SUMS OF FOURIER COEFFICIENTS OF AUTOMORPHIC CUSP FORMS by Theran Bassett B.Sc., University of Northern British Columbia, 2019 THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF MASTER OF SCIENCE IN MATHEMATICS UNIVERSITY OF NORTHERN BRITISH COLUMBIA April 2022 © Theran Bassett, 2022 Abstract Let f be a primitive cusp form of weight k and level N . Let L ( s, sym2 f ) be the symmetric square L-function associated to f . We denote by Asym 2 f ( n ) the n-th coef ficient in the Dirichlet series representation of L ( s, sym2 f ) . The work in this thesis is motivated by the relation between the size of partial sums of the coefficients and the first negative coefficient in the sequence {Asym2 f ( n )}n . We prove that for x large enough, if Asym2 f ( n ) 0 for all n x, then Y_ Asym ( n) cNx ( logx ) for 2f 2 n x ^ ( n,N ) = 1 some positive constant CM depending only on N . Assuming the Generalized Riemann Hypothesis, we also determine a range of x, in terms of k and N , for which n $:x Asym 2 f ( n ) = o ( x ( logx ) 2 ) . These results are GL3 analogues of recent work of Y. Lamzouri on "Large Sums of Hecke Eigenvalues of Holomorphic Cusp Forms". We also extend the work of Lamzouri to the number field setting by proving analo- gous results for the sequence {Af ( ti )}„ of normalized Fourier coefficients associated with a primitive Hilbert cusp form f defined over a totally real number field of narrow class number 1. ii TABLE OF CONTENTS Abstract ii Table of Contents iii Acknowledgements iv 1 Introduction 1.1 Motivation 1.2 Research Problem 1.3 Thesis Overview 1.4 Conventions and Notation 1 1 3 4 5 Classical Modular Forms 2.1 Definitions and Basic Properties of Modular Forms 2.2 L-functions and Symmetric Square L-functions . . 6 16 3 Review of Literature 21 4 Sums of Coefficients of Symmetric Square L-functions 4.1 Analytic Tools and Preliminary Lemmas 4.2 Proof of Theorem 4.0.1 4.3 Proof of Theorem 4.0.2 25 26 29 36 5 Sums of Fourier Coefficients of Hilbert Modular Forms 5.1 Definition and Basic Properties of Hilbert Modular Forms Proof of Theorem 5.1.2 5.2 56 56 61 6 Concluding Remarks 76 2 Bibliography 6 76 iii Acknowledgements First, I would like to begin by expressing my profound appreciation to my bril- liant supervisor, Dr. Alia Hamieh. The past few years have been a turbulent time for myself and much of the world . One of the few constants throughout the duration of this thesis, was the consistent readiness of Dr. Hamieh to provide me with her invaluable insight or much needed encouragement . Due to her commitment to being a wonderful supervisor, I have learned priceless skills that will forever benefit me as a mathematician and as a person. I would also like to extend my sincere gratitude towards my committee members, Dr. Edward Dobrowolski, and Dr. Margot Mandy for their very valuable time and advice. For the former, were it not for his passion of mathematics and engaging teaching style, I would not have pursued mathematics. So a special thank you to Dr. Dobrowolski for bringing me into the world of mathematics. In addition, I would like to thank the entire Department of Mathematics and Statistics at UNBC. Working alongside its members was always easy due to their friendly natures and bountiful advice. Of course, my sincere gratitude goes to my biggest supporter and anchor, my wonderful wife. iv Chapter 1 Introduction 1.1 Motivation There is a quote attributed to the German number theorist Martin Eichler which says "There are five fundamental operations of mathematics: addition, subtraction, multiplication, division, and modular forms". This bold statement conveys the importance Eichler, and incidentally many other mathematicians, place on the topic of modular forms. Modular forms appear in many areas of mathematics including algebraic geometry, representation theory and number theory as well as areas outside of pure mathematics with applications to string theory [20]. This field of research is most closely associated with number theory since many important arithmetic results have been derived from the theory of modular forms. Most no- tably, Andrew Wiles proved Fermat's Last Theorem by connecting modular forms to elliptic curves. Within the short history of the theory, many mathematicians have been allured to researching modular forms. This is owed to the potential strength of results that lie within the theory in addition to the broad spectrum of mathematical techniques that are used to study modular forms. In this work, we explore questions related to the Fourier coefficients of modular forms from an ana- 1 lytic perspective. Interest in the sign changes of these coefficients, and in particular the first sign change, spiked when Kowalski, Lau, Soundarajan, and Wu [15] made a strong connection between this problem and the problem of least quadratic non- residue, which is a classical topic in analytic number theory. The Fourier coefficients of modular forms encode a lot of arithmetic information and are worthy of deep research. By just scratching the surface, one can show several deep results. As an example, Jacobi's 4-square problem asks how many ways an integer can be written as the sum of four squares. For instance, there are eight ways to represent 1, since we count all the ordered 4-tuples ( a, b, c, d ) such that a2 + b2 + c2 + d2 = 1 . Jacobi gave a wonderful elementary answer by inves- tigating the coefficients of a famous modular form, the theta function. The theta function is defined as the infinite sum 00 0 ( z) = X neZ qn2 = 1 + 2 e2mn " = neZ 2 qn, where q = e 2mz , Im ( z ) > 0 n=1 and is known to be a modular form of weight \. Jacobi noted that the n-th Fourier coefficients of 04 ( z ) , which is a modular form of weight 2, would be exactly the number of ways n can be represented as the sum of four squares since 04 ( z ) = Y_ qn Y- qm Z 2 2 neZ meZ TGZ = SGZ Y , ,r ,seZ nm qn2+m2+r2+s 2 = 1 + 00 ^ r4 ( u ) qn . n =1 Jacobi then had the task of finding an expression for the theta function from which he could easily deduce the coefficients of 04 ( z ) . He accomplished his task by writing the theta function in terms of another classical modular form, the eta function. 2 From there it was simple to establish the incredible formula r4 ( n ) = SY_ d, d |n | d4 where r4 ( n ) represents the number of ways we can write n as a sum of four squares. Other less obvious arithmetic properties can be deduced from understanding coefficients of modular forms. Let dk ffk ( n ) = d|n be the sum of the k-th powers of the positive divisors of n. If asked to prove the relation TL — 1 0 be given, and letx Suppose that Asym 2 f ( n ) 0 /or all n ^ ^ l be such that logx > max o> ( N ) 1 + e, x. Then 2 Y- Asym ( n ) cNx (logx ) , 2f nsjx ( n,N ) =1 where cN is a positive constant depending only on N . We also prove that the sum appearing in Theorem 1.2.1 is o (x ( logx ) 2 ) in a certain range of x depending on k and N . Theorem 1.2.2. Assume that L ( s, sym2 f ) satisfies the Generalized Riemann Hypothesis . * In the range loglog log kN —y oo, we have HAsym fW = 0 (x ( logx ) 2) . n x ^ 2 These theorems can be viewed as GL3 analogues of [16, Theorem 1.1] and [16, Corollary 1.2]. We also extend this work to the number field setting by proving the analogue of Theorem 1.2.2 for the sequence {Af ( n )}„ of normalized Fourier coefficients associated with a primitive Hilbert cusp form f defined over a totally real number field of narrow class number 1. 1.3 Thesis Overview This thesis is organized into six chapters. In Chapter 2, we introduce the basic theory for modular forms and their associated L-functions and symmetric square 4 L-functions. In Chapter 3 we briefly present the major contributions leading up to the work of Lamzouri [16] which inspired this research project. Our aim in this chapter is to focus on implications of the results, the methods of their proofs, and the constraints the authors worked within, rather than providing a rigorous mathematical account. In Chapter 4, we prove Theorem 1.2.1 and Theorem 1.2.2 after presenting all the required technical results from complex analysis and an- alytic number theory. In Chapter 5 we provide an overview of Hilbert modular forms and prove a main theorem related to their Fourier coefficients. This theorem mirrors Theorem 1.2.2, but in the number field setting. In Chapter 6 we describe possible avenues for future research. 1.4 Conventions and Notation In this work, we adopt the following conventions and notation. Given two functions f ( x ) and g ( x ) we write f ( x ) < g ( x ), g ( x ) » f ( x ) orf ( x ) = 0 ( g ( x ) ) to mean there exists some positive constant M such that |f ( x ) | y M| g ( x ) | for x large enough. We write f ( x ) x g ( x ) when both estimates f ( x ) C g ( x ) and g ( x ) < f ( x ) hold simultanef (x) ously. We write f ( x ) ~ g ( x ) if lim —— = 1 . We write f ( x ) = o ( g ( x ) ) when g ( x ) 0 < ^ f fx ) for sufficiently large x and lim —— = 0. We follow the convention of denoting an arbitrarily small positive constant by e, which may change from instance to instance. The letter p ( resp . p ) will be exclusively used to represent a prime number ( resp . prime ideal) . 5 Chapter 2 Classical Modular Forms In this chapter, we provide a brief account of the theory of modular forms. The reader is referred to standard textbooks on the topic such as [3] or [11] for a com- plete exposition. All the theorems discussed in this chapter can be found in the aforementioned references. 2.1 Definitions and Basic Properties of Modular Forms Definition (Upper Half Plane ). The upper half plane, which we denote by H, is the set of complex numbers with positive imaginary part. More precisely, we have H = (z e C : Im ( z ) > 0}. Let S1_2 ( Z ) be the group of 2-by-2 matrices with integer entries and determinant 1. This group is referred to as the modular group. For y e S1_ 2 ( Z ) and z e H, we set az + b where y = a b c d 6 Observe that for z = x + iy e H and y e SL2 ( Z ), we have az + b _ a ( x + ty ) + b _ ( ax + b + lay ) ( cz + d — icy ) ( cx + d ] 2 + ( cy ) 2 cz + d c ( x + iy ) + d and so Im ( yz ) = V ( cx + d ) 2 + ( cy ) 2 > 0. Hence, yz e H for all z e H and y e SL2 ( Z ) . Definition (Modular Transformation). Let k be a positive integer. Given a function f : H -> C be a function that satisfies (2.1) . We say that f is a modular form of weight k if , in addition, it is holomorphic on H and at oo . Next, we explain the requirement that f ( z ) is holomorphic at z = oo . The map g defined by z ^ e 2niz transforms H to the punctured open unit disk D * = {q e C : 0 < | q | < 1 ). Since f is periodic of period 1 , we can write f ( z ) = g ( e2mz ), where g ( q ) 7 is holomorphic on D * . The function g has a Laurent series expansion g( q) = Y - a( ^ cin TIEZ rL _ for q e D * . Since | q | = e 27lIm ( z ) , we see that q 0 as Im ( z ) oo. Hence, thinking of z = oo as lying far in the imaginary direction, we say that f is holomorphic at oo if g extends holomorphically to the puncture point q = 0 in which case its Laurent series sums over n e N . This means that a modular form f has a Fourier series expansion OO t( z) = 22 n=0 where q = e 2mz . The coefficients a ( n ) are called the Fourier coefficients of the — modular form f . Showing that a holomorphic function f : H » C that satisfies (2.1) is holomorphic at oo doesn't require computing the Fourier expansion of f . In view of the relation q —> 0 Im ( z ) oo, it suffices to show that lim Im ( z)-Kx> f (z) exists. Definition (Cusp Form ). A cusp form of weight k is a modular form f of weight k whose Fourier expansion has leading coefficient a ( 0 ) = 0, i.e. OO f ( z) = ^ a ( u ) qn, q = e2niz . n=1 It is well-known that the set of all modular forms of weight k forms a finite dimensional vector space over C denoted by Mk ( SL2 ( Z ) ) or simply Mk . The set of all cusp forms of weight k forms a subspace of Mk, denoted by Sk ( SL2 ( Z ) ) or 8 simply Sk . The dimension of Mk as a vector space over C is given by dim ( Mk ) = if k = 2 ( mod 12 ) LfsJ [£ J + 1 if k 2 ( mod 12 ) . ^ In particular, it easy to see that M2 = {0} and that Mk is one dimensional for all even integers 2 < k 10. Let k > 2 be an even integer. For non-trivial examples of modular forms in Mk, consider the Eisenstein series Gk given by 1 Gk ( z ) = ( m,n ) y ( 0,0 ) ( mz + ri ) k ' Z H. This can be viewed as a 2-dimensional analogue of the Riemann zeta function 00 C( k ) = 1 One can show that Gk is indeed a modular form of weight k. More- n=1 over, it admits the Fourier expansion Gk ( z ) = 2 C ( k ) + 2 - ( 2m ) k ^ In fact, for k e N, one can prove that any f e Mk can be written as a polynomial in G 4 and Gg . In practice, it is common to work with the normalized Eisenstein series of weight k GkW F M] 2C ( k) ' for which the leading Fourier coefficient is 1. The space of cusp forms Sk together with the space EkC gives Mk = Sk © EkC if k > 2. In particular, Mg = EgC since we know that it is 1-dimensional. This observation is at the heart of the proof of (1.1) presented in Chapter 1. Observe that | E e Mg, and so Eg = | E . Hence, the identity in (1.1) follows by comparing Fourier coefficients. This simple proof gives us a glimpse to the infinitely many other arithmetic facts that we can derive from Eisenstein's series alone. 9 In what follows we focus our discussion on the space of cusp forms Sk . In order to study this space further, we make it into an inner product space. Definition (Petersson Inner Product ). The Petersson inner product ( , ) : Sk x Sk C is given by ( f, g ) = ]J ykf z g z dxda a ' ( ) ( ) 2 SL 2 ( Z ) \ H where z = x + iy . One can verify that the integral appearing in the above definition is well-defined because the integrand is SL2 fZ ) -invariant. Moreover, the integral is finite since both f and g are cusp forms. In particular, the space Sk is a finite dimensional Hilbert space with this inner product. We now introduce the Hecke operators Tn on the linear space Mk of modular forms of weight k. Let M ( m ) be the set of all 2-by-2 matrices with integer entries and determinant m. The modular group SL2 ( Z ) acts on M ( m ) from both sides so that M ( m ) = SL2 ( Z ) M ( m ) = M ( m ) SL2 ( Z ) . The collection a b A ( m) = : ad = m, 0 b< d 0 d forms a complete set of representatives of M ( m ) modulo SL2 ( Z ), i.e. we have the disjoint partition M ( m ) = [J SI_2 ( Z ) p. The m-th Hecke operator acting on f e pGA ( m ) Mk is given by the formula Tmf = mH £V pGA ( m ) _ where we set f |y = det ( p ) 2 ( c z + d ) kf ( yz ) when y — ( £ jj ) . In what follows we present a formula for Tm that is more suitable for computational purposes. Definition (Hecke Operator ). The m-th Hecke operator Tm acting on a modular 10 form f G Mk is defined by Theorem 2.1.1. The Hecke operator Tm is a linear transformation on Mk . It maps a modular form to a modular form and a cusp form to a cusp form: Tm : Mk and Mk Tm : Sk Sk . Next, we present a result that shows how Tm acts on the Fourier coefficients of f . Let f be given by the Fourier expansion 00 2=>( u ) e f (z) = 2 2 (2.2) . n 0 Then Tmf is given by the Fourier expansion OO Tmf ( z ) = ^ bm ( u ) e2 z, n=0 where bm ( u ) = Y_, dk l a ( r ) - d| ( m n ) ^ (2.3) This relation is key to studying the coefficients of modular forms because it allows the Fourier coefficients to be expressed in terms of the eigenvalues of the operator. This is made possible by the following important result. Theorem 2.1.2. In the space Sk of cusp forms of weight k there exists an orthogonal basis which consists of eigenvectors of all the Hecke operators Tm . Since each such vector is a modular form, we refer to it as a Hecke eigenform instead . Theorem 2.1.2 follows from the spectral theorem of linear algebra because 11 Theorem 2.1.3. The Hecke operators commute, i.e. TmTn = TnTm . Moreover, the Hecke operators acting on Sk are self-adjoint , i .e. (Tnf , g ) = (f , Tng ) for all f , g e Sk . Another important result about Hecke operators is that they satisfy the multi- plicative property Tmri = TmTn for all ( m., n ) = 1 . Moreover, they satisfy the recur_ rence formula Tpj + i = TpTpl - pk 1 Tpj _ i for all j 1. Let f be a Hecke eigenform satisfying Tmf = A ( m ) f VmeN . Suppose that f is given by the Fourier expansion (2.2). Using (2.3), we see that A ( m) a ( n ) = ^ dk 1 a ( mnd 2 ) . d | ( ra,n ) For n = 1, this yields A ( m ) a ( l ) = a ( m ) . Notice that a ( l ) f 0 for otherwise f vanishes identically. Hence, the Fourier coefficients of a Hecke eigenform are pro- portional to the eigenvalues of the Hecke operators. We say that f is normalized if a ( l ) = 1 . The set of all normalized Hecke eigenforms of weight k is commonly denoted by dfk . The entire theory that we have presented so far is centered around a function f : H C that is holomorphic on H and at oo and satisfies the transformation property f ( yz ) = ( cz + d ) kf ( z ) , Vy = ( “ * ) e SL2 ( Z ) . Replacing the modular group SL2 ( Z ) by a subgroup F in the last condition generalizes the notion of modular forms and allows for more examples. In fact, the modular form 0 that was mentioned in relation to Jacobi's 4-square problem in Chapter 1 is one such example. In what follows we will revisit most of the above constructions when SLifZ j is 12 replaced by a certain subgroup. The exposition, however, will be less detailed than the previous. Definition (Modular Transformation w.r.t F). Let k be a positive integer, and let F — be a finite index subgroup of SL2 ( Z ) . Given a function f : H » C, we say that f satisfies the modular transformation property of weight k with respect to F if f ( yz ) = ( cz + d ) kf ( z ) , When (2.4) is applied with y = ( Q1 ^ Vy = ( “ ^ ) er . (2.4) ) , we see that f = 0 unless k is an even integer. Since V has a finite index in SL2 ( Z ), there exists M y 1 such that ( J ' ) e h Y In particular, (2.4) implies that f (z + M) = f ( ( J Y ) ) = C, ) kf (z) = f ( z) . Z Hence, f can be written as a function in which we denote by g . More precisely, there is a function g such that f ( z ) = g ( qjvi ) . The function g is holomor- phic in the punctured unit disk. If g extends to a holomorphic function at 0, we say that f is holomorphic at oo. This gives rise to the Fourier series expansion OO f (z) = g ( q M ) = Y=- n 0 where qM = e 2 ^ . We say that f vanishes at infinity if a ( 0 ) = 0. If a G SL2 ( Z ), then a-1 Fa is a finite index subgroup of SL2 ( Z ) . For y G F, one can verify that f a satisfies the modular transformation property for the subgroup a-1 Fa. Hence, f |a has a Fourier expansion at oo. We say that f is holomorphic at the cusps if f | is holomorphic at oo for all a G SL2 ( Z ) . We say that f vanishes at the CT cusps if f | vanishes at oo for all cr G SL2 ( Z ) . ff 13 The most commonly used subgroups V are of the form r0 ( N ) = { ( “ § ) G SL2 ( Z ) : c = 0 ( mod N ) j and r, ( IM ) = { ( “ S ) e S L2 ( Z ) : c = 0 ( m o d N ) / a = d = 1 (m o d N ) } for positive integers N . Notice for example that Fo ( 1 ) = SL2 ( Z ) is the full modular group. In this chapter we will only discuss modular forms arising from the subgroups r0 ( N ) . Definition (Modular Form of Level N ). Let f : H C be a function that satisfies (2.4) for Fo ( N ) . We say that f is a modular form of weight k and level N if, in addition, it is holomorphic on H and at the cusps. Definition (Cusp Form of Level N ). A cusp form of weight k and level N is a modular form f of weight k and level N which vanishes at the cusps. It is well-known that the set of all modular forms of weight k and level N forms a finite dimensional vector space over C denoted by Mk ( N ) . The set of all cusp forms of weight k and level N forms a subspace of Mk ( N ) , denoted by Sk ( N ) . There is a well-developed theory of Eisenstein series of weight k and non-trivial level N , but we will not discuss it here. In what follows we focus our discussion on the space of cusp forms Sk ( N ) . In order to study this space further, we make it into an inner product space. Definition (Petersson Inner Product ). The Petersson inner product ( , ) : Sk x Sk C is given by < f , 9) = y k n z ) g ( z ) dxdy ' | J D 2 r0 ( N ) \ H where z = x + iy . 14 Much like the case of Sk, the space Sk ( N ) is a finite dimensional Hilbert space with this inner product . One can also define Hecke operators on Sk ( N ) much like we did for Sk . For economy of exposition, we will not go into the details of this construction. Instead, we will state the most important theorems that result from this construction. Theorem 2.1.4. The Hecke operators commute , i.e. TmTn = TnTm . Moreover, for all f , g G Sk, we have ( Tnf , g ) = ( f , Tng ) if ( n, N ) = 1 . Therefore, Theorem 2.1.5. In the space Sk ( N ) there exists an orthogonal basis which consists of eigenvectors of all the Hecke operators Tn with ( n, N ) = 1 . Since each such vector is a modular form, we refer to it as a Hecke eigenform instead . Let f be a Hecke eigenform satisfying Tmf = A ( m ) f V ( m, N ) = 1 . (2.5) Suppose that f is given by the Fourier expansion (2.2). Using a formula similar to (2.3) for the Fourier expansion of Tmf ( z ) , we get A ( m ) a ( 1 ) = a ( m ) for all ( m, N ) = 1 . Unfortunately, this doesn't imply that a ( 1 ) f 0, as there are Hecke eigenforms f fO such that a ( l ) = 0. Such forms are known to come from lower levels and are called oldforms. In fact, one notices that a modular form is not always unique to its level. For example, a modular form of level N will necessarily be a modular form of level M if N |M. When working in the level aspect, it is often useful to ignore those modular from that come from lower levels. More precisely, the space of cusp forms Sk ( N ) can be decomposed as Sk ( N ) = S°ld ( N ) ® S£ew ( N ) where