=0 j=0
i! n0 j!
and thus the expression B(b) can be written as
b+ 4t log n0
X∞ X ∞
(b + 2t log n0 ) j
B(b) = n0 Bi, j
i=0 j=0
j!
where
b(N+1)/2c
X ( 4t log2 (1 + h i
n0 )) log j (1 + h
n0 )
Bi, j B 1+y−iX
.
h=−bN/2c+1 (n0 + h) i!2

If we truncate the i, j summations at some cutoff E, we obtain the approximation


b+ 4t log n0
E−1 X
X E−1
(b + 2t log n0 )i
B(b) ≈ n0 Bi, j (n0 ) .
i=0 j=0
i!

The quantities Bi, j , i, j = 0, . . . , E − 1 may be evaluated in time O(NE 2 ), and then the sums B(b)
for M values of b may be evaluated in time O(ME 2 ), leading to a total computation time of
O((N + M)E 2 ) which can be significantly faster than O(N M) even for relatively large values of
E. We took E = 50, which is more than adequate to obtain extremely high accuracy for ft (s);
see ???. The code for implementing this may be found in the file
dbn upper bound/pari/barrier multieval t agnostic.txt
in the github repository [19].

9. A new upper bound for the de Bruijn-Newman constant


In this section we prove Theorem 1.1. As stated in the introduction, it suffices to verify
the conditions (i), (ii), (iii) of Theorem 1.2 t0 B 0.2, X B X0 − 0.5, and y0 B 0.2, where
X0 B 6 × 1010 + 83952.
Claim (i) is immediate from the result of Platt [17] that all the non-trivial zeroes of ζ with
imaginary part between 0 and 3.06×1010 lie on the critical line {Re(s) = 1/2}. For the remaining
claims (ii), (iii), we need to verify that Ht (x + iy) , 0 for two regions of (x, y, t):
√ √
(ii) x ≥ X0 − 0.5 + 0.96, 0.2 ≤ √ y ≤ √0.6, and t = 0.2. √
(iii) X0 − 0.5 ≤ x ≤ X0 − 0.5 + 0.96, 0.04 + 2(0.2 − t) ≤ y ≤ 0.6, and 0 ≤ t ≤ 0.2.
3It is also possible to proceed by just performing Taylor expansion on the second exponential and leaving the first
exponential untouched; this turns out to lead to a comparable numerical run time.
42 D.H.J. POLYMATH

Both of these regions lie in (5). Enlarging the regions slightly, and applying Corollary 1.4, it
thus suffices to establish the following claim:
Proposition 9.1. Let (x, y, t) lie in one of the following regions:

(ii) (Asymptotic zero free region) x ≥ X0 , 0.2 ≤ y ≤ √ 0.6, and t = 0.2.
(iii) (Barrier) X0 − 0.5 ≤ x ≤ X0 + 0.5, 0.2 ≤ y ≤ 0.6, and 0 ≤ t ≤ 0.2.
Define
N N
X btn X
y bn
t
ft (x + iy) B + γ n
n=1
n s∗ n=1
n s∗ +κ
t
btn B exp( log2 n),
 r 4  
 x t  x πt 1/2
NB 
 + 
 ≤ (1 + ) ,
4π 16  4π 4x
so in particular
x πt
(84) log ≥ 2 log N − log(1 + ),
4π 4x
and let κ, s∗ , γ, eA , eB , eC0 be as in Theorem 1.3, and eA , eB , eC0 . Then
(85) | ft (x + iy)| > eA + eB + eC,0 .
Note that if x ≥ X0 − 0.5 then N ≥ N0 B 69098, with N = N0 when X0 − 0.5 ≤ x ≤ X0 + 0.5.
We calculate a somewhat crude upper bound for the right-hand side of (85):
Lemma 9.2. For (x, y, t) in either of the regions (ii), (iii) in Proposition 9.1, one has
eA + eB + eC,0 ≤ 1.25 × 10−3 .
Proof. From Theorem 1.3 we have
(86) eA + eB ≤ (eδ1 − 1)(F N,t (Re(s∗ )) + |γ|F N,t (Re(s∗ ) − y − |κ|))
where
N
X btn
(87) F N,t (σ) := .
n=1

and
t2
log2 4π
x
+ 0.626
16
(88) δ1 B .
x − 6.66
From Lemma 5.1(vi), the quantity δ1 is monotone decreasing in x in the region (5). Thus we
have
(0.2)2
−0.5
log2 X04π 16+ 0.626
(89) δ1 ≤
X0 − 0.5 − 6.66
whenever x ≥ X0 − 0.5 ≥ 200 and 0 ≤ t ≤ 0.2. Computing the right-hand side, we conclude that
δ1 ≤ 3.12 × 10−11
and hence by Taylor expansion
eδ1 − 1 ≤ 1.001δ1
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 43

(say). Also, from Theorem 1.3 and (87) we can bound


|γ|F N,t (Re(s∗ ) − y − |κ|) ≤ |γ|N y N |κ| F N,t (Re(s∗ ))
!
1 x ty
≤ exp 0.02y + y(log N − log ) + log N F N,t (Re(s∗ ))
2 4π 2(x − 6)
πt
!
ty 1 ty x
≤ exp 0.02y + (y + ) log(1 + ) + log F N,t (Re(s∗ )).
2(x − 6) 2 4x 4(x − 6) 4π
For 0.2 ≤ y ≤ 1, 0 ≤ t ≤ 0.2, and x ≥ X0 − 0.5 we see from Lemma 5.1(vi) that
ty x 0.2 X0 − 0.5
log ≤ log ≤ 1.86 × 10−11
4(x − 6) 4π 4(X0 − 0.5 − 6) 4π
and
ty 1 πt 0.2 1 0.2π
(y + ) log(1 + ) ≤ (1 + ) log(1 + ) ≤ 1.31 × 10−12
2(x − 6) 2 4x 2(X0 − 0.5 − 6) 2 4(X0 − 0.5)
and thus
(90) |γ|F N,t (Re(s∗ ) − y − |κ|) ≤ 1.021F N,t (Re(s∗ )).
Thus
eA + eB ≤ 1.023δ1 F N,t (Re(s∗ )).
To estimate Re(s∗ ), we use Proposition 6.6(ii) or (22), together with the inequality
4y(1 + y)
! !
t 0.2 8
1 − 3y + ≤ 1 − 3 × 0.2 + ≤ 1.2 × 10−23
2x2 x2 + 2(X 0 − 0.5) 2 (X 0 − 0.5)2
+
to obtain
t x
Re(s∗ ) ≥ 0.6001 + log
4 4π
(say). Since F N,t (σ) is non-increasing in σ, we conclude
t x
eA + eB ≤ 1.023δ1 F N,t (0.6001 + log ).
4 4π
Since  r 
x t 
N =  + ,

4π 16 
0 ≤ t ≤ 0.2, and x ≥ X0 − 0.5, it is easy to see that
x
N≤

and hence by (15)
btn
t x ≤ 1
n 4 log π
for all 1 ≤ n ≤ N. Therefore
N
t x X 1
F N,t (0.6001 + log ) ≤ 0.6001
4 4π n=1
n
and hence by the integral test
Z N+1
t x ds 1
F N,t (0.6001 + log ) ≤ = (N + 1)0.3999
4 4π 1 s0.6001 0.3999
44 D.H.J. POLYMATH

so that (by (88))


(0.2)2
16 log2 4π
x
+ 0.626 1
eA + eB ≤ 1.023 (N + 1)0.3999 .
x − 6.66 0.3999
We have
x − 6.66 1/2
N + 1 ≤ (1.001)( )

(say), and hence
0.0025 log2 4π + 0.626
x
eA + eB ≤ 0.7220
(x − 6.66)0.79995
From Lemma 5.1(vi), the right-hand side is monotone decreasing in the region x ≥ X0 − 0.5,
thus
X0 −0.5
0.0025 log2 4π + 0.626
eA + eB ≤ 0.7220
(X0 − 0.5 − 6.66)0.79995
≤ 3.444 × 10−9 .
Meanwhile, from Proposition 6.6(vi) one has
 x − 1+y x
3| log 4π + i π2 | + 3.58 1.24 × (3y + 3−y )
! !
4 t 2 x 6.92
eC,0 ≤ exp − log + 1+ + .
4π 16 4π x − 8.52 N − 0.125 x − 6.66
2
log2 4π
x
4 +π | log4π
x
2 +i π |
From Lemma 5.1(vi), the quantity (x−8.52) 2 is monotone decreasing in x in (5), hence x−8.52
is also monotone decreasing. Also the expression is monotone decreasing in y. We conclude
that
(91)
√ √
+ i π2 | + 3.58  
!− 1+0.2 −0.5
 3| log X04π
  
X0 − 0.5 4 1.24 × (3 0.6 + 3− 0.6 ) 6.92
 1 + +

eC,0 ≤ exp  
4π X − 0.5 − 8.52
0 N − 0.125
0 X − 0.5 − 6.66 
0
−3
≤ 1.249 × 10
X0 −0.5
where we have discarded the negative term − 16t log2 4π . Combining the estimates, we obtain
the claim. 
It now suffices to establish the bound
(92) | ft (x + iy)| > 1.25 × 10−3
in the following regions:
(i) When 6 × 1010 + 83952 − 0.5 ≤ x ≤ 6 × 1010 + 83952 + 0.5, 0 ≤ t ≤ 0.2, and 0.2 ≤ y ≤ 1.
(ii) When x ≥ 6 × 1010 + 83952 − 0.5, 69098 ≤ N ≤ 80000, t = 0.2, and 0.2 ≤ y ≤ 1.
(iii) When 80000 ≤ N ≤ 1.5 × 106 , t = 0.2, and 0.2 ≤ y ≤ 1.
(iv) When N ≥ 1.5 × 106 , t = 0.2, and 0.2 ≤ y ≤ 1.
We begin with claim (i). The function ft (x + iy) is holomorphic, so by Rouche’s theorem,
it suffices to show that for each time 0 ≤ t ≤ 0.2, as x + iy traverses the boundary ∂R of the
rectangle
R B {x + iy : 1010 + 83952 − 0.5 ≤ x ≤ 6 × 1010 + 83952 + 0.5; 0.2 ≤ y ≤ 1},
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 45

the function ft (x + iy) stays outside of the ball B B {z : |z| ≤ 1.25 × 10−3 }, and furthermore has
a winding number of zero around the origin.
To verify this claim numerically for a given value of t, we subdivide each edge of ∂R into
some number n of equally spaced mesh points, thus approximating ∂R by a discrete mesh x j +iy j ,
j = 1, . . . , 4n, with any two adjacent points on this mesh separated by a distance at most 1/n.
Using the techniques in Section 8.1, we evaluate ft (x j +iy j ) numerically for each such value of j.
The polygonal path connecting these points then winds around the origin with winding number
4n
1 X
arg( ft (x j+1 + iy j+1 )/ ft (x j + iy j ))
2π j=1

which can be easily computed (and verified to be zero). To pass from this polygonal path to the
true trajectory ft (∂R) of ft (x + iy) on ∂R, we again use Rouche’s theorem. If one has a derivative

bound | ∂z ft (z)| ≤ Dz on the boundary of the rectangle, the polygonal path and the true trajectory
ft (∂R) differ by a distance of at most D z
2n , and the latter will have the same winding number
around the origin (and stay outside of the ball B) as long as
Dz
| ft (x j + iy j )| > 1.25 × 10−3 +
.
2n
Furthermore, the same is true for nearby times t ≤ t0 0.2 to t, as long as one has the stronger
bound
Dz Dt |t0 − t|
(93) | ft (x j + iy j )| > 1.25 × 10−3 + +
2n 2n
and a bound of the form | ∂t∂ ft˜(z)| ≤ Dt for t ≤ t˜ ≤ 0.2 and z ∈ ∂R.
This gives the following algorithm to verify (i) for the entire range 0 ≤ t ≤ 0.2. We start with
t = 0 and obtain bounds Dt , Dz for the derivatives of ft˜(z) in the indicated ranges. Because of the
way the barrier location X was selected, we expect | ft (x + iy)| to stay well above 1 in magnitude.
We thus choose n so that D 2n ≤ 1, and evaluate ft (x j + iy j ) at all the mesh points (in particular
z

confirming that | ft (x j + iy j )| does stay well above 1). Using the minimum value of | ft (x j + iy j )|,
we can then use the condition (93) to establish the claim for times in the interval [t, t0 ) where
t0 > t is chosen so that (93) holds (or t0 = 0.2, if that is also possible); the most aggressive choice
of t0 would be one in which (93) held with equality, but in practice we can afford to take more
conservative values of t0 and still obtain good runtime performance. If t0 < 0.2, we then repeat
the process, replacing t by t0 , until the entire range 0 ≤ t ≤ 0.2 is verified.
To run this algorithm, we need bounds on Dt and Dz . This is achieved by the following
lemma, which gives bounds which are somewhat complicated but which can be easily upper
bounded numerically on ∂R:
Lemma 9.3. In the region (5), and away from the jump discontinuities of N, we have
N
∂ ft ≤ btn
X !
log n t log n
∂z +
n=1
nRe(s∗ ) 2 4(x − 6)
N
btn ny |1 + y + ix|
! !!
X t log n 3 1 t
+ |γ|N |κ|
+ log +π+ +
n=1
nRe(s∗ ) 4(x − 6) 4π x 2 4(x − 6)
46 D.H.J. POLYMATH

and
N
∂ ft ≤ btn 1 π
X !
x 2 log n
∂t log n log + log n +
n=1
nRes∗ 4 4πn 8 x−6
N
btn ny 1 π 2 log n 1 π
! !!
X x 8 x 8
+ |γ|N |κ|
log n log + log n + + + log + .
n=1
nRe(s∗ ) 4 4πn 8 x−6 4 2 x−6 4π x − 6

Proof. We begin with the first estimate. Write


1 − y + ix t 1 − y + ix
!
s∗∗ B s∗ − y + κ = + α
2 2 2
then
N N
X btn X btn
(94) ft = + γ .
n=1
n s∗ n=1
n s∗∗
One can check that s∗ , s∗∗ , γ are holomorphic functions of x + iy, hence by the Cauchy-Riemann
equations
∂ ft = ∂ ft .

∂x ∂y
By the product and chain rules, we may calculate
N N
∂ ft btn ∂s∗ btn ∂ ∂s∗∗
X X !
=− log n + γ log γ − log n .
∂x n=1
n s∗ ∂x n=1
n s∗∗ ∂x ∂x
From (17), (43) we have
∂s∗ 1 − y + ix
!
i it
= − − α0
∂x 2 4 2
!
i t
= − + O≤ .
2 4(x − 6)
Similarly we have
∂s∗∗
!
i t
= + O≤ .
∂x 2 4(x − 6)
1−y+ix
Writing s = 2 , we have from (16), (10) that
t
log γ = (α(s)2 − α(1 − s)2 ) + log M0 (s) − log M0 (1 − s)
4
and hence by (8)
∂γ it i i
= (α(s)α0 (s) + α(1 − s)α0 (1 − s)) + α(s) + α(1 − s).
∂x 4 2 2
From the triangle inequality and (43), we thus have
N N
∂ ft btn btn |α(s) + α(1 − s) − log n| t(|α(s)| + |α(1 − s)|)
! !
X log n t log n X t log n
| |≤ + +|γ| + + .
∂x n=1
nRe(s∗ ) 2 4(x − 6) n=1
nRe(s∗∗ ) 4(x − 6) 2 4(x − 6)
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 47

We have from (9) that


1 1 |1 − y + ix| π 3
|α(s)|, |α(s) − log n| ≤ log + +
2 2 4π 2 2x
x |1−y+ix|
since n ≤ N ≤ 4π ≤ 4π . Similarly
1 1 |1 + y + ix| π 3
|α(1 − s)|, |α(1 − s) − log n| ≤ log + +
2 2 4π 2 2x
and thus
|1 + y + ix| 3
|α(s) + α(1 − s)|, |α(s) + α(1 − s) − log n| ≤ log +π+ .
4π x
Writing Re(s∗∗ ) = Re(s∗ ) − y + Re(κ), we then have the first estimate.
Now we estimate the time derivative. Since
∂ 1
log btn = log2 n
∂t 4
∂ 1 1 + y − ix
s∗ = α( )
∂t 2 2
∂ 1 1 − y + ix
s∗∗ = α( )
∂t 2 2
∂ 1 − y + ix 2 1 + y − ix
!
1
log γ = α( ) − α2 ( )
∂t 4 2 2
we see from differentiating (94) that, we obtain
1+y−ix
∂ ft X btn  log2 n α( 2 )
N N  log2 n α( 1−y+ix
   
 X btn ) 1 1 − y + ix 1 + y − ix
= +γ 2
log n + (α( ) −α ( )) .
2 2

− log n −
∂t
 
n s∗  4 n s∗∗
 
n=1
2 
n=1
4 2 4 2 2
From (43), (9) we have
1 ± y + ix ix 1
α( ) = α( ) + O≤ ( )
2 2 x−6
1 x πi 4
= log + + O≤ ( )
2 4π 4 x−6
and hence (since α = α∗ )
1 ± y − ix 1 x πi 4
α( ) = log − + O≤ ( )
2 2 4π 4 x−6
1+y−ix 1−y+ix
so in particular (recalling that s+ = 2 and s− = 2 )
πi 8
α(s− ) − α(s+ ) = + O≤ ( )
2 x−6
and
x 8
α(s− ) + α(s+ ) = log + O≤ ( )
4π x−6
so that
π
! !
8 x 8
α(s− )2 − α(s+ )2 ≤ + log + .
2 x−6 4π x − 6
48 D.H.J. POLYMATH

We conclude from the triangle inequality that


N
∂ ft ≤ btn 1 π
X !
x 2 log n
∂t log n log + log n +
n=1
nRes∗ 4 4πn 8 x−6
N
btn π 2 log n 1 π
!
X 1 x 8 x 8
+ |γ| log n log + log n + + ( + )(log + )
n=1
nRe(s∗∗ ) 4 4πn 8 x−6 4 2 x−6 4π x − 6
giving the second claim. 
The next few graphs summarize the numerical output of the algorithm for the following Bar-
rier parameters: x = 6 × 1010 + 83925 ± 0.5, y = 0.2 . . . 1, t = 0 . . . 0.2.
The first step in the Barrier verification process was to precalculate a ’stored sum’ of Taylor
expansion terms that allows for fast recreation of ft (x + iy) during the execution of the algortihm.
The number of Taylor terms required was determined through an iterative process targeted to
achieve a 20 decimal accuracy. The following plot illustrates that the achieved accuracy for all
rectangular mesh points at t = 0.

Figure 12. Achieved error term in the Taylor expansion at t = 0. Target was set
at 20 decimal places accuracy.

The derivative bounds determine the number of mesh points required on each xy-rectangle
and in the t-direction. The next two plots ilustrate that these bounds have been chosen quite
conservatively:
The number of rectangle mesh points varies with t ranging from 11076 at t = 0 to 56 at
t = 0.195:
The overall winding number for the Barrier at this specific location came out at 0. The next
graph shows the winding process at t = 0. Due to the choice of the Barrier location and the
small Barrier width compared to the wave length of the x variable, little oscillation is expected
to occur in each rectangle:
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 49

Figure 13. Derivative bounds versus their actual values at all required steps of t.

Figure 14. The number of mesh points required per rectangle for each step of t.

Now we prove claim (ii).


need to detail this more
For N ∈ [69098, 80000], it was computationally verified that | fN | > 0.042, since the Lemma
bound (eqn 78) using the Euler 5 mollifier monotonically increased with N, from 0.04278 to
0.08011.

Now we prove claim (iii).


need to detail this more
We attempt here to find N dependent positive lower bounds for | ft (x + iy)| or corresponding
mollified sums through an incremental approach,
50 D.H.J. POLYMATH

Figure 15. Winding the rectangle at t = 0.

X
| fN+1 | >= | fN | − (additional terms corresponding to N+1)
with the bound for | fNmin | computed similarly to (77) or (78).

If the incremental bound goes below a positive threshold at N, we reset Nmin to N and restart
the process, generating a sawtooth pattern.

For verifyingPthe lower bounds for [Nmin , Nmax ], and with


βn,N = λd btn/d ,
1≤d≤D:d|n:n≤dN
αn,N = λd atn/d ,
P
1≤d≤D:d|n:n≤dN
1−|γ| γNmax
g(n, N, Nmax ) = max( 1+|γ|NNmax |βn,N + αn,N |, |βn,N − αn,N |, |βn,N − | γN |αn,N |),
max

For the Triangle inequality bound,

DN N
1 X min
|βn,Nmin | + |αn,Nmin | Xmin
btn (n|κ| − 1)
| fNmin | ≥ PD (1 − |γNmin | − ) − |γ |
nσNmin nσNmin −y
|λd | Nmin
d=1 dσNmin n=2 n=1

1 X dN+d
X |βn,N+1 | + |αn,N+1 | btN+1 ((N + 1)|κ| − 1)
| fN+1 | ≥ | fN | − PD ( ) − |γ N+1 |
|λd | nσN+1 (N + 1)σN+1 −y
d=1 dσN+1 1≤d≤D:d|D n=dN+1

For the Lemma inequality bound,

DN N
1 X min
g(n, Nmin , Nmax ) Xmin
btn (n|κ| − 1)
| fNmin | ≥ PD (1 − |γNmin | − ) − |γ |
nσNmin nσNmin −y
|λd | Nmin
d=1 dσNmin n=2 n=1
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 51

1 X dN+d
Xg(n, Nmin , Nmax ) btN+1 ((N + 1)|κ| − 1)
| fN+1 | >= | fN | − PD ( ) − |γN+1 |
|λd | nσN+1 (N + 1)σN+1 −y
d=1 dσN+1 1≤d≤D:d|D n=dN+1

Note: The incremental bounds work since,


α α
1) βn,N = βn,N+1 and γn,NN
= γn,N+1
N+1
when n < [dN, dN + d] for d : d|D
2) g(n, N, Nmax ) can be seen to have the following properties with n fixed and N varying,
resulting in it being constant or decreasing with increasing N,
•g(n, N, Nmax ) = |βn,N − αn,N | when βn,N αn,N ≤ 0
1−|γ| γ
•g(n, N, Nmax ) = max( 1+|γ|NNmax |βn,N +αn,N |, |βn,N −| Nγmax
N
|αn,N |)) when βn,N αn,N > 0 and |βn,N | ≥
max
|αn,N |
1−|γ| γ
•g(n, N, Nmax ) = max( 1+|γ|NNmax |βn,N + αn,N |, |βn,N − αn,N |, |βn,N − | Nγmax
N
|αn,N |)) when βn,N αn,N >
max
0 and |αn,N | > |βn,N |

Now, with Nmax = 1.5million, y = 0.2, t = 0.2, Euler 5 mollifier, and threshold = 0.03
and starting with Nmin = 80000, and resetting Nmin whenever the incremental lemma bound
goes below 0.03,
it was computationally verified that | fN | ≥ 0.03 for N ∈ [80000,1.5 million]

..

..

..
...
Finally, we prove claim (iv). In this regime we have
πt
x ≥ xN B 4πN 2 − ;
4
in particular,
x ≥ x1.5×106 ≥ 2.82 × 1013 .
In view of (79), it suffices to show that
N N
X b0.2 X b0.2
(95) n
+ |γ| n
< 2 − 1.25 × 10−3
n=1
nσ n=1
nσ−0.2−|κ|

where σ B Res∗ . Our main tool here will be Lemma 8.1 (with N0 = 69098). First observe from
(22) that
0.2 × 0.2
|κ| ≤ ≤ 7.08 ∗ 10−16
2(2.82 × 1013 − 6)
while from (20) one has
 x −0.1
N
|γ| ≤ 1.005

52 D.H.J. POLYMATH

and from (21) one has


 
0.2 xN 0.2 0.96
σ ≥ 0.6 + 0.4 +
 
log − 2 2

4 4π 2x x1.5∗106
1.5∗106 +
!  
t 0.2 0.96
≥ 0.6 + 0.1 log N + 0.05 log 1 − 0.4 +
 
6 2
− 2 2

16(1.5 ∗ 10 ) 2x1.5∗106 x1.5∗106
+
≥ 0.6 + 0.1 log N − 5.1 × 10−29 .
We can then apply Lemma 8.1 twice to bound the left-hand side of (95) by A + B, where
N0 0.2 N0 0.2
 x −0.1 X
X b N b
AB n
+ 1.005 n
,
nσ nσ
0 00
n=1
4π n=1
0 0
 x −0.1 00 00 N
N
N0 , N
B B (max(N01−σ b0.2 bN ) + 1.005 N0 , N
1−σ 0.2
max(N01−σ b0.2 1−σ 0.2
bN )) log
4π N0
σ := 0.6 + 0.1 log N − 5.1 × 10
0 −29

σ00 := 0.4 + 0.1 log N − 7.09 × 10−16 .


The quantity A is decreasing in N, so we may bound it by its value at N = 1.5 × 106 . Performing
the sum numerically, we obtain
A ≤ 1.88.
Finally, the quantity B can also be seen to be decreasing in N in the range N ≥ 1.5 × 106 , and
obeys the bound
B ≤ 0.075.
The claim follows.

Figure 16. Plot of B vs N


UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 53

10. Asymptotic results


In this section we use the effective estimates from Theorem 1.3 to obtain asymptotic informa-
tion about the function Ht , which improves (and makes more effective) the results of Ki, Kim,
and Lee [9], by establishing Theorem 1.5.
We begin with an asymptotic
Proposition 10.1. Let 0 < t ≤ 1/2, x ≥ 200, and −10 ≤ y ≤ 10.
(i) If x ≥ exp( Ct ) for a sufficiently large absolute constant C, then
1 + y − ix 1 − y + ix
! !
Ht (x + iy) = (1 + O(x−ct ))Mt + (1 + O(x−ct ))Mt
2 2
for an absolute constant c > 0, where Mt is defined in (10).
(ii) If instead we have 3 ≤ y ≤ 4 and x ≥ C for a sufficiently large absolute constant C, then
1 + y − ix
!
Ht (x + iy) = (1 + O≤ (0.7))Mt .
2
(iii) If x = x0 + O(1) for some x0 ≥ 200, then
1 + y − ix ix0
Ht (x + iy) = O(x0O(1) |Mt ( )|) = O(x0O(1) |Mt ( )|).
2 2
Proof. We begin with (i). Since Ht = Ht and Mt = Mt , we may assume without loss of
∗ ∗

generality that y ≥ 0. Using (16), (11) we may write the desired estimate as
Ht (x + iy)
= 1 + O(x−ct ) + γ.
B0 (x + iy)
We apply Theorem 1.3. (Strictly speaking, the estimates there required y ≤ 1 rather than y ≤ 10;
however, as remarked at the beginning of Section 6, all the estimates in that section would
continue to hold under this weaker hypothesis if one adjusted all the numerical constants appro-
priately.) This gives
N N
Ht (x + iy) X btn X
y bn
t
= + γ + O≤ eA + eB + eC,0

(96) n
B0 (x + iy) n=1 n s∗
n=1
n s∗ +κ

where
γ = O(x−y/2 )
κ = O(x−1 )
1+y t x
Re(s∗ ) ≥ + log − O(x−2 )
2 2 4π
 N

2 
 X 1−y t x −1 log x
eA = O  x−y/2 btn n− 2 − 2 log 4π −O(x )


n=1
x 
 N 
2 
1+y t
+O(x −1 ) log x 
X x
eB = O  bn n 2 2 4π
t − − log
 
n=1
x 
 1+y 
eC,0 = O x− 4
54 D.H.J. POLYMATH

−1
Since N = O(x1/2 ), we have x−y/2 ny = O(1) and nO(x ) = O(1) for all 1 ≤ n ≤ N. We conclude
that
 N

Ht (x + iy)  log2 x X bt 1+y 
= 1 + γ + O  + n
+ x− 4 

B0 (x + iy) x 1+y t
+ log 4π
n=2 n 2 2
x

so it will suffice (for c small enough) to show that


N
X btn
1+y t
= O(x−ct ).
2 +2
x
log
n=2 n 4π

By (15) we can write the left-hand side as


N
X 1
1+y t
= O(x−ct ).
2 +2
x√
log
n=2 n 4π n

For 2 ≤ n ≤ N, we have
1+y t x
+ log √ ≥ ct log x
2 2 4π n
for some absolute constant c > 0. By the integral test, the left-hand side is then bounded by
Z ∞
1 1
ct log x
+ ct log x
du
2 2 u

which, for x ≥ exp(C/t) and C large, is bounded by O(2−ct log x ). The claim then follows after
adjusting c appropriately.
Now we prove (ii). As before we have the expansion (96). We have
N
 N

t t
y bn b
X  X 
γ n s +κ = O  x −y/2
1−y
n

n ∗ − 2 + 2t log 4π
x 
n=1 n=1 n
 N

 X y−1 
= O  x−y/2

n 2 
n=1
− y−1
= O(x 4 );
2
similar arguments give eA = O( logy−1x ), while
x 4

 N

 log2 x X 1+y t x 

eB = O   t − −
bn n 2 2 4π  log
x n=1
 N

 log2 x X 
= O  n−2 
x n=1
log2 x
!
=O .
x
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 55

We conclude that
N
Ht (x + yi) X btn y−1
= + O(x− 4 )
B0 (x + yi) n=1 n s∗

 N 
X 1+y t x −1
 y−1
= 1 + O≤  n− 2 − 2 log 4π −O(x )  + O(x 4 )
n=2
 N 
X 
= 1 + O≤  n  + O(x−1/2 )
−2

n=2
π2
!
= 1 + O≤ − 1 + O(x−1/2 )
6
= 1 + O≤ (0.7)
as claimed, if x ≥ C for C large enough.
Finally we prove (iii). Again our starting point is (96). The right-hand side can be bounded
crudely by O(xO(1) ) = O(x0O(1) ), hence
1 + y + ix
Ht (x + iy) = O(x0O(1) |Mt ( )|).
2
However, from (10), (6), (9) it is not hard to see that the log-derivative of Mt (s) is of size
O(log x0 ) in the region s = ix20 + O(1). Thus
1 + y + ix ix0
|Mt ( )| = O(x0O(1) |Mt ( )|),
2 2
giving the claim. 
To understand the behavior of Mt (x + iy) we make the following simple observations:
Lemma 10.2. Let 0 < t ≤ 1/2, let x∗ > 0 be sufficiently large, and let x + iy = x∗ + O(1). Then
1 + y + ix 1 + ix∗ x∗ πi
! !!
1 log x∗
Mt ( ) = Mt ( ) exp (i(x − x∗ ) + y) log + +O .
2 2 4 4π 8 x∗
 
Also, there is a continuous branch of argMt 1+ix 2

for all large real x∗ such that
1 + ix∗
!
tπ x∗ 7π x∗ x∗ x∗ log x∗
argMt = log + + log − + O( ).
2 16 4π 8 4 4π 4 x∗
Proof. By (10), (8), the log-derivative of Mt is given by
Mt0 t
(97) = α + αα0 .
Mt 2
For s = ix∗
2 + O(1), we have from (9) that
x∗ πi
!
1 1
(98) α(s) = log + +O
2 4π 4 x∗
and from this and (43) we conclude that
Mt0 (s) 1 x∗ πi
!
log x∗
= log + +O
Mt (s) 2 4π 4 x∗
56 D.H.J. POLYMATH

whenever s = ix2∗ + O(1). The first claim then follows by applying the fundamental theorem of
calculus to a branch of log Mt .
For the second claim, we calculate
!2
1 + ix∗ 1 + ix∗ −1 + ix∗ 1 + ix∗ 1 + ix∗
! !
t x∗
argMt = Imα + π − log π + Im log −
2 4 2 4 4 4 4
t π −1 + ix∗
! !!!
x∗ log x∗ x∗ x∗ iπ i 1 x∗
= log + O( ) + π − log π + Im log + − +O 2 −
4 4 4π x∗ 4 4 4 2 x∗ x∗ 4
x∗ π
!
tπ x∗ x∗ x∗ log x∗
= log + π + log π + log − + O
16 4π 4 4 4 8 x∗
!
tπ x∗ 7π x∗ x∗ x∗ log x∗
= log + + log − +O
16 4π 8 4 4π 4 x∗
as desired. 
Now we can prove Theorem 1.5. We begin with (ii). Let n ≥ exp( Ct ), and suppose that
x + iy = xn + O(1). By Proposition 10.1(i) and Lemma 10.2 we have
1 + ixn xn πi
! ! !
1
Ht (x + iy) = Mt exp (−i(x − xn ) + y) log − + O(xn )
−ct
2 4 4π 8
(99)
1 + ixn xn πi
! ! !
1
+ Mt exp (i(x − xn ) − y) log + + O(xn−ct ) .
2 4 4π 8
From Lemma 10.2 and (26) one has
1 + ixn π
! !
log xn
argMt =− +O mod π
2 2 xn
and hence
1 + ixn 1 + ixn
! ! !
log xn
(100) Mt = − exp O( ) Mt .
2 xn 2
 
If we now make the further assumption y = O log1 xn , we can thus simplify the above approxi-
mation as
(101)
1 + ixn −π(x−xn )/8
! !
1 xn
Ht (x + iy) = −Mt e exp (−i(x − xn ) + y) log + O(|y| log xn + xn )−ct
2 4 4π
1 + ixn −π(x−xn )/8
! !
1 xn
+ Mt e exp (i(x − xn ) − y) log + O(|y| log xn + xn )
−ct
2 4 4π
1 + ixn −π(x−xn )/8   x + iy − xn
!
xn  
= 2iMt e sin log + O(|y| log xn + xn−ct ) .
2 4 4π
In particular, if x + iy traverses the circle {xn + logc n eiθ : 0 ≤ θ ≤ 2π} once anti-clockwise and c
is small enough, the quantity Ht (x + iy) will wind exactly once around the origin, and hence by
the argument principle there is precisely one zero of Ht inside this circle. As the zeroes of Ht
are symmetric around the real axis, this zero must be real. This proves (ii).
Now we prove (i). Suppose that Ht (x + iy) = 0 and x ≥ exp( Ct ). We can assume |y| ≤ 1 since
it is known (e.g., from [4, Theorem 13]) that there are no zeroes with |y| > 1.
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 57
 
Let n be a natural number that minimises |x − xn |, then x = xn + O log1 xn since the derivative
of the left-hand side of (26) in xn is comparable to log xn . From (99) we have
1 + ixn xn πi
! ! !
1
0 = Mt exp (−i(x − xn ) + y) log − + O(xn )
−ct
2 4 4π 8
1 + ixn xn πi
! !
1
+ Mt ( ) exp (i(x − xn ) − y) log + + O(xn−ct ) .
2 4 4π 8
Thus both summands on the right-hand side have the same magnitude, which on taking loga-
rithms and canceling like terms implies that
1 xn 1 xn
y log + O(xn−ct ) = −y log + O(xn−ct )
4 4π 4 4π
 −ct 
x
and hence y = O logn xn . We can now apply (101) to conclude that
 x + iy − x xn 
n
sin log + O(xn−ct ) = 0
4 4π
 −ct 
x
which (when combined with the hypothesis that |x − xn | is minimal) forces x − xn = O logn xn .
This gives the claim.
Next, we prove (iii). In view of parts (i) and (ii), and adjusting C if necessary, we may assume
that X takes the form X = xn + logc xn for some n ≥ exp( Ct ). By the argument principle, Nt (X) is
2π times the variation in the argument of Ht on the boundary of the rectangle {x + iy :
equal to −1
0 ≤ x ≤ X; −3 ≤ y ≤ 3} traversed clockwise, since there are no zeroes with imaginary part of
magnitude greater than one. By compactness, the variation on the left edge {iy : −3 ≤ y ≤ 3} is
O(1), and similarly for any fixed portion {x+3i : 0 ≤ y ≤ C} of the upper edge. From Proposition
10.1 (and (99)), we see that the variation of Ht (x + iy)/Mt ( 1+y−ix 2 ) on the remaining upper edge
{x + 3i : C ≤ x ≤ X} and on the top half {X + iy : 0 ≤ y ≤ 3} of the right edge are both equal to
O(1). Since Ht = Ht∗ , the variation on the lower half of the rectangle is equal to that of the upper
half. We thus conclude that
!
1 1 − iX
Nt (X) = − argMt + O(1)
π 2
 
where we use a continuous branch of the argument of Mt 1−iX 2 that is bounded at 3i. The claim
now follows from Lemma 10.2.
Finally, we prove (iv). From (4) and the rapid decrease of Φ it is easy to verify that the entire
function Ht has order 1, while from the positivity of Φ we see that Ht has no zero at the origin.
Thus by the Hadamard factorization theorem we have
!
Y z z
Ht (z) = exp(a + bz) 1− exp( )
n
zn zn
for some complex numbers a, b, where zn are the zeroes of Ht counting multiplicity; using the
functional equation Ht (z) = Ht (−z) we can index the zeros as zn = (zn )n∈Z\{0} with z−n = −zn ,
and conclude that b = 0, and
z2
Y !
Ht (z) = exp(a) 1− 2 .
n>0
zn
58 D.H.J. POLYMATH

Taking logarithmic derivatives, we conclude that


Ht0 (z) X
!
1 1
(102) = + .
Ht (z) n>0 z − zn z + zn
Setting z = X+4i, we see from Proposition 10.1 and the generalized Cauchy integral formula that
 1+y−ix
the logarithmic derivative of Ht (x + iy)/Mt 2 is equal to O(1) at X + 4i for all sufficiently
large X, and hence for all X by symmetry and compactness. On the other hand, from Stirling’s
formula (or the logarithmic
 1+y−ix growth of the digamma function) one easily verifies that the logarith-
H 0 (X+4i)

mic derivative of Mt 2 is equal to O(log(2 + X)) at X + 4i. Hence Htt (X+4i) = O(log(2 + X)).
Taking imaginary parts, we conclude that
X (4 − yn ) (4 − yn )
+ = O(log(2 + X))
n>0
(X − xn ) + (4 − yn )
2 2 (X + xn )2 + (4 + yn )2
where we write zn = xn + iyn ; equivalently one has
X (4 − yn )
= O(log(2 + X))
n
(X − xn )2 + (4 − yn )2
where the sum now ranges over all zeroes, including any at the origin. Since |yn | ≤ 1, every
zero in [X, X + 1] makes a contribution of at least 100 1
(say). As the summands are all positive,
the first part of claim (iv) follows. To prove the second part, we may assume by compactness
that x ≥ C. Repeating the proof of (iii), and reduce to showing that the variation of argHt on
the short vertical interval {X + iy : 0 ≤ y ≤ 3} is O(log X). If we let θ be a phase such that
eiθ Ht (X + 3i) is real and positive, we see that this variation is at most π(m + 1), where m is the
number of zeroes of Reeiθ Ht (X + yi) for 0 ≤ y ≤ 3, since every increment of π in argeiθ Ht must
be accompanied by at least one such zero. As Ht = Ht∗ , this is also the number of zeroes of
eiθ Ht (X + yi) + e−iθ Ht (2X − (X + yi)). On the other hand, from Proposition 10.1(ii), (iii) and
Jensen’s formula we see that the number of such zeroes is O(log X), and the claim follows.
Remark 10.3. Theorem 1.5 gives good control on Ht (x + iy) whenever x ≥ exp(C/t). As a
consequence (and assuming for sake of argument that the Riemann hypothesis holds), then for
any Λ0 > 0, the bound Λ ≤ Λ0 should be numerically verifiable in time O(exp(O(1/Λ0 ))), by
applying the arguments of previous sections with t and y set equal to small multiples of Λ0 . We
leave the details to the interested reader.
Remark 10.4. Our discussion here will be informal. In view of the results of [8], it is expected
that the zeroes z j (t) of Ht (x + iy) should evolve according to the system of ordinary differential
equations
0
X 1
∂t zk (t) = 2
z (t) − z j (t)
j,k k
where the sum is evaluated in a suitable principal value sense, and one avoids those times where
the zero zk (t) fails to be simple; see [8, Lemma 2.4] for a verification of this in the regime t > Λ.
In view of the Riemann-von Mangoldt formula (as well as variants such Corollary 1.5, it is
expected that the number of zeroes in any region of the form {x + iy : x + iy = x∗ + O(1)} for
large x∗ should be of the order of log x∗ . As a consequence, we expect a typical zero zk (t) to
move with speed O(log |zk (t)|), although one may occasionally move much faster than this if two
zeroes are exceptionally close together, or less than this if the zeroes are close to being evenly
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 59

spaced. As a consequence, if the Riemann hypothesis fails and there is a zero x + iy of H0 with
y comparable to 1, it should take time comparable to log1 x for this zero to move towards the real
axis, leading to the heuristic lower bound Λ  log1 x . Thus, in order to obtain an upper bound
Λ ≤ Λ0 , it will probably be necessary to verify that there are no zeroes x + iy of H0 with y
comparable to 1 and |x| ≤ c log Λ1 for some small absolute constant c > 0. This suggests that the
time complexity bound in Remark 10.3 is likely to be best possible (unless one is able to prove
the Riemann hypothesis, of course).
In [8, Lemma 2.1] it is also shown that the velocity of a given zero z(t) is given by the formula
Ht00 (z(t))
∂t z(t) =
Ht0 (z(t))
assuming that the zero is simple. By using the asymptotics in Proposition 10.1 and Corollary
1.5 together with the generalized Cauchy integral formula to then obtain asymptotics for Ht0 and
Ht00 , it is possible to show that for the zeroes x(t) that are real and larger than exp(C/t), and
move leftwards with velocity
π
∂t x(t) = − + O(x−ct );
4
we leave the details to the interested reader.

Figure 17. Real zeros moving up leftwards and getting ‘solidified’.


60 D.H.J. POLYMATH

Figure 18. Actual trajectories of some real and complex zeros.

11. Further numerical results


By Theorem 1.5, one can verify the second hypothesis of Theorem 1.2 when X ≥ exp(C/t0 )
for a large constant C. If we ignore for sake of discussion the third hypothesis of Theorem 1.2
(which turns out to be relatively easy to verify numerically in practice), this suggests that one
can obtain a bound of the form Λ ≤ O(t0 ) provided that one can verify the Riemann hypothesis
up to a height exp(C/t0 ). In other words, if one has numerically verified theRiemann
 hypothesis
up to a large height T , this should soon lead to a bound of the form Λ ≤ O log1 T .
Aside from improving the implied constant in this bound, it does not seem easy to improve
this sort of implication without a major breakthrough on the Riemann hypothesis (such as a
massive expansion of the known zero-free regions for the zeta function inside the critical strip).
We shall justify this claim heuristically as follows. Suppose that there was a counterexample to
the Riemann hypothesis at a large height T , so that H0 (2T + iy) = 0 for some positive y, which
for this discussion we will take to be comparable to 1. The Riemann von Mangoldt formula
indicates that the number of zeroes of H0 within a bounded distance of this zero should be
comparable to log T ; the majority of these zeroes should obey the Riemann hypothesis and thus
stay at roughly unit distance from our initial zero 2T + iy. Proposition 3.1 then suggests that
as time t advances, this zero should move at speed comparable to log T . Thus one should not
expect this zero to reach the real axis until a time comparable to log1 T . This heuristic analysis
 
therefore indicates that it is unlikely that one can significantly improve the bound Λ ≤ O log1 T
without being able to exclude significant violations of the Riemann hypothesis at height T .
UPPER BOUND FOR DE BRUIJN-NEWMAN CONSTANT 61

In this section we collect some numerical results verifying the second two hypotheses of
Theorem 1.2 for larger values of X, and smaller values of t0 , than were considered in Section
9. This leads to improvements to the bound Λ ≤ 0.22 conditional on the assumption that the
Riemann Hypothesis can be numerically verified beyond the height T ≈ 1.2 × 1011 used in
Section 9. Our conclusions may be summarised as follows:

Table 1. Conditional Λ Results

X t0 y0 Λ Winding Number Na Moll2 Bound


2 × 1012 + 129093 0.198 0.15492 0.21 0 398942 0.0341
5 × 1012 + 194858 0.186 0.16733 0.20 0 630783 0.0376
2 × 1013 + 131252 0.180 0.14142 0.19 0 1261566 0.0349
6 × 1013 + 123375 0.168 0.15492 0.18 0 2185096 0.0377
3 × 1014 + 188911 0.161 0.13416 0.17 0 4886025 0.0369
2 × 1015 + 122014 0.153 0.11832 0.16 0 12615662 0.0532
7 × 1015 + 68886 0.139 0.14832 0.15 0 23601743 0.0350
6 × 1016 + 156984 0.132 0.12649 0.14 0 69098829 0.0307
6 × 1017 + 88525 0.122 0.12649 0.13 0 218509686 0.0347
9 × 1018 + 35785 0.113 0.11832 0.12 0 846284375 0.0318
2 × 1020 + 66447 0.102 0.12649 0.11 0 3989422804 0.0305
9 × 1021 + 70686 0.093 0.11832 0.1 0 26761861742 0.0321

The computations for X = 2 × 1020 + 66447 and X = 9 × 1021 + 70686 in the above table were
massive, and performed using a Boinc based grid computing setup, in which a few hundred vol-
unteers participated. Their contributions can be tracked at http://anthgrid.com/dbnupperbound
...

References
[1] J. Arias de Reyna, High-precision computation of Riemann’s zeta function by the Riemann-Siegel asymptotic
formula, I, Mathematics of Computation, 80 (2011), 995–1009.
[2] J. Arias de Reyna, J. Van de lune, On the exact location of the non-trivial zeroes of Riemann’s zeta function,
Acta Arith. 163 (2014), 215–245.
[3] W. G. C. Boyd, Gamma Function Asymptotics by an Extension of the Method of Steepest Descents, Proceedings:
Mathematical and Physical Sciences, Vol. 447, No. 1931 (Dec. 8, 1994), 609–630.
[4] N. C. de Bruijn, The roots of trigonometric integrals, Duke J. Math. 17 (1950), 197–226.
[5] G. Csordas, T. S. Norfolk, R. S. Varga, A lower bound for the de Bruijn-Newman constant Λ, Numer. Math. 52
(1988), 483–497.
[6] G. Csordas, A. M. Odlyzko, W. Smith, R. S. Varga, A new Lehmer pair of zeros and a new lower bound for the
De Bruijn-Newman constant Lambda, Electronic Transactions on Numerical Analysis. 1 (1993), 104–111.
[7] G. Csordas, A. Ruttan, R.S. Varga, The Laguerre inequalities with applications to a problem associated with
the Riemann hypothesis, Numer. Algorithms, 1 (1991), 305–329.
[8] G. Csordas, W. Smith, R. S. Varga, Lehmer pairs of zeros, the de Bruijn-Newman constant Λ, and the Riemann
hypothesis, Constr. Approx. 10 (1994), no. 1, 107–129.
[9] H. Ki, Y. O. Kim, and J. Lee, On the de Bruijn-Newman constant, Advances in Mathematics, 22 (2009), 281–
306.
[10] D. H. Lehmer, On the roots of the Riemann zeta-function, Acta Math. 95 (1956), 291–298.
62 D.H.J. POLYMATH

[11] H. L. Montgomer y, The pair correlation of zeros of the zeta function, Analytic number theory (Proc. Sympos.
Pure Math., Vol. XXIV, St. Louis Univ., St. Louis, Mo., 1972), 181–193. Amer. Math. Soc., Providence, R.I.,
1973.
[12] H. L. Montgomery, R. C. Vaughan, Multiplicative number theory. I. Classical theory. Cambridge Studies in
Advanced Mathematics, 97. Cambridge University Press, Cambridge, 2007.
[13] C. M. Newman, Fourier transforms with only real zeroes, Proc. Amer. Math. Soc. 61 (1976), 246–251.
[14] T. S. Norfolk, A. Ruttan, R. S. Varga, A lower bound for the de Bruijn-Newman constant Λ II., in A. A. Gonchar
and E. B. Saff, editors, Progress in Approximation Theory, 403–418. Springer-Verlag, 1992.
[15] A. M. Odlyzko, An improved bound for the de Bruijn-Newman constant, Numerical Algorithms 25 (2000),
293–303.
[16] T. K. Petersen, Eulerian numbers. With a foreword by Richard Stanley. Birkhäuser Advanced Texts: Basler
Lehrbücher. Birkhäuser/Springer, New York, 2015.
[17] D. J. Platt, Isolating some non-trivial zeros of zeta, Math. Comp. 86 (2017), 2449–2467.
[18] G. Polya, Über trigonometrische Integrale mit nur reelen Nullstellen, J. Reine Angew. Math. 58 (1927), 6–18.
[19] D. H. J. Polymath, github.com/km-git-acc/dbn upper bound
[20] B. Rodgers, T. Tao, The De Bruijn-Newman constant is nonnegative, preprint. arXiv:1801.05914
[21] Y. Saouter, X. Gourdon, P. Demichel, An improved lower bound for the de Bruijn-Newman constant, Mathe-
matics of Computation. 80 (2011), 2281–2287.
[22] J. Stopple, Notes on Low discriminants and the generalized Newman conjecture, Funct. Approx. Comment.
Math., vol. 51, no. 1 (2014), 23–41.
[23] J. Stopple, Lehmer pairs revisited, Exp. Math. 26 (2017), no. 1, 45–53.
[24] H. J. J. te Riele, A new lower bound for the de Bruijn-Newman constant, Numer. Math., 58 (1991), 661–667.
[25] E. C. Titchmarsh, The Theory of the Riemann Zeta-function, Second ed. (revised by D. R. Heath-Brown),
Oxford University Press, Oxford, 1986.

http://michaelnielsen.org/polymath1/index.php

Vous aimerez peut-être aussi