11.07.2015 Views

ZERO FREE REGIONS FOR DIRICHLET SERIES 1. Introduction In ...

ZERO FREE REGIONS FOR DIRICHLET SERIES 1. Introduction In ...

ZERO FREE REGIONS FOR DIRICHLET SERIES 1. Introduction In ...

SHOW MORE
SHOW LESS
  • No tags were found...

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 5)dtt that vanish almost ev-1−2σthe subspace of functions in L ( 2 (0, +∞),erywhere on (1, +∞).We defineψ(u) = res (L(s) ˆϕ(s)u s , s = 1) − ∑ ( na n ϕu)n σ 0 . The following gives a criterion for this property.Theorem 2.<strong>1.</strong> Let r > σ 0 . If m L ≥ 1, we assume furthermore thatr ≠ <strong>1.</strong> The following are equivalent:(i) The function ψ belongs to L 2 du((1, +∞), ).u 1+2r(ii) The function t ↦−→ L(r + it) ˆϕ(r + it) belongs to L 2 (R).We will see in Corollary 3.4 that the conditions (i) and (ii) above aresatisfied whenever r > <strong>1.</strong> Furthermore, Theorem 2.1 is a generalizationof [dR07a, Proposition 3.3] when ϕ = χ (0,1) and r = 1/2.We see that the second condition of the theorem above dependson some growth estimates of L in vertical strip which is, in its turn,linked with questions related to the convexity bound and to the Lindelöfhypothesis. <strong>In</strong> particular, if L is a function in the Selberg class, thenone can prove that L satisfies the generalized Lindelöf hypothesis if andonly if for every k ∈ N, we have(2.1) t ↦−→ Lk ( 1 + it) 21+ it ∈ L 2 (R).2Note that with the special choice of ϕ = χ (0,1) , then ˆϕ(s) = 1/s and thecondition (ii) above for r = 1/2 means exactly that (2.1) is satisfiedfor k = <strong>1.</strong> Moreover, in [dR07a] using the functional equation, it isshown that the condition ψ ∈ L 2 ((1, +∞), du ) (in the case ϕ = χu 2 (0,1) )is necessary for the generalized Riemann Hypothesis for L-functions inthe Selberg class to hold. <strong>In</strong> [dR07b], it is also shown that the conditionon ψ (still with ϕ = χ (0,1) ) is satisfied for L-functions in the Selbergclass of degree less than 4.Fix an integer m ≥ 0 and let W = ⋃ n≥1 (0, 1]n . We say that α ∈ Wis of length n if α belongs to (0, 1] n . For each α in W , its length isdenoted by l(α). Now let α ∈ W and c ∈ C l(α) , we say that A = (α, c)is an m-admissible sequence if(2.2)l(α)∑c j α j (log α j ) k = 0 for all 0 ≤ k ≤ m − <strong>1.</strong>j=1


8 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTAn other consequence of Theorem 2.2 is, in a way, a Beurling-Nymancriterion for Dirichlet L-functions. More precisely, let χ be a Dirichletcharacter with conductor q and L(χ, s) its L-function. Then, for 1/2 ≤r < 1, we define d r byd r = minl,c,α⎛⎝∫ 10∣ t1−r − t rl∑j=1c j∑n 0 and some 1/2 ≤ r ≤ 1, then L(χ, σ) does notvanish in the real-interval σ > 1 − C/ log q.If we could take C independent of χ in (2.4), then Theorem 2.5 wouldsolve the Siegel zero problem.<strong>In</strong> order to obtain the criterion for the Siegel zero problem, then weconsider all Dirichlet characters χ and an absolute constant C independantof χ.The next section is devoted to the proof of Theorem 2.1 and tothe study of the function ψ. Section 4 will focus on the admissiblesequences and the functions f A,r . Theorem 2.2 and Corollary 2.3 willbe proven in Section 5 and Theorem 2.4 will be proven in Section 6.Some explicit examples will be studied in Section 7 in which we willprove Theorem 2.5.2dtt⎞⎠123. The function ψ and proof of theorem 2.1We define the functions ψ 1 and ψ 2 byψ 1 (u) = res (L(s) ˆϕ(s)u s , s = 1) (u ∈ R + ),ψ 2 (u) = ∑ ( na n ϕ (u ∈ R + ),u)n


write<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 9(3.1) L(s) ˆϕ(s) =with H analytic in Π σ0 .Lemma 3.<strong>1.</strong> We have(3.2) H(s) =∫ 1where φ(t) = ψ(1/t)χ (0,1) (t).0∑m Lk=1p −k(s − 1) k − H(s),( ) 1ψ t s−1 dt =t ˆφ(s), R(s) > 1,Proof of Lemma 3.<strong>1.</strong> On the one hand, we have(3.3) L(s) ˆϕ(s) =∫ +∞On the other hand, the equality(3.4)∫ +∞11ψ 2 (u)u −s−1 du, R(s) > <strong>1.</strong>ψ 1 (u)u −s−1 du =∑m Lk=1p −k(s − 1) kfollows by linearity using the classical identities( )us(3.5) res(s − 1) , s = 1 u(log u)k−1= k (k − 1)!and1(k − 1)!∫ +∞1(log u) k−1 u −s du =1(s − 1) k .Equations (3.3) and (3.4) imply now the expected result.□Remark 3.2. From (3.1) and (3.5), we easily see that we can writeψ 1 (t) = tP (log t) where P is a polynomial of degree < m L (P ≡ 0 ifm L = 0). More precisely we have∑m Lp −kψ 1 (t) = t(k − 1)! (log t)k−1 .k=1Lemma 3.3. The function s ↦→ H(s) is of finite order on Π r . Moreover,for all σ > 1, the function t ↦→ H(σ + it) belongs to L 2 (R) and∫( ) 1|H(σ + it)| 2 dt = Ofor σ → +∞.Rσ 1−2σ 1


10 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTProof of Lemma 3.3. The function s ↦→ H(s) is bounded on a neighborhoodV 1 of s = <strong>1.</strong> The functions s ↦→ ∑ p −jjand (by assumption)s ↦→ L(s) are of finite order on Π σ0 V 1 . Moreover, since(s−1) jϕ(x) = O(x −σ 0(1 − x) −σ 1) and r > σ 0 , it is clear that s ↦→ ˆϕ(s) isbounded on the closure of Π r . Hence we can conclude that H is offinite order on Π r , which proves the first part of the lemma.Let σ > 1 be a fixed real number. We easily check that, for every j ≥ 1,the function t ↦−→ (σ − 1 + it) −j belongs to L 2 (R) and we have∫∣ 1 ∣∣∣2 ( ) 1(3.6) ∣dt = O , σ → +∞.(σ + it − 1) j σ 2j−1RNow, Plancherel’s formula and the estimate ϕ(x) = O(x −σ 0(1 − x) −σ 1)yield∫∫ 1| ˆϕ(σ + it)| 2 dt = 2π |ϕ(x)| 2 x 2σ dxR0x(∫ 1)= O x 2σ−2σ0−1 (1 − x) −2σ 1dx .Then, the classical identityβ(A, 1 − 2σ 1 ) :=0∫ 1and Stirling’s formula give∫ 1(x A−1 (1 − x) 1−2σ 1 1dx = O00x A−1 (1 − x) 1−2σ 1dx = Γ(A)Γ(1 − 2σ 1)Γ(A + 1 − 2σ 1 )A 1−2σ 1), A → +∞,and we can conclude (recall that s ↦→ L(s) is bounded in Π σ ).Now we can prove Theorem 2.<strong>1.</strong>Assume that (i) is satisfied. Then the function φ(t) = ψ ( )1 χ(0,1) (t)tbelongs to L 2 dt∗((0, 1), ) and thus the function G := ˆφ belongs tot 1−2rH 2 (Π r ). The analytic continuation principle implies that the equality(3.2) is satisfied for every s ∈ Π r , that is H(s) = ˆφ(s) = G(s), s ∈ Π r .Since G ∈ H 2 (Π r ), we know that the function G ∗ , defined byG ∗ (t) := lim G(σ + it),σ→> rexists almost everywhere on R and belongs to L 2 (R). ButG ∗ (t) = lim H(σ + it) = H(r + it),σ→> rbecause H is continuous on the closed half-plane R(s) ≥ r. Thereforet ↦−→ H(r + it) belongs to L 2 (R). It remains to notice that for every□


<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 111 ≤ k ≤ m L , the function t ↦−→ (r − 1 + it) −k belongs to L 2 (R) (this iswhere we have to assume that r ≠ 1 whenever m L ≥ 1). Thus accordingto (3.1), we get that t ↦→ L(r + it) ˆϕ(r + it) belongs to L 2 (R).Conversely assume that (ii) is satisfied. Then the function t ↦−→H(r + it) belongs to L 2 (R).Let σ 2 > max(1, r). The function H is analytic on Ω := {s : r max(1, r), we have ˆφ 1 (s) = H(s) = ˆφ(s).By injectivity of the Mellin transform, we get that( ) 1ψ χ (0,1) (t) = φ(t) = φ 1 (t).tThus t ↦−→ ψ ( ) ( )1 χ(0,1) (t) belongs to L 2 t ∗ (0, 1),dtt , which implies1−2rthat ψ belongs to L ( ) 2 du(1, +∞),u and that concludes the proof of1+2rTheorem 2.<strong>1.</strong>□Corollary 3.4. Let r > <strong>1.</strong> Then ψ ∈ L 2 ( (1, +∞),‖ψ‖ 2 = O ( r σ 1−1/2 ) , as r → +∞.)duu and we have1+2rProof. Let r > 1 be a fixed real number. By Lemma 3.3, the functiont ↦−→ H(r+it) belongs to L 2 (R) and for every 1 ≤ j ≤ m L , the functiont ↦−→ (r − 1 + it) −j belongs also to L 2 (R). Therefore it follows from(3.1) that t ↦−→ L(r + it) ˆϕ(r + it) belongs to L 2 (R) and Theorem 2.1implies that ψ ∈ L 2 du((1, +∞), ). Moreover, if we let as beforeu 1+2rφ(t) = ψ(1/t)χ (0,1) (t), we have by Plancherel’s formula‖ψ‖ 2 2 =∫ +∞1|ψ(u)| 2 du ∫ 1( )∣u = 1 ∣∣∣21+2r ∣ ψ t= ‖φ‖ 2 L 2 ∗ ((0,1), dtt 1−2r )= 1 ∫|2πˆφ(r + it)| 2 dtR= 1 ∫|H(r + it)| 2 dt,2πR0dtt 1−2r


12 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTand Lemma 3.3 gives the result.□Remark 3.5. <strong>In</strong> fact, if the coefficients of L(s) have some nice arithmeticalproperties, for example so that a Wiener-Ikehara type theoremcan be applied to L(s), one can expect that the main contributions ofψ 1 will be compensated by the main contributions of ψ 2 and so that thefunction ψ will belong to the space L ( ) 2 du(1, +∞),u for smaller value1+2rof r. For example, if L(s) = ζ(s) and if ϕ(t) = χ 0,1 (t) for all t ∈ (0, 1),then ψ(u) = u − ⌈u⌉ + 1 (so ψ(u) = {u} for almost all u ∈ (1, ∞))and ψ ∈ L ( ) 2 du(1, +∞),u for all r > 0. <strong>In</strong> this case ˆϕ(s) = 1/s and1+2rt ↦→ |ζ(r + it)|/|r + it| ∈ L 2 (R) for all r > 0 as expected by Theorem2.<strong>1.</strong> We will discuss other examples in Section 7.Remark 3.6. Assume that ϕ is bounded at t = 1 (which correspondsto σ 1 = 0). Thus, for every α > 1, we have ψ(t) = ψ 1 (t) − ψ 2 (t) =O(t α ), t → +∞. <strong>In</strong>deed, on the one hand, ψ 1 (t) = tP (log t), where Pis a polynomial of degree < m L . On the other hand, let α > 1 be afixed real number. Then there exists a constant C > 0 such thatHence we get|ψ 2 (t)| ≤ ∑ n


<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 13(1) For any polynomial P of degree d < m, for any real t > 0, wehavel(α)∑c j α j P (log(α j /t)) = 0.j=1(2) For any positive real number λ 1 and for any real number λ 2 ≥max j (α j ), the sequences ((α λ 1j ) j, (c j α 1−λ 1j ) j ) and ((α j /λ 2 ) j , c)are both m-admissible sequences.Proof. (1) If t = 1, the equality comes from the definition. If t > 0,we apply the result for t = 1 to the polynomial P t (X) = P (X − log t).The rest of the proof follows immediately from the definition. □For an m-admissible sequence A = (α, c) we define the entire functionl(α)∑g A (s) = c j αjsj=1(s ∈ C).Using (2.2), we notice that these functions satisfy(4.1) g (k)A(1) = 0 (0 ≤ k ≤ m − 1).Lemma 4.2. Let m be a nonnegative integer. Then we have the following(1) For every integer l ≥ m+1 and every 0 < α 1 < α 2 < · · · < α l ≤1, there exists c ∈ C l such that A = (α, c) is an m-admissiblesequence and g (m)A (1) ≠ 0. Furthermore we can choose c l ≠ 0.(2) For every s 1 ∈ C {1}, there exists an m-admissible sequenceA such that g A (s 1 ) ≠ 0.Proof. (1) Let l ≥ m+1, then there is a unique vector (b 1 , b 2 , . . . , b m , b l ) ∈C m+1 such that(4.2)⎛⎜⎝⎞ ⎛1 . . . . . . 1 1(log α 1 ) . . . . . . (log α m ) (log α l )⎟ ⎜.. .⎠ ⎝(log α 1 ) m . . . . . . (log α m ) m (log α l ) mb <strong>1.</strong>b mb l⎞ ⎛⎟⎠ = ⎜⎝since the corresponding van der Monde matrix is invertible. Setting⎧b j⎪⎨ α j, 1 ≤ j ≤ mc j = 0, m + 1 ≤ j ≤ l − 1⎪⎩ b lα l, j = l.0.01⎞⎟⎠ ,


14 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTThen A = (α, c) is an m-admissible sequence and g (m)A(1) = <strong>1.</strong> Furthermorewe see that b l ≠ 0 (otherwise, we would have easily b 1 = b 2 =· · · = b m = b l = 0, which contradicts (4.2)), therefore c l ≠ 0.(2) Let A = (α, c) be a non trivial m-admissible sequence.For 0 < λ < 1, consider g Aλ (s) = g A (1 − λ + λs). Note thatl∑g Aλ (s) = c j α 1−λj (αj λ ) s ,j=1and, by Lemma 4.1, A λ = ((αj λ ) j , (c j α 1−λj ) j ) is an m-admissible sequence.Now let s 1 ≠ 1; then there necessarily exists 0 < λ < 1 suchthat g Aλ (s 1 ) ≠ 0, since g A is analytic and non identically zero. □Theorem 4.3. Let r > σ 0 such that ψ ∈ L 2 du((1, +∞), ) and letu 1+2rA = (α, c) be an m L -admissible sequence. We definel(α)∑f A,r (t) = t r−σ 0c j ψj=1Then we have:(1) f A,r (t) = 0, if t > max j α j .(2) f A,r (t) ∈ L 2 dt∗((0, 1), ). t 1−2σ 0(3) For R(s) > σ 0 , we have(4.3)( αjt)(t > 0).̂f A,r (s) = −L(s + r − σ 0 ) ˆϕ(s + r − σ 0 )g A (s + r − σ 0 ).Proof. (1) We write ψ = ψ 1 − ψ 2 with ψ 1 (u) = res (L(s) ˆϕ(s)u s , s = 1)and ψ 2 (u) = ∑ n


<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 15and∫ +∞α j‖α‖|ψ(u)| 2 du ∫ 1u = |ψ 1+2r α 1 (u)| 2 du ∫ +∞j u + |ψ(u)| 2 du1+2r 1 u , 1+2r ‖α‖since ψ(u) = ψ 1 (u) for u < <strong>1.</strong> Now, by hypothesis, the second integralis finite and the first integral is also finite because ψ 1 (u) = uP (log u)and α j> 0. Since f ‖α‖ A,r(t) = 0 for t > ‖α‖, the expected result followsby linearity. Furthermore, we see that(4.4) ‖f A,r ‖ L 2 ∗ ((0,1),⎛⎛l(α)∑|c j αj|r ⎜⎝⎝j=1dtt 1−2σ ) ≤0∫ 1⎞|ψmin j α 1 (u)| 2 du⎠j u 1+2r max j α j1/2+ ‖ψ‖ L 2 ((1,+∞),⎞⎟duu 1+2r )⎠ .(3) For R(s) > 1, we have by (3.3)L(s) ˆϕ(s) =Hence∫ +∞1ψ 2 (u)u −s−1 du.l(α)∑L(s) ˆϕ(s) c j αj s =j=1==l(α)∑∫ +∞( ) −s uc j ψ 2 (u) du α j uj=11l(α)∑∫ αjc jj=10l(α)∑∫ 1c jj=10( αj) dtψ 2t t 1−s( αj) dtψ 2t t , 1−sthe last equality follows from ψ 2 (α j /t) = 0 if α j < t ≤ <strong>1.</strong>Using once more Lemma 4.1 (1), we havel(α)∑c j ψj=1( αjt) l(α)∑ ( αj)= − c j ψ 2 , (t > 0),tj=1


16 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTwhencel(α)∑L(s) ˆϕ(s) c j αj s = −j=1= −= −∫ 10∫ 10∫ 10⎛l(α)∑⎝ c j ψj=1( αjtt σ 0−r f A,r (t) dtt 1−s) ⎞ ⎠ dtf A,r (t)t s+σ 0−r−1 dtt 1−s= − ̂f A,r (s + σ 0 − r), R(s) > <strong>1.</strong>So, for R(s) > 1 + σ 0 − r, we have ̂f A,r (s) = −L(s + r − σ 0 ) ˆϕ(s + r −σ 0 )g A (s + r − σ 0 ). By the analytic continuation principle, the equalityholds for R(s) > σ 0 . (Note that from (4.1) the pole of L(s + r − σ 0 ) iskilled by the zero of g A (s + r − σ 0 ) at s = 1 − r + σ 0 .)□Remark 4.4. Let m ≥ 1, and let A = (α, c) be an m-admissiblesequence. Then, we havel∑ ( αj) l∑ ( αj) ( )) 1c j ψ = c j(ψ − α j ψ .tt tj=1Thus, if m = 1 and(l∑K r = span t r−σ 0c j ψj=1we easily check thatK r = span(t r−σ 0( αjtj=1)): A = (α, c) a 1-admissible sequence ,( ( α) ( )))1ψ − αψ : 0 < α ≤ 1 .t t<strong>In</strong> particular, if L(s) = ζ(s), ϕ(t) = χ (0,1) (t) (that is σ 1 = 0), we obtainthat ψ(u) = {u} for almost all u ∈ (1, +∞) and then K r = ˜K r , thesubspace introduced in [Nik95].5. Zeros free regions for Dirichlet seriesBefore proving Theorem 2.2 we need some well known tools concerningthe Hardy space H 2 (Π σ0 ). For these we refer to [Hof62, Chapter8]. Actually, in [Hof62], the following facts are stated for the spaceH 2 (Π 0 ) but it is easy to obtain the corresponding results for H 2 (Π σ0 ),for instance using the unitary map h(z) ↦−→ h(z − σ 0 ) from H 2 (Π σ0 )onto H 2 (Π 0 ).


<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 17Recall that if h ∈ H 2 (Π σ0 ), thenh ∗ (σ 0 + it) := limσ→σ0>h(σ + it)exists for almost every t ∈ R (with respect to the Lebesgue measure).Moreover we have h ∗ ∈ L 2 (σ 0 +iR) and ‖h‖ 2 = ‖h ∗ ‖ 2 . We can thereforeidentify (unitarily) H 2 (Π σ0 ) with a (closed) subspace of L 2 (σ 0 +iR). <strong>In</strong>the following we use the symbol h not only for the function in H 2 (Π σ0 )but also for its ”radial ” limit (in other words we forget the star). Thisidentification enables us to consider H 2 (Π σ0 ) as an Hilbert space, withscalar product given by(5.1) 〈h, g〉 2 =∫ +∞−∞h(σ 0 + it)g(σ 0 + it) dt, h, g ∈ H 2 (Π σ0 ).Now for λ ∈ Π σ0 , we have the following integral representationh(λ) = 12π∫ +∞so that we can write, using (5.1),−∞h(σ 0 + it)λ − σ 0 − it dt,(5.2)h(λ) = 〈h, k λ 〉 2 ,where k λ is the function in H 2 (Π σ0 ) defined by(5.3)k λ (z) := 1 12π z − 2σ 0 + ¯λ , z ∈ Π σ 0.The function k λ is called the reproducing kernel of H 2 (Π σ0 ) and wehave(5.4)‖k λ ‖ 2 2 = 〈k λ , k λ 〉 2 = k λ (λ) =14π(R(λ) − σ 0 ) .<strong>In</strong> particular, the linear functional h ↦−→ h(λ) is continuous on H 2 (Π σ0 ).Recall now a useful property of factorization for functions in the Hardyspace.Lemma 5.<strong>1.</strong> Let h ∈ H 2 (Π σ0 ) and µ ∈ Π σ0 such that h(µ) = 0. Thenthere exists g ∈ H 2 (Π σ0 ) such that ‖h‖ 2 = ‖g‖ 2 andh(z) =z − µz + ¯µ − 2σ 0g(z), z ∈ Π σ0 .Proof. See the results in [Hof62, Chapter 8, pp. 132] and apply thetransform h(z) ↦−→ h(z − σ 0 ).□


18 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTProof of Theorem 2.2 Denote by E r = MK r , where K r is defined by(2.3). Since the Mellin transform is a unitary map from L 2 dt∗((0, 1), ) t 1−2σ 0onto H 2 (Π σ0 ), we haveE r = span H 2 (Π σ0 ) (h A,r : A a m L -admissible sequence) ,where h A,r (s) = Mf A,r (s), for R(s) > σ 0 . It follows from Theorem 4.3thath A,r (s) = − 1 √2πL(s + r − σ 0 ) ˆϕ(s + r − σ 0 )g A (s + r − σ 0 ), R(s) > σ 0 .Now assume that there is µ ∈ Π σ0 such that L(µ + r − σ 0 ) = 0. Sinceµ ∈ Π σ0 , we get that h A,r (µ) = 0, for every m L -admissible sequence A.Thus, by continuity of h ↦−→ h(µ) in H 2 (Π σ0 ), we also have h(µ) = 0,for all h ∈ E r . Since h ∈ H 2 (Π σ0 ), we know from Lemma 5.1 that thereis g ∈ H 2 (Π σ0 ) such that ‖g‖ 2 = ‖h‖ 2 andz − µh(z) =g(z) (z ∈ Π σ0 ).z + ¯µ − 2σ 0Hence with Cauchy-Schwarz inequality and (5.2), we deduce that∣ ∣ |h(λ)| =λ − µ ∣∣∣ ∣|g(λ)| ≤λ − µ ∣∣∣λ + ¯µ − 2σ 0∣‖g‖ 2 ‖k λ ‖ 2 ,λ + ¯µ − 2σ 0so∣ sup |h(λ)| ≤µ − λ ∣∣∣∣h∈E r µ + ¯λ ‖k λ ‖ 2 .− 2σ 0‖h‖ 2 =1By contraposition, we have proved that L does not vanish on⎧⎫⎨ ∣ ∣ ∣∣∣r − σ 0 +⎩ µ ∈ C : µ − λ ∣∣∣sup h∈Er |h(λ)| ⎬‖h‖µ + ¯λ


<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 19Hencesup h∈Er |h(λ)| 2‖h‖ 2 =1‖k λ ‖ 2 2But for R(s) > σ 0 , we have()M t¯λ−2σ 0χ (0,1) (t) (s) === 1 − dist2 (k λ , E r ).‖k λ ‖ 2 2∫1 1√ t¯λ−2σ 0t s−1 dt2π01 1√2π ¯λ − 2σ 0 + s = √ 2πk λ (s),according to (5.3). Since the Mellin transform is an isometry fromL 2 dt∗((0, 1), ) onto t 1−2σ 0 H2 (Π σ0 ), we obtaindist(k λ , E r ) = 1 √2πdist(t¯λ−2σ 0χ (0,1) (t), K r ) = d r(λ)√2π.We conclude the proof of Theorem 2.2 using ‖k λ ‖ 2 12 = .4π(R(λ)−σ 0 )□Proof of Corollary 2.3 It follows from the proof of Theorem 2.2 and(5.4) that L does not vanish on⎧⎪⎨ ∣ ∣∣∣r − σ 0 +⎪⎩ µ ∈ C : µ − λµ + ¯λ − 2σ 0∣ ∣∣∣< √ 4π(R(λ) − σ 0 ) suph∈E r‖h‖ 2 =1⎫⎪⎬|h(λ)|⎪⎭ .So in particular (with h = h A,r ), we get that L does not vanish on{∣ r − σ 0 + µ ∈ C :µ − λ ∣∣∣∣µ + ¯λ < √ 4π(R(λ) − σ 0 ) |h }A,r(λ)|,− 2σ 0 ‖h A,r ‖ 2which proves corollary 2.3 because h A,r = Mf A,r = 1 √2π̂fA,r .The following could be interesting in applications.Corollary 5.2. Let M be a subspace of K r and λ ∈ Π σ0 . Then L doesnot vanish on the disc{∣ }r−σ 0 + µ ∈ C :µ − λ ∣∣∣2∣µ + ¯λ < 1 − 2(R(λ) − σ 0 )dist 2 (t¯λ−2σ 0χ (0,1) , M) .− 2σ 0Proof. It is sufficient to note thatdist(t¯λ−2σ 0χ (0,1) , M) ≥ dist(t¯λ−2σ 0χ (0,1) , K r ) = d r (λ)and then apply Theorem 2.2.Adapting the proof of Theorem 2.2, we could obtain immediatelythe following generalization.□□


20 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTTheorem 5.3. Let k ∈ N and λ ∈ Π σ0 . The function L does not haveany zero of order greater or equal to k on{∣ r − σ 0 + µ ∈ C :µ − λ ∣∣∣∣µ + ¯λ < ( 1 − 2(R(λ) − σ 0 )d 2− 2σr(λ) ) }12k.06. A Beurling–Nyman type theorem for Dirichlet seriesOne of the main steps for proving Theorem 2.4 is to prove thatthe space E r is a closed subspace of H 2 (Π σ0 ) that is invariant undermultiplication operator τ v , v ≥ 0, and then to apply Lax–Beurling’stheorem. Before recalling this theorem, we give some notations andresults. We refer to [Nik02, Part A, Chap. 2 & 6] for more details.For v ∈ R, let τ v be the operator of multiplication on L 2 (σ 0 + iR)defined by(τ v f)(σ 0 + it) = e −ivt f(σ 0 + it) (f ∈ L 2 (σ 0 + iR)).We also denote by H ∞ (Π σ0 ) the Hardy space of bounded analyticfunctions on Π σ0 ; as in H 2 (Π σ0 ), functions in H ∞ (Π σ0 ) admit ”radial”limits at almost every points of the boundary R(s) = σ 0 and iff ∈ H ∞ (Π σ0 ) and f ∗ is its boundary limits, then f ∗ ∈ L ∞ (σ 0 + iR)and ‖f ∗ ‖ ∞ = ‖f‖ ∞ (= sup z∈Πσ0 |f(z)|). A function Θ in H ∞ (Π σ0 ) issaid to be inner if |Θ(σ 0 + it)| = 1 for almost every point t ∈ R.It is easy to see that if Θ is inner and E = ΘH 2 (Π σ0 ), then E isa closed subspace of H 2 (Π σ0 ) invariant by τ v , v ≥ 0. Lax–Beurling’stheorem gives the converse.Theorem 6.1 (Lax–Beurling). Let E be a closed subspace of H 2 (Π σ0 )such that τ v E ⊂ E, ∀v ≥ 0. Then there is an inner function Θ ∈H ∞ (Π σ0 ) unique (up to a constant of modulus one) such that E =ΘH 2 (Π σ0 ).Remark 6.2. When E is spanned by a family of functions, we canprecise a little bit more the conclusion of Theorem 6.<strong>1.</strong> <strong>In</strong>deed, letE be a closed subspace of H 2 (Π σ0 ) spanned by a family of functions(f i ) i∈I , f i ∈ H 2 (Π σ0 ) and let f i = h i g i be the factorization of f i in innerfactor h i and outer factor g i . If τ v E ⊂ E, ∀v ≥ 0, then E = ΘH 2 (Π σ0 ),where Θ = gcd(h i : i ∈ I) is the greatest common inner divisor of thefamily (h i ) i∈I .Proof of Theorem 2.4 First of all, note that⋂(6.1) h −1 ({0}) = {s ∈ Π σ0 : L(s + r − σ 0 ) = 0},h∈E r


22 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTthe functions f A,r span the subspace K r , we deduce (6.2). Thereforewe obtain that E r = MK r is a closed subspace of H 2 (Π σ0 ) which isinvariant under the semi-group of operators MS β M −1 , 0 < β ≤ <strong>1.</strong>Now let us show that(6.3)MS β M −1 = τ v ,with v = − log β. If h ∈ H 2 (Π σ0 ) and if f ∈ L 2 dt∗((0, 1), ) is sucht 1−2σ 0that M −1 h = f, we have for R(s) > σ 0 ,Hence we have(MS β M −1 h)(s) = √ 1 ∫ ∞(S β f)(t)t s−1 dt2π 0= √ 1 ∫ +∞( )β −σ 0 tf t s−1 dt2π β0∫=β s−σ 1 +∞0√ f(u)u s−1 du2π0=β s−σ 0(Mf)(s) = β s−σ 0h(s).(MS β M −1 h)(σ 0 +it) = β it h(σ 0 +it) = e it log β h(σ 0 +it) = (τ v h)(σ 0 +it),with v = − log β, which proves (6.3). We therefore obtain that τ v E r ⊂E r , for all v ≥ 0 and then Lax–Beurling’s Theorem (see Theorem 6.1)implies that there is an inner function Θ in the half-plane Π σ0 suchthat E r = ΘH 2 (Π σ0 ). Moreover we know that Θ = BS, where Bis a Blaschke product and S is a singular inner function ([Nik02]).But according to Remark 6.2, the zeros of B coincide with the set ofcommon zeros of functions h ∈ E r . Hence it follows from (6.1) andthe hypothesis that B has no zeros. <strong>In</strong> other words, B ≡ <strong>1.</strong> Wewill now show that S ≡ <strong>1.</strong> First note that since h A,r is analytic onR(s) > 2σ 0 − r and since S is a common inner divisor of all functionsin E r , it follows that S can be continued analytically through the axisσ 0 + iR. <strong>In</strong> particular, this forces S to be of the form S(s) = e −a(s−σ0) ,for some a ≥ 0 (see for instance [Nik02, Part A, Chap. 4 & 6]). Nowlet h A,r be a function in E r and write h A,r = Sh, with h ∈ H 2 (Π σ0 ).Lemma 6.3. Let h ∈ H 2 (Π σ0 ). Thenlim supσ→+∞log |h(σ)|σ≤ 0.Proof of lemma 6.3. Using (5.2) and (5.4), we get|h(σ)| ≤ ‖h‖ 2 ‖k σ ‖ 2 =‖h‖ 22 √ π(σ − σ 0 ) 1/2 ,


<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 23for every σ > σ 0 . Hencelog |h(σ)|≤ O(1)σ σ − 1 2which gives the result letting σ → +∞.log(σ − σ 0 ),σ□On the one hand, according to the previous lemma, we have(6.4)lim supx→+∞log |h A,r (x)|xAnd on the other hand, writing≤ lim supx→+∞log |S(x)|x= −a ≤ 0.l(α)∑h A,r (x) = L(x + r − σ 0 ) ˆϕ(x + r − σ 0 ) c j α x+r−σ 0j ,we can assume, by Lemma 4.2, that 0 < α 1 < · · · < α l(α) = 1 and thatc l(α) ≠ 0. Then we havej=1|h A,r (x)| ∼ |a 1 || ˆϕ(x + r − σ 0 )c l(α) |, x → +∞.Hence we can find x 0 > 1 such that|h A,r (x)| ≥ |a 1|2 | ˆϕ(x + r − σ 0)c l(α) |, (x > x 0 ),which giveslog |h A,r (x)| ≥ − log 2+log |a 1 |+log |c l(α) |+log | ˆϕ(x+r−σ 0 )|, (x > x 0 ).Therefore we obtain thatlim supx→+∞log |h A,r (x)|x≥ lim supx→+∞log | ˆϕ(x + r − σ 0 )|x= 0.Using (6.4), we conclude that a = 0. Hence S ≡ 1, which givesE r = H 2 (Π σ0 ), that is K r = L 2 dt∗((0, 1), ).□t 1−2σ 0Remark 6.4. <strong>In</strong> fact, with the hypothesis on ϕ, one can easily showlog | ˆϕ(x+r−σthat we always have lim sup 0 )|x→+∞ ≤ 0. Note that therexexists some ϕ for which the lim sup is negative.7. Some examples and applications7.<strong>1.</strong> The Riemann zeta function. Letζ(s) = ∑ 1(R(s) > 1),n s n≥1be the Riemann zeta function. Then it is well known that ζ can bemeromorphically continued in the whole plane C, with a unique poleof order 1 at the point s = 1 ([Tit86]). Thus, the function ζ satisfies


24 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTour hypothesis with m L = 1 (and for instance, σ 0 = 0). Now let usconsider the function ϕ defined on [0, +∞[ by{(1 − t) −σ 1if 0 ≤ t < 1ϕ(t) =,0 if t ≥ 1where σ 1 < 1/2 is fixed.Then an elementary computation shows that ˆϕ(s) = Γ(s)Γ(1−σ 1),Γ(s+1−σ 1 )R(s) > 0. Hence,(7.1) | ˆϕ(s)| ∼ Γ(1 − σ 1)|t| 1−σ 1as |t| = |I(s)| → ∞.We now illustrate 1 the use of Theorem 2.1 in order to determine thevalues of r such that)ψ ∈ L((1, 2 du∞), .u 1+2rAn obvious computation shows thatψ(u) =u − ∑ 1( )1 − σ 1n 0.uLet µ = µ(1/2) be a convexity bound for ζ(1/2 + it) ([Ten08]), so thatwe have for all ε > 0,){O ε(|t| 1 2 −(1−2µ)r+ε if 0 ≤ r ≤ 1/2,|ζ(r + it)| = (O ) ε |t|2µ(1−r)+εif 1/2 ≤ r < <strong>1.</strong>We get from (7.1)|ζ(r + it) ˆϕ(r + it)| 2 =So a direct application of Theorem 2.1 gives:{O ε(|t|−1−(2−4µ)r+2σ 1 +ε ) if 0 ≤ r ≤ 1/2,O ε(|t|−2+2σ 1 +4µ(1−r)+ε ) if 1/2 ≤ r < <strong>1.</strong>Proposition 7.<strong>1.</strong> With the notation above, then ψ ∈ L ( 2 (1, ∞),if one the following holds:• max(0, σ 1 /(1 − 2µ)) < r ≤ 1/2;• max(1/2, 1 − (1 − 2σ 1 )/(4µ)) < r < 1;• r > <strong>1.</strong>)duu 1+2r1 It is an illustration since here Theorem 2.1 is useless as it can be seen in Theorem7.2.


<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 25By a result of Huxley ([Hux05]), we can take µ = 32 . <strong>In</strong> particular,205one may have r = 1/2 if σ 1 is chosen such that σ 1 < 141 ≈ 0.343 . . . .410If we assume the Lindelöf Hypothesis, we have µ = 0 and a similarcomputation as above implies that if r ≠ 1 and r > max(0, σ 1 ), then)ψ ∈ L((1, 2 du∞), .u 1+2r<strong>In</strong> fact, the work above is useless for the Riemann zeta function sincewe can prove directly the following:Theorem 7.2. Let r ≠ <strong>1.</strong> Then ψ ∈ L 2 du((1, ∞), ) if and only ifu 1+2rr > max(0, σ 1 ). Moreover, if the condition holds we havewhere‖ψ‖ 2 ≤ 1 L 2 du((1,∞),u 1+2r ) 2r + C(σ 1)ζ(1 + 2r − 2σ 1 )C(σ 1 ) =11+1 − 2σ 1 (1 − σ 1 ) 2 (3 − 2σ 1 ) + ε 1(1 − σ 1 ) 2and ε 1 = 1 if σ 1 ≥ 0 and ε 1 = −1 if σ 1 < 0.Proof. <strong>In</strong> the case of the zeta function and for the previous choice ofϕ, we haveψ(u) =u − ∑ (1 − n −σ1(u > 0).1 − σ 1 u)n


26 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTWith the definition of ε 1 , the function t ↦→ ε 1 (u − t) −σ 1is increasingfor t ∈ [0, u). Then∫ lε 1 u −σ 1+ε 1 (u−t) −σ 1dt ≤ ε 1whence0l∑∫ l(u − n) −σ 1≤ ε 1 (u−l) −σ 1+ε 1 (u−t) −σ 1dt,n=0ε 1 u −σ 1− ε 1 (u − l) −σ 1+ ε 1(u − l) 1−σ 11 − σ 1≤ ε 1 ∆ l (u) ≤ ε 1(u − l) 1−σ 11 − σ 1,which immediately gives (7.2). Applying (7.2) gives∞∑∫ 1()‖ψ‖ 2 2 ≤ |(u + k) −σ 1− u −σ 1| + u1−σ 1 2du1 − σ 1 (u + k) .1+2r−2σ 1k=10Using the fact that |(u + k) −σ 1− u −σ 1| = ε 1 (u −σ 1− (u + k) −σ 1) andexpanding each integral, we get(7.3) ‖ψ‖ 2 2 ≤where we setandIt is now clear that∫ 10θ A (u) = ∑ k≥1(θ2σ1 (u)ξ σ1 (u) 2 − 2θ σ1 (u)ξ σ1 (u) + θ 0 (u) ) du,ξ σ1 (u) = u −σ 1+ ε 1u 1−σ 11 − σ 11, (A < 2r).(u + k)1+2r−A(7.4) θ σ1 (u)ξ σ1 (u) ≥ 0 (0 ≤ u ≤ 1, σ 1 < 1/2),(note that by assumption σ 1 < 1/2). Hence‖ψ‖ 2 2 ≤∫ 10(θ2σ1 (u)ξ σ1 (u) 2 + θ 0 (u) ) du.A simple computation yields ∫ 1θ 0 0(u) du = 1 . For the remaining term,2rnoting that for any A < 2r, we have θ A (u) ≤ ζ(1 + 2r − A), we obtain‖ψ‖ 2 2 ≤ 1 2r + ζ(1 + 2r − 2σ 1)∫ 10ξ σ1 (u) 2 du,which gives the first part of the theorem after some obvious computations.On the other hand, suppose that σ 1 > 0. (If σ 1 ≤ 0 there is nothingto prove since by hypothesis r > σ 0 and σ 0 = 0 in our case).0


<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 27It sufficies to find U(σ 1 ) > 0 such that(7.5)∫ k+1k|∆ k (u)| 2 du ≥ U(σ 1 )uniformely in k. <strong>In</strong> that case, we shall have∞∑∫ k+1‖ψ‖ 2 2 = |∆ k (u)| 2 duk=1ku 1+2r−2σ 1∫ k+1∞∑ 1≥|∆(k + 1) 1+2r−2σ k (u)| 2 du1k=1k∞∑ 1≥ U(σ 1 )(k + 1) 1+2r−2σ 1k=1hence r > σ 1 .It now remains to prove (7.5). First by (7.2), for any k < u ≤ k + 1,we have|∆ k−1 (u)| ≤ ∣ ∣u−σ 1− (u + 1 − k) −σ 1∣ +(u + 1 − k) 1−σ 11 − σ 1≤ 2 + 21−σ 11 − σ <strong>1.</strong>Let δ > 0 fixed such that δ −σ 1= 2(2 + 21−σ 11−σ 1). Then δ < 1, and forsuch a choice of δ, we have|∆ k (u)| = |(u − k) −σ 1− ∆ k−1 (u)| ≥ 1 2 (u − k)−σ 1,for any k ≥ 1 and any k < u < k + δ. Hence,∫ k+1k|∆ k (u)| 2 du ≥ 1 4∫ k+δk(u − k) −2σ 1du = δ1−2σ 14(1 − 2σ 1 ) > 0. □7.<strong>1.</strong><strong>1.</strong> Zero-free discs for ζ. Let r > max(0, σ 1 ), r ≠ <strong>1.</strong> Then accordingto Theorem 7.2, the function ψ ∈ L ( ) 2 du(1, ∞),u . Take λ ∈ C,1+2rR(λ) > 0. It follows from Theorem 2.2 that the Riemann zeta functions ↦−→ ζ(s) does not vanish in the disc{r + µ ∈ C :µ − λ∣µ + ¯λ∣ < √ }1 − 2R(λ)d 2 r(λ) ,where d r (λ) = dist(t¯λχ (0,1) , K r ), withK r = span(f A,r : A a 1-admissible sequence).As mentioned in remark 4.4, with the choice ϕ(t) = χ (0,1) (t), we recover[Nik95, Theorem 0.1].


28 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTWe will now try to give more explicit zero free discs for the Riemannzeta function. So, let A = (α, c) be a 1-admissible sequence, whichmeans thatl(α)∑c j α j = 0,j=1with 0 < α j ≤ 1, c j ∈ C.Applying Corollary 2.3 gives that s ↦→ ζ(s) does not vanish in thedisc⎧⎨ ∣ ∣∣∣r +⎩ µ ∈ C: µ − λµ + ¯λ∣ < √ ∣∣ζ(λ + r) ˆϕ(λ + r) ∑ l(α)j=1 c jα λ+r⎫j ∣2R(λ)‖f A,r ‖ L 2 ∗ ((0,1),dt/t)But the estimate (4.4) gives for 0 < r < 1:l(α)∑(‖f A,r ‖ L 2 ∗ ((0,1),dt/t) ≤ |c j αj|r 1(1 − σ 1 ) √ 2 − 2r + ‖ψ‖ L 2 ((1,∞),and thus we deducej=1⎬⎭ .)duu 1+2r )Proposition 7.3. Let max(0, σ 1 ) < r < 1 and let A = (α, c) be a1-admissible sequence. If R(λ) > 0, then s ↦→ ζ(s) does not vanish inthe disc⎧⎨ ∣ ∣∣∣r +⎩ µ ∈ C: µ − λµ + ¯λ∣ < √ ∣2R(λ))j=1 c jα λ+r⎫j ˆϕ(λ + r)ζ(λ + r) ∣ ⎬∑ l(α)j=1 |c jαj r|(‖ψ‖ 12 +(1−σ 1 ) √ ) ⎭ .2−2r( ∑l(α)<strong>In</strong> particular, if we consider only admissible sequences of length 2,we obtain zero-free discs of the form(7.6) {r + µ ∈ C:µ − λ∣µ + ¯λ∣ < √ ∣ ( α λ+r − α ) ˆϕ(λ + r)ζ(λ + r) ∣ }2R(λ)(α r 1+ α)(‖ψ‖ 2 +(1−σ 1 ) √ ) .2−2rSince | ˆϕ(λ+r)| = O(1/|I(λ)| 1−σ 1) we obtain larger disc than in [Nik95]whenever I(λ) is large enough and σ 1 > 0. We emphasize on the factthat our zero-free discs are explicit.For example, if we take α = 1/4 in (7.6), we obtain the followingzero-free region for zeta which can be improved but has the merit tobe quite explicit:Corollary 7.4. Let r > max(0, σ 1 ) and let λ ∈ C such that R(λ) > 0.Then the Riemann zeta function does not vanish in the disc{}r + µ ∈ C:µ − λ∣µ + ¯λ∣ < F (λ, r, σ 1) ,


<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 29whereF (λ, r, σ 1 ) =√2R(λ)∣ ∣∣( ( 14(( 14) )λ+r−14) r ) (+11C(r, σ41 ) +(1−σ 1 ) √ 2−2rΓ(λ + r)Γ(1 − σ 1 )ζ(λ + r) ∣) ∣∣Γ(λ + r + 1 − σ 1 ) ∣ .with C(r, σ 1 ) = √ 1/2r + C(σ 1 )ζ(1 + 2r − 2σ 1 ) (recall that C(σ 1 ) isdefined in Theorem 7.2).Recall that the discs in Corollary 7.4 are euclidean discs of center(x, y) and radius R where:x = r + R(λ) 1 + F (λ, r, σ 1) 21 − F (λ, r, σ 1 ) 2 , y = I(λ)R = 2R(λ)F (λ, r, σ 1)1 − F (λ, r, σ 1 ) 2 .Taking, for example, λ = 0.01 + 50i, r = 0.49 and σ 1 = 0.4 then asimple evaluation of F (λ, r, σ 1 ) implies that ζ has no zero in the discof center 1 + 50i and radius 5.13 × 2 10−6 . Note that ζ has a zero ats = 1 + 49.773 . . . i and at s = 1 + 52.970 . . . i.2 27.2. Dirichlet L-functions. Let χ be a non trivial Dirichlet characterof conductor cond(χ) = q and L(χ, s) = ∑ n≥1 χ(n)n−s be the (degree1) associated Dirichlet function. This function has an analytic continuationto the whole complex plane. For simplicity, we take σ 1 = σ 0 = 0and{1 if 0 ≤ t < 1,ϕ(t) =0 if t ≥ <strong>1.</strong>So we have ψ(u) = − ∑ n 0. The admissibility condition for theu 1+2rsequence A is empty and we may takel∑ ∑f A,r (t) = t rχ(n)j=1c jn


30 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTFurthermore, we will write d r,χ if we need to specify the dependanceon χ. Remark that we have trivially d 2 r ≤ 1/(2 − 2r) (just take all thec j to be 0). <strong>In</strong> fact, we can be a little bit more precise.Proposition 7.5. We have d 2 r < 1/(2 − 2r).Proof. Let c ∈ C and 0 < α ≤ <strong>1.</strong> A short calculation gives∫ 1d 2 r ≤0 ∣ t1−r − ct ∑2r χ(n)dtn 1 − (1 − r) √ 1 − 2(1 − r)d 2 r.We expect (by the Riemann hypothesis for Dirichlet functions) thatd r = 0 which would imply that L(χ, s) does not vanish on Π r . If wehad “only”d 2 r ≤ 12 − 2r − C 22(log q) 2 (1 − r) 3for some constant C > 0, it would imply (using (7.8)) that L(χ, σ)does not vanish in the real-interval σ > 1 − C/ log q. That provesTheorem 2.5 and we also get immediately the following.Theorem 7.6. If there exists an absolute constant C such that for all(real) character χ there exists r ∈ [1/2, 1) withd 2 r,χ < 12 − 2r − C 22(log cond(χ)) 2 (1 − r) 3 ,


<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 31then there is no Siegel’s zero for Dirichlet L-functions.The previous theorem can be seen as a Beurling-Nyman criterion forSiegel’s conjecture.7.2.<strong>1.</strong> Explicit zero-free discs for L(χ, s). We now apply Corollary 2.3.Note that for the special choice of ϕ we have done (ϕ = χ (0,1) ), we haveˆϕ(s) = 1/s. Moreover, since the function ψ is bounded, say ‖ψ‖ ≤ B,we have‖ψ‖ L 2 ((1,∞),du/u 1+2r ) ≤ B/ √ 2r.Then Corollary 2.3 asserts that s ↦→ L(χ, s) does not vanish in the disc{r + µ ∈ C:µ − λ∣µ + ¯λ∣ < √ √ }|L(χ, λ + r)| 2r2R(λ) .|λ + r| BLet us remark that the Pólya-Vinogradov’s theorem implies that B ≤2 √ q log q where q = cond(χ) is the conductor of χ. Note also thatthere are some improvements of the Pólya-Vinogradov’s inequality forsome characters (see for example [GS07]).7.3. The Selberg class. Let L(s) = ∑ n≥1 a nn −s be a L-function inthe Selberg class S ([dR07a]). We denote by d its degree and by m Lthe order of its pole at s = <strong>1.</strong>As for the Riemann zeta function, we take{(1 − t) −σ 1if 0 ≤ t < 1ϕ(t) =0 if t ≥ 1where σ 1 < 1/2. By the Phragmen-Lindelöf principle, we have thatL(r + it) = O ε (t d 2 (1−r)+ε ) for 0 < r ≤ 1/2. Then we deduce fromTheorem 2.1 and (7.1) that)ψ ∈ L((1, 2 du∞),u 1+2rif the following inequality holdsσ 1 < 1 2 − 1 − r2 d.<strong>In</strong> particular, for r = 1/2, this gives σ 1 < 1/2 − d/4 which is betterthan σ 1 < −d/4 obtained in [dR07a]. Nevertheless, we should mentionthat for d < 4, we can in fact take σ 1 = 0 (see [dR07b]).


32 C. DELAUNAY, E. FRICAIN, E. MOSAKI, AND O. ROBERTReferences[BD03] L. Báez-Duarte. A strengthening of the Nyman-Beurling criterion forthe Riemann hypothesis. Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat.Natur. Rend. Lincei (9) Mat. Appl., 14(1):5–11, 2003.[BDBLS00] L. Báez-Duarte, M. Balazard, B. Landreau, and E. Saias. Notes sur lafonction ζ de Riemann. III. Adv. Math., 149(1):130–144, 2000.[Beu55] A. Beurling. A closure problem related to the riemann zeta-function.Proc. Nat. Acad. USA, 41(5):312–314, 1955.[BF84] H. Bercovici and C. Foias. A real variable restatement of Riemann’shypothesis. Israel J. Math., 48(1):57–68, 1984.[BS98] M. Balazard and E. Saias. Notes sur la fonction ζ de Riemann. <strong>1.</strong> Adv.Math., 139(2):310–321, 1998.[Bur02] J.-F. Burnol. A lower bound in an approximation problem involving thezeros of the Riemann zeta function. Adv. Math., 170(1):56–70, 2002.[dR06] A. de Roton. Une approche hilbertienne de l’hypothèse de Riemanngénéralisée. Bull. Soc. Math. France, 134(3):417–445, 2006.[dR07a] A. de Roton. Généralisation du critère de Beurling-Nyman pourl’hypothèse de Riemann. Trans. Amer. Math. Soc., 359(12):6111–6126(electronic), 2007.[dR07b] A. de Roton. On the mean square of the error term for an extendedSelberg class. Acta Arith., 126(1):27–55, 2007.[GS07] Andrew Granville and K. Soundararajan. Large character sums: pretentiouscharacters and the Pólya-Vinogradov theorem. J. Amer. Math.Soc., 20(2):357–384 (electronic), 2007.[HIP27] G. Hardy, A. <strong>In</strong>gham, and G. Pólya. Theorems concerning mean valuesof analytic functions. Proc. of the Royal Math. Soc., 113:542–569, 1927.[Hof62] K. Hoffman. Banach spaces of analytic functions. Prentice-Hall Seriesin Modern Analysis. Prentice-Hall <strong>In</strong>c., Englewood Cliffs, N. J., 1962.[Hux05] M. N. Huxley. Exponential sums and the Riemann zeta function. V.Proc. London Math. Soc. (3), 90(1):1–41, 2005.[IK04] H. Iwaniec and E. Kowalski. Analytic number theory, volume 53 ofAmerican Mathematical Society Colloquium Publications. AmericanMathematical Society, Providence, RI, 2004.[Nik95]N. Nikolski. Distance formulae and invariant subspaces, with an applicationto localization of zeros of the Riemann ζ-function. Ann. <strong>In</strong>st.Fourier (Grenoble), 45(1):143–159, 1995.[Nik02] N. Nikolski. Operators, functions, and systems: an easy reading. Vol. 1,volume 92 of Mathematical Surveys and Monographs. American MathematicalSociety, Providence, RI, 2002. Hardy, Hankel, and Toeplitz,Translated from the French by Andreas Hartmann.[Nym50][Ten08][Tit86]B. Nyman. On some groups and semi-groups of translations. PhD thesis,Upsala, 1950.G. Tenenbaum. <strong><strong>In</strong>troduction</strong> à la théorie analytique et probabiliste desnombres. 3ème édition. Belin, 2008.E. C. Titchmarsh. The theory of the Riemann zeta-function. TheClarendon Press Oxford University Press, New York, second edition,1986. Edited and with a preface by D. R. Heath-Brown.


<strong>ZERO</strong> <strong>FREE</strong> <strong>REGIONS</strong> <strong>FOR</strong> <strong>DIRICHLET</strong> <strong>SERIES</strong> 33Christophe Delaunay, Université de Franche-Comté ; Laboratoirede Mathématiques de Besançon ; CNRS UMR 6623 ; 16, route de Gray,F-25030 Besançon, France.E-mail address: christophe.delaunay@univ-fcomte.frEmmanuel Fricain, Université de Lyon ; Université Lyon 1; <strong>In</strong>stitutCamille Jordan CNRS UMR 5208; 43, boulevard du 11 Novembre 1918,F-69622 Villeurbanne, France.E-mail address: fricain@math.univ-lyon<strong>1.</strong>frElie Mosaki,Université de Lyon ; Université Lyon 1; <strong>In</strong>stitut CamilleJordan CNRS UMR 5208 ; 43, boulevard du 11 Novembre 1918, F-69622Villeurbanne, France.E-mail address: mosaki@math.univ-lyon<strong>1.</strong>frOlivier Robert, Université de Lyon & Université de Saint-Etienne ;LaMUSE (EA 3989) ; 23, rue du Dr P. Michelon, F-42000, Saint-Etienne,France.E-mail address: olivier.robert@univ-st-etienne.fr

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!