<<

HECKE OPERATORS AND L-FUNCTIONS OF MODULAR FORMS

AMR UMERI, MUZE REN

1. Motivation

We want to dene Hecke operators, which are linear maps, indexed by m ∈ N and sending modular forms of a given weight k to modular forms of the same weight and cusp forms to cusp forms:

Tm : Mk(SL(2, Z)) → Mk(SL(2, Z)), Tm : Sk(SL(2, Z)) → Sk(SL(2, Z)). Hecke operators enable us to apply methods from linear algebra into the theory of modular forms: Hecke operators are pairwise commuting and self-adjoint on , Sk(SL(2, Z)) where the hermitian form is taken to be the Petersson inner product. Hence by Spec- tral Theorem we have the existence of an orthonormal basis of consist- Sk(SL(2, Z)) ing of simultaneous eigenfunctions (called Hecke eigenforms) of all of the operators . Tm, m ∈ N 2. Definition of Hecke Operators and their first properties Fix and consider the space: . We have m ∈ N Mm = {A ∈ Mat(2, Z) | det A = m} an action of on by left-multiplication: SL(2, Z) Mm SL(2, Z) × Mm → Mm.

Hence we have a decomposition of Mm into orbits. Hecke operators can be thought of as averaging operators on the space of modular forms, where the sum is taken over orbit representatives of the above action. Before dening the Hecke operators we want to study the above action and give explicitly a set of orbit representatives of the group action. This will be done in the following proposition: Proposition 2.1. A set of orbit representatives of the above action is given by the following matrices: a b M 0 = { | ad = m, 0 ≤ b ≤ d − 1, d > 0}. m 0 d 0 Moreover two dierent elements of Mm are inequivalent, meaning that each element gives rise to a distinct orbit.

Proof. First we want to show that each element A ∈ Mm appears in some orbit. a b a b Hence we want to nd γ ∈ SL(2, ) and ∈ M 0 such that γ−1A = . Z 0 d m 0 d x x  First let 1 2 and set: x1 , x3 . Hence A = ∈ Mm a = (x1, x3), γ1 = a γ3 = a x3 x4 and by lemma of Bézout we can nd such that . (γ1, γ3) = 1 γ2, γ4 ∈ Z γ1γ4 −γ2γ3 = 1   γ1 γ2 Hence ∈ SL(2, Z). Now consider the following calculations: γ3 γ4 1 2 AMR UMERI, MUZE REN

γ γ −1 x x  γ x − γ x γ x − γ x  a γ x − γ x  1 2 1 2 = 4 1 2 3 4 2 2 4 = 4 2 2 4 . γ3 γ4 x3 x4 γ1x3 − γ3x1 γ1x4 − γ3x2 0 γ1x4 − γ3x2 This follows from:

γ4x1 − γ2x3 = γ4γ1a − γ2γ3a = a(γ1γ4 − γ2γ3) = a. γ1x3 − γ3x1 = γ1γ3a − γ3γ1a = 0. 0 Now set: b = γ4x2 − γ2x4, d = γ1x4 − γ3x2. a b0 Hence we have γ−1A = . We can ensure that d > 0 by multiplying from the 0 d −1 0  left by the matrix . By the division algorithm we can also nd b = b0 −kd, 0 −1 1 −k where k ∈ and 0 ≤ b ≤ d − 1 then we just multiply from the left by Z 0 1   to obtain the matrix with the desired properties: a b . It remains to prove that 0 d a b ad = m. This follows by simple calculations or by observing that γ0A = , for 0 d some γ0 ∈ SL(2, Z) and that the of A equals m. 0 Now we want to prove the second claim, namely that dierent elemens in Mm give a b a0 b0  dierent orbits. Let , ∈ M 0. Assume there is γ ∈ SL(2, ) such 0 d 0 d0 m Z that a b a0 b0  γ = . 0 d 0 d0 This means that    0 0  aγ1 bγ1 + dγ2 a b = 0 . aγ3 aγ3 + dγ4 0 d

This implies that γ3 = 0, but then since γ1γ4 = 1, we have that: γ1, γ4 = 1 or− 1. 0 0 0 Since d, d > 0, we have that γ1 = γ4 = 1. But then d = d and 0 ≤ b, b ≤ d − 1 0 0 implies that −d + 1 ≤ b − b ≤ d − 1. But b − b = dγ2. Hence γ2 = 0. Hence γ a b a0 b0  is necessarily the identity matrix and = . The converse implication 0 d 0 d0 proves the claim.  Before dening the Hecke operators we need to recollect some denitions:

a b Denition 2.2. Let γ = ∈ GL+( ) and z ∈ . Dene j(γ, z) = cz + d. c d 2 R H 1 Now let f : H → C and an integer k xed. Dene the function

f[γ]k : H → C,

k/2 −k f[γ]k(z) = (det γ) j(γ, z) f(γz). This function has the following properties:

+ f[γ1γ2]k = f[γ1]k[γ2]k ∀γ1, γ2 ∈ GL2 (R), is a linear operator + f 7→ f[γ]k ∀γ ∈ GL2 (R). 1 In literature one often nds the following notation for this function: f|kγ. HECKE OPERATORS AND L-FUNCTIONS OF MODULAR FORMS 3

Remark 2.3. In general we cannot expect f[A]k to be a for arbitrary + and . Hecke operators (which will be dened below) A ∈ GL2 (R) f ∈ Mk(SL(2, Z)) are special in that regard: By summing elements of the form f[A]k for a given modular form f we indeed get back a modular form! Now we are ready to dene the Hecke operators: Denition 2.4. Let . We dene the m-th Hecke operator as follows: m ∈ N Tm For let f ∈ Mk(SL(2, Z)) k/2−1 X (2.1) Tmf := m f[A]k.

A∈Mm/∼ The sum is meant to be taken over arbitrary orbit representatives.2 After showing that the above statement is well-dened, i.e independent of the choice of orbit represen- tatives, we can insert our distinguished set of orbit representatives from proposition 2.1 to obtain: d−1 X X  a b  T f = mk/2−1 f . m 0 d ad=m b=0 k d>0 For z ∈ H we have: d−1 X X az + b T f(z) = mk−1 d−kf . m d ad=m b=0 d>0 Lemma 2.5. The Hecke operators are well-dened, i.e the denition does not depend on the choice of orbit representatives. Proof. We want to show now that the Hecke operators are well-dened. For this let be the (nite) number of orbits and Let 0 N {Ai | i = 1,...,N}, {Ai | i = 1,...,N} be two arbitrary orbit representatives. Then up to reordering of the indices we have that 0 for some Then the claim follows directly Ai = γiAi γi ∈ SL(2, Z), i = 1,...,N. from the weak modularity of f: N N N N X X 0 X 0 X 0 f[Ai]k = f[γiAi]k = f[γi]k[Ai]k = f[Ai]k. i=1 i=1 i=1 i=1  Theorem 2.6. Let , . Then is m ∈ N f ∈ Mk(SL(2, Z)) Tmf : H → C (i) holomorphic on H, (ii) holomorphic at ∞, (iii) weakly modular of weight k on SL(2, Z). Hence and is a linear Tmf ∈ Mk(SL(2, Z)) Tm : Mk(SL(2, Z)) → Mk(SL(2, Z)) operator on the space of modular forms of weight k. Proof. The rst claim is obviously true. The second claim will follow from the next lemma, where we will discuss the action of Tm on the Fourier expansion of f ∈ . Let us prove the third claim. Given we must show Mk(SL(2, Z)) f ∈ Mk(SL(2, Z)) that: Tmf[γ]k = Tmf ∀γ ∈ SL(2, Z). 2 This notation may seem ambiguous, since the elements in Mm/ ∼ are orbits. However this should not lead to any confusion. 4 AMR UMERI, MUZE REN Consider the following:

 k/2−1 X  k/2−1 X Tmf[γ]k = m f[A]k [γ]k = m f[Aγ]k = Tmf.

A∈Mm/∼ A∈Mm/∼ The second equality follows from known properties. The last equality follows from the fact that is another set of orbit representatives. {Aγ | A ∈ Mm/ ∼}  To prove holomorphicity at ∞, we need to know how the Hecke operator acts on the Fourier expansions of modular forms. For that let us consider the next lemma: Lemma 2.7. Let have the Fourier expansion: f ∈ Mk(SL(2, Z)) ∞ X 2πinz f(z) = fne . n=0

Then the Fourier expansion of Tmf is given by the following formula: ∞   X X k−1 2πinz Tmf(z) = a f mn e . a2 n=0 a|n,m

Hence Tmf is holomorphic at ∞. Proof. We need the following easy result for the proof:

d−1 ( X d, if d | n e2πinb/d = . 0, otherwise b=0 Let P∞ 2πinz be the Fourier expansion. Applying the Hecke operator we f(z) = n=0 fne obtain:

d−1 ∞ d−1 ∞ k−1 X X −k X 2πin az+b 1 X X mk X 2πin az 2πin b T f(z) = m d f e d = f e d e d m n m d n ad=m b=0 n=0 ad=m b=0 n=0 d>0 d>0 ∞ ∞ X mk−1 X 2πin az X mk−1 X 2πinaz = f e d = f e d n d nd ad=m n=0 ad=m n=0 d>0 d|n d>0 ∞ ∞ X k−1 X 2πinaz X X k−1 2πirz = a f nm e = a f rm e . a a2 a|m n=0 r=0 a|m a|r

 Corollary 2.8. If , then . Hence the Hecke f ∈ Sk(SL(2, Z)) Tmf ∈ Sk(SL(2, Z)) operators maps cusp forms to cusp forms and we have the restriction map:

Tm : Sk(SL(2, Z)) → Sk(SL(2, Z)).

The next goal is to show that the Hecke operators Tm form a commuting family of endomorphisms on and . Mk(SL(2, Z)) Sk(SL(2, Z))

Lemma 2.9. If (m, n) = 1, then TmTn = Tmn = TnTm. HECKE OPERATORS AND L-FUNCTIONS OF MODULAR FORMS 5

Proof. Let . Let P∞ 2πirz be the Fourier expansion. f ∈ Mk(SL(2, Z)) f(z) = r=0 fre Consider the mn-th Hecke operator Tmn. It has the following Fourier expansion: ∞   X X k−1 2πirz Tmnf(z) = t f rmn e . t2 r=0 t|r,mn Now there is a bijection (exercise) between the tuples of divisors {(a, b) | a|r, b|r, a|n b|m} and the divisors {t | t|r, t|mn} given by (a, b) 7→ ab. (Here we use the rst time that (n, m) = 1). Hence we can rewrite the above sum in the following way:

∞   ∞   X X k−1 k−1 2πirz X X X k−1 k−1 2πirz Tmnf(z) = a b f rmn e = a b f rmn e . a2b2 a2b2 r=0 ab|r,mn r=0 b|r,m a|r,n

Now consider the n-th Hecke operator Tn. It has the following Fourier expansion: ∞   X X k−1 2πirz Tnf(z) = a f rn e . a2 r=0 a|r,n ˜ Now set f = Tnf and apply the m-th Hecke operator. This gives: ∞   ˜ X X k−1 ˜ 2πirz Tm(Tnf)(z) = Tmf(z) = b f rm e . b2 r=0 b|r,m ˜ Now notice that fk is the k-th Fourier-coecient of Tnf. Hence: X k−1 f˜rm = a f rmn . b2 b2a2 a| rm ,n b2 Now inserting the Fourier-coecient into the above sum we get: ∞   X X X k−1 k−1 2πirz a b f rmn e . a2b2 r=0 b|r,m a| rm ,n b2 But now consider the following: rm implies that . By assumption . Hence by using for the second time a| b2 a|rm a|n that (n, m) = 1, we have that (a, m) = 1. Then a|rm implies that a|r. (This can be proven by lemma of Bézout). Hence the above sum simplies to: ∞   X X X k−1 k−1 2πirz a b f rmn e . a2b2 r=0 b|r,m a|r,n What we have shown is that for arbitrary it holds that f ∈ Mk(SL(2, Z)) TmTnf = Tmnf. Now the claim follows easily from the commutativity of the natural numbers:

TmTn = Tmn = Tnm = TnTm.  Lemma 2.10. Let p be a prime number. Then we have the following:

∀r, s ∈ N : Tpr Tps = Tps Tpr . 6 AMR UMERI, MUZE REN

Proof. See for example: [2] or see [1] for a dierent approach.  Theorem 2.11. The Hecke operators commute:

∀m, n ∈ N : TmTn = TnTm.

Proof. Let . By prime factorization, put: α1 αj β1 βk . m, n ∈ N n = p1 . . . pj , m = q1 . . . qk Then:

α α TnTm = Tp 1 ...Tp j T β1 . . . t βk . 1 j q1 qk

Now, if pi = ql for some i, l, we can commute them by lemma 2.10. If they are not equal, we can commute them by lemma 2.9. Hence the claim follows.  We will state an important property of the Hecke operators on , Sk(SL(2, Z)) namely that they are self-adjoint w.r.t the Petersson inner product: Theorem 2.12. The Hecke operators are self- Tm : Sk(SL(2, Z)) → Sk(SL(2, Z)) adjoint w.r.t the Petersson inner product, i.e

hTmf, gi = hf, Tmgi ∀ f, g ∈ Sk(SL(2, Z)). Proof. First one shows that the statement is true for Poincaré-series. Then one notices that the Poincaré-series span (last talk). See for example: [6], theorem Sk(SL(2, Z)) 6.12. 

3. associated to modular forms Each modular form has an associated Dirichlet series, its L- f ∈ Mk(SL(2, Z)) function. Let P∞ n 2πinz be its Fourier expansion, let be a complex f(z) = n=0 a e s ∈ C variable, and write formally

∞ X −s L(s, f) = ann . n=1 Convergence of L(s, f) in a half plane of s-values follows from estimating the Fourier coecents of f. Note that the Dirichlet series begains from 1. Proposition 3.1. If is a then its fourier coecients f ∈ Mk(SL(2, Z))) k/2 satisfy an = O(n ). If f is not a cusp form then its fourier coeents satisfy k an = O(n ). Proof. let 2πiτ 2πi(x+iy), let P∞ n, a holomorphic function on the q = e = e g(q) = n=1 anq unit disk q : |q| < 1, then for any r ∈ (0, 1), we have Z Z 1 1 −n −2πin(x+iy) an = g(q)q dq/q = f(x + iy)e dx 2πi |q|=r x=0 for any y > 0, and letting y = 1/n Z 1 2π −2πinx an = e f(x + i/n)e dx x=0 Since f is a cusp form, we know that it decays rapidly when it approaches the cusp. Thus Im(τ)k/2|f(τ)| is bounded on the upper half plane, so estimating the last k/2 integral, we kown that |an| ≤ Cn , then we get the result. HECKE OPERATORS AND L-FUNCTIONS OF MODULAR FORMS 7 If is an Eisenstein series in , its th Fourier coecient of is Ek Mk(SL(2, Z)) n f σk−1(n), and X X n X 1 σ = dk−1 = ( )k−1 = nk−1 < nk−1ζ(k − 1) k−1 d dk−1 d|n d|n d|n k−1 we have that |an| ≤ Cn . Since every modular form is the sum of a cusp form and an Eisenstein series, the rest follows.  Corollary 3.2. if is a cusp form then converges absultely f ∈ Mk(SL(2, Z)) L(s, f) for all s with <(s) > k/2+1. If f is not a cusp form then L(s, f) converges absolutely for all s with <(s) > k

k/2 Remark 3.3. Daniel Bump in [3] remarked that: The estimate |an| ≤ Cn called the trival estimate, is due to Hardy (1927) and (more simply) Hecke (1937). The (k−1)/2+ correct estimate |an| ≤ Cn for any  > 0 was conjectured (for f = ∆) by Ra- manujan (1916); this famous statement, the Ramanujan conjecture, was nally proved around 1970 by Deligne(1971) using dicult techniques from algebraic geometry. In order to estimate the convergence of some integrals which we will use later, we need the following lemma Lemma 3.4. For a sequence ∞ of complex numbers, for , put {an}n=0 z ∈ H ∞ X 2πinz (3.1) f(z) = ane , n=0 v and an = O(n ) with some v > 0. Then the right-hand side of equation 3.1 convergent absolutely and uniformly on any compact subset of H, and f(z) is holomorphic on H. Moreover, f(z) = O(Im(z)−v−1) (Im(z) → 0) −2π Im(z) f(z) − a0 = O(e ) (Im(z) → ∞) uniformly on <(z) Proof. By the formula n!nv+1 Γ(v + 1) = lim , n→∞ (v + 1)(v + 2) ... (v + n + 1) we have for v > 0, −v − 1 lim nv/(−1)n = Γ(v + 1), n→∞ n v And because an = O(n ), there exisits L > 0 such that −v − 1 |a | ≤ L(−1)n n n for all n ≥ 0. Put z = x + iy, then we have ∞ ∞ X X −v − 1 |a ||e2πinz| ≤ L( (−1)n e−2πny) n n n=0 n=0 = L(1 − e−2πy)−v−1 Because (1 − e−2πy) = O(y) as y → 0, we see that |f(z)| = O(y−v−1), also we have f(z) is bounded when y → ∞. 8 AMR UMERI, MUZE REN

As 2πinz P∞ 2πinz and P∞ 2πinz also satisfy the as- f(z) − a0 = e ( n=0 an+1e ), n=0 an+1e sumption of the lemma, by the conclusion we just proved, it is bounded when y → ∞. Therefore, we have as y → ∞ ∞ 2πinz X 2πinz −2πy f(z) − a0 = e ( an+1e ) = O(e ) n=0  Proposition 3.5. The L-function L(s, f) has meromorphic continuation to all s and satises a functional equation. In fact, if Λ(s, f) = (2π)−sΓ(s)L(s, f) then Λ(s, f) extends to an analytic function of s if f is a cusp form; if it is not a cusp form, then it has simple poles at s = 0 and s = k. In general, it satises Λ(s, f) = (−1)k/2Λ(k − s, f) Proof. By the lemma 3.4, we know that for and P∞ −2πnt t > 0 <(s) > k + 1, n=1 ane and P∞ R ∞ s −2πnt −1 converges absolutely. n=1 0 ant e t dt For <(s) > k + 1, we have Λ(s, f) = (2π)−sΓ(s)L(s, f) ∞ Z ∞ X −s −t s−1 = an(2πn) e t dt n=1 0 ∞ Z ∞ X s −2πnt −1 = ant e t dt 0 (3.2) n=1 Z ∞ ∞ s X −2πnt −1 = t ( ane )t dt 0 n=1 Z ∞ s −1 = t (f(it) − a0)t dt 0 Z ∞ Z ∞ a0 −s −1 s −1 = − + t f(i/t)t dt + t (f(it) − a0)t dt s 1 1 Because f is a modular form, we have f(it) = (−1)k/2t−kf(i/t), then k/2 Z ∞ Z ∞ a0 (−1) a0 k/2 k−s −1 s −1 Λ(s; f) = − − +(−1) t (f(it)−a0)t dt+ t (f(it)−a0)t dt s k − s 1 1 By the lemma 3.4, we have −2πt f(it) − a0 = O(e ), so that Z ∞ k−s −1 t (f(it) − a0)t dt 1 and Z ∞ s −1 t (f(it) − a0)t dt 1 converges absolutely and uniformly on any vertical strip. Therefore they are holo- morphic on the whole s-plane. If we dene Λ(s, f) for any s ∈ C by the integral, it is then a meromorphic function on the whole s-plane. The functional equation follows.  HECKE OPERATORS AND L-FUNCTIONS OF MODULAR FORMS 9 Now we consider the Euler property of the L-function. As Daniel Bump remarks in the [3]: the rst historical hint that a should be associated with the L-series of a modular form came from Ramanujan's investigation of ∆. The Fourier coecients of ∆ comprise Ramanujan's tau function: ∆(z) = P τ(n)qn. Ramanujan (1916) conjectured, and Mordell (1917) proved shortly afterward, that ∞ X Y τ(n)n−s = (1 − τ(p)p−s + p11−2s)−1. n=1 p The true explanation of this identity requires the theory of Hecke operators. The Hecke algebra is a commutative family of self-adjoint operators on the nite- dimensional Sk(SL2(Z)), by the spectral theorem, we konw there exists a basis of functions which are eigenfunction of all the Hecke operators. We call them the Hecke eigenform. We will show that these eigenforms has the property of Euler product. And also it is interesting to notice that after some normalization. We can nd that the eigenvalues of the Hecke operator are hidden in the Fourier coeents of the so called normalized Hecke eigenform.

Proposition 3.6. Now we suppose that f is a Hecke eigenform. Let fn denotes its Fourier coecients. Then fn satisfy the following:

(1) f1 6= 0 (2) if f1 = 1, then λ(n) = fn for all n (3) if f1 = 1 then the coeents fn are multiplicative; that is, if (m, n) = 1, we have fmn = fmfn Proof. By lemma 2.7 , we have that

X k−1 (3.3) λ(n)fm = a f mn a2 a|n,a|m Suppose that (m, n) = 1. Then the only a that divides both m and n is a = 1, so it reduces to

(3.4) λ(n)fm = fnm Taking , this implies that , then we get the conclusion. m = 1 fn = λ(n)f1 

Thus we can adjust the coeent of the Hecke eigenform by making f1 = 1, such a Hecke eigenform will be called normalized. Theorem 3.7. If f is a normalized Hecke eigenform, then X −s Y −s k−1−2s −1 (3.5) L(s, f) = fnn = (1 − fpp + p ) n≥1 p Proof. Because the coeents of f is multiplicative, then we have ∞ X −s Y X −rs L(s, f) = fnn = ( fpr p ) n≥1 p r=0 In order to prove equation 3.5, we only need the equity ∞ −s k−1−2s X −rs (1 − fpp + p )( fpr p ) = 1 r=0 10 AMR UMERI, MUZE REN To prove this, rst observe that by lemma 2.7, we have

X k−1 λ(n)fm = a f mn a2 a|n,a|m by taking m = pr, n = p, then we have k−1 fpr+1 − fpfpr + p fpr−1 = 0 Multipling by Xr+1, and summing this equation for r ≥ 1, we have X r+1 X r X k−1 2 r−1 fpr+1 X − (fpX)(fpr X ) + (p X )(fpr−1 X ) = 0 r≥1 r≥1 r≥1 by adding the corresponding term, we have

X r X r X k−1 2 r fpr X − (fpX)(fpr X ) + (p X )(fpr X ) = 1 r≥0 r≥0 r≥0 that is ∞ k−1 2 X r (1 − fpX + p X )( fpr X ) = 1 r=0 taking X = p−s, we get the Euler product.  This section is based on [3] and [5].

4. Hecke's converse theorem We rst write the following condition for conveience to quote later.

Condition 4.1. let f(z) be a function on H, f(z) has a fourier expansion ∞ X 2πinz f(z) = ane , n=0 which converges absolutely and uniformly on any compact subset of H. And there exists v > 0 such that: f(z) = O(Im(z)−v), (Im(z) → 0), uniformly on <(z).

If f is holomorphic on H, satises the condition 4.1, then by a similar proof as v propostion 3.1, we have that an = O(n ). Remark 4.2. By the above lemma 3.4, all holomorphic functions f(z) on H satisfying the condition 4.1 correspond bijectively to all sequences ∞ of complex numbers {an}n=1 v such that an = O(n ) with v > 0.

Theorem 4.3 (Phragmen-Lindelöf). For two real numbers v1, v2 v1 < v2, put

F = {s ∈ C|v1 ≤ <(s) ≤ v2}. Let φ be a holomorphic function on a domain contating F satisfying |φ(s)| = O(e|τ|δ )(|τ| → ∞), s = σ + iτ uniformly on F with δ > 0. For real number b, if on <(s) = v1 and <(s) = v2, |φ(s)| = O(|τ|δ)(|τ| → ∞) HECKE OPERATORS AND L-FUNCTIONS OF MODULAR FORMS 11 then uniformly on F, we have: |φ(s)| = O(|τ|δ)(|τ| → ∞)

Proof. By assumption, there exists L > 0 such that |φ(s)| ≤ Le|τ|δ . First consider the case when b = 0. Then there exist M > 0, such that |φ(s)| ≤ M on the lines <(s) = v1 and <(s) = v2. Let m be a positive integer such that m ≡ 2 mod 4. Put s = σ + iτ. Since <(sm) = <((σ + iτ)m) is a polynomial of σ and τ, and the highest term of τ is −τ m, we have <(sm) = −τ m + O(|τ|m−1)(|τ| → ∞), uniformly on F , so that <(sm) has an upper bound on F . Taking m and N so that m m > δ and <(s ) ≤ N, we have, for any  > 0, on <(s) = v1 and <(s) = v2, |φ(s)esm | ≤ MeN , and |φ(s)esm | = O(e|τ|δ−τ m+K|τ|m−1 ) → 0, (|τ| → ∞) uniformyly on F. By the maximum principle, we see |φ(s)esm | ≤ MeN , (s ∈ F ) Letting  → 0, we obtain that |φ(s)| ≤ M, namely, φ(s) = O(|τ|0). Next assume b 6= 0. We dene a holomorpic function:

b b log(s−v1+1) ψ(s) = (s − v1 + 1) = e , where log takes the principal value. Since

<(log(s − v1 + 1) = log(|s − v1 + 1|), we have uniformly on F b b |ψ(s)| = |s − v1 + 1| ∼ |τ| (|τ| → ∞)

Put φ1(s) = φ(s)/ψ(s). Then φ1(s) satises the same condition as φ with b = 0, so that by the above result, φ1(s) is bounded on F , thus we get the result for arbitrary b.  For a holomorphic function P∞ 2πinz on satisfying condition 4.1, f(z) = n=0 ane H we put P∞ −s, since v , converges absolutely and L(s, f) = n=1 ann an = O(n ) L(s, f) uniformly on any compact subset of <(s) > 1 + v, so that it is holomorphic on . We call the Dirichlet series associated with f. For , we <(s) > 1 + v √ L(s, f) N > 0 −s put ΛN (s; f) = (2π/ N) Γ(s)L(s, f) Theorem 4.4. Let P∞ 2πinz and P∞ 2πinz be holomorphic f(z) = n=0 ane g(z) = n=0 bne funtions satisfying the condition 4.1. For positive k and N, the following condition (1) and (2) are equivalent. √ (1) g(z) = (−i Nz)−kf(−1/Nz) (2) Both ΛN (s; f) and ΛN (s; g) can be analytically continued to the whole s-plane, satisfy the functional equation

ΛN (s; f) = ΛN (k − s; g), and a b Λ (s; f) + 0 + 0 N s k − s 12 AMR UMERI, MUZE REN

is holomorphic on the whole s-plane and bounded on any vertical strip σ1 ≤ <(s) ≤ σ2. Proof. From (1)⇒ (2): similar to the proof of proposition 3.5. From (2)⇒(1) By the Mellin inversion formular: For σ > 0, 1 Z e−t = Γ(s)t−sds, 2πi <(s)=σ let t = 2πny, we have ∞ 1 X Z f(iy) = a (2πny)−sΓ(s)ds + a 2πi n 0 n=1 <(s)=α for any If then P∞ −s is uniformly convergent and α > 0. α > v + 1, L(s, f) = n=1 ann bounded on Re(s) = α, so that by the Stirling's estimate √ Γ(s) ∼ 2πτ σ−1/2e−π|τ|/2, (s = σ + iτ, |τ| → ∞), √ −s ΛN (s; f) = (2π/ N) Γ(s)L(s; f) is absolutely integrable. Therefore we can ex- change the order of summation and integration, and Z √ 1 −s f(iy) = ( Ny) ΛN (s; f)ds + a0. 2πi <(s)=α Since L(s; f) is bounded on <(s) = α, we see, for any µ > 0, −µ |ΛN (s; f)| = O(|Im(s)| )(|Im(s)| → ∞) on <(s) = α by Stirling's estimate. Next take β so that k − β > v + 1. A similar argument implies that for any µ > 0, −µ |ΛN (s; f)| = |ΛN (k − s; g)| = O(|Im(s)| )(|Im(s)| → ∞) on <(s) = β. By assumtion, a b Λ (s; f) + 0 + 0 N s k − s is bounded on the domain β ≤ <(s) ≤ α. Hence for any µ > 0, we see by 4.3 −µ |ΛN (s; f)| = O(|Im(s)| )(|Im(s)| → ∞) holds unifomly on the domain Furthermore we assume that √ β ≤ <(s) ≤ α. α > k and Since −s has simple poles at and with residues β < 0.√ ( Ny) ΛN (s; f) s = 0 s = k −k −a0 and ( Ny) b0, respectively, we can change the integral paths from <(s) = α to <(s) = β and obtain Z √ √ 1 −s −k f(iy) = ( Ny) ΛN (s; f)ds + ( Ny) b0. 2πi <(s)=β By the function equation, Z √ √ 1 −s −k f(iy) = ( Ny) ΛN (k − s; g)ds + ( Ny) b0 2πi <(s)=β Z √ √ 1 s−k −k = ( Ny) ΛN (s; g)ds + ( Ny) b0 2πi <(s)=k−β √ = ( Ny)−kg(−1/(iNy)) HECKE OPERATORS AND L-FUNCTIONS OF MODULAR FORMS 13

Since f(z) and g(z) are holomorpic on H, we obtain √ f(z) = ( Nz/i)−kg(−1/Nz), √ g(z) = (−i Nz)−kf(−1/Nz)  Since Γ(1) is generated by two elements T and S, we can easily characterize an element f(z) of Mk(Γ(1)) by the functional equation of L(s, f) and obtain an Corollary 4.5. Let k be an even integer ≥ 2, assume a holomorphic function f(z) on satises the condition. Then belongs to if and only if H f(z) Mk(Γ(1)) Λ(s; f) = (2π)−sΓ(s)L(s, f) can be analytically continued to the whole s-plane, a (−1)k/2a Λ(s; f) + 0 + 0 s k − s is holomorphic on H and bounded on any vertical strip and satises the functional equation : Λ(s; f) = (−1)k/2Λ(k − s; f).

Moreover if a0 = 0, then f(z) is a cusp form. This part is based on [4] and [3]. I also used some proofs from the book [5]. References [1] Serre, J-P. (1973). A Course in Arithmetic. Springer. [2] Murty, R. (2016). Problems in the Theory of Modular Forms. Hindustan Book Agency/ Springer. [3] Bump, D. (1997). Automorphic Forms and Representations. Cambridge University Press. [4] Miyake, T. (1989). Modular Forms. Springer. [5] Diamond, F. & Shurman, J. (2005). A First Course in Modular Forms. Springer. [6] Iwaniec, H. (1997). Topics in Classical Automorphic Forms. American Mathematical Society.