Fact-checked by Grok 2 weeks ago

Modular form

A modular form is a f: \mathcal{H} \to \mathbb{C} on the upper half-plane \mathcal{H} = \{\tau \in \mathbb{C} : \Im(\tau) > 0\} that satisfies the transformation property f\left(\frac{a\tau + b}{c\tau + d}\right) = (c\tau + d)^k f(\tau) for all \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \mathrm{SL}_2(\mathbb{Z}) and \tau \in \mathcal{H}, where k \in 2\mathbb{Z}_{\geq 0} is the weight, and which remains bounded as \Im(\tau) \to \infty. These functions admit a Fourier expansion f(\tau) = \sum_{n=0}^\infty a_n q^n with q = e^{2\pi i \tau}, where the coefficients a_n often encode arithmetic data such as divisor sums or class numbers. A subclass known as cusp forms consists of those modular forms that vanish at the cusp \tau = i\infty, meaning a_0 = 0. The theory of modular forms originated in the 19th century through studies of elliptic and theta functions by mathematicians such as Jacobi and Eisenstein, who explored their symmetries under linear fractional transformations. By the early 20th century, the modern analytic definition emerged, with significant contributions from Hecke on operators that act on spaces of modular forms, revealing their algebraic structure and connections to L-functions. These spaces are finite-dimensional vector spaces over \mathbb{C}, spanned by examples like the Eisenstein series E_k(\tau) for even weights k \geq 4, which have explicit formulas involving zeta values, and the discriminant cusp form \Delta(\tau) of weight 12. Modular forms play a central role in number theory, particularly through their association with elliptic curves via the modularity theorem (formerly the Taniyama-Shimura conjecture), which states that every elliptic curve over \mathbb{Q} corresponds to a modular form of weight 2. This connection underpinned Wiles's proof of in 1994. Their Fourier coefficients satisfy multiplicative properties under Hecke operators, linking them to problems in arithmetic statistics, such as Ramanujan's congruences for the function p(n), and broader frameworks like the , where they relate Galois representations to automorphic forms. Beyond number theory, modular forms appear in physics (e.g., string theory functions) and (e.g., linking the to weight-0 forms).

Core Definitions

Holomorphic Modular Forms

A holomorphic modular form of weight k \in \mathbb{Z} for the modular group \Gamma = \mathrm{SL}_2(\mathbb{Z}) is a function f: \mathbb{H} \to \mathbb{C} that is holomorphic on the upper half-plane \mathbb{H} = \{\tau \in \mathbb{C} \mid \Im(\tau) > 0\} and satisfies the transformation property f\left(\frac{a\tau + b}{c\tau + d}\right) = (c\tau + d)^k f(\tau) for all \gamma = \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \Gamma. Additionally, f must be holomorphic at the cusp i\infty, meaning it admits a Fourier q-expansion f(\tau) = \sum_{n=0}^\infty a_n q^n with q = e^{2\pi i \tau}, where the series converges uniformly on compact subsets of \mathbb{H}. This boundedness as \Im(\tau) \to \infty ensures no poles or essential singularities at the cusp. The space of such forms, denoted M_k(\Gamma), forms a finite-dimensional . For odd k, M_k(\Gamma) = \{0\}, while for even k \geq 0, the is \dim M_k(\Gamma) = \lfloor k/12 \rfloor if k \equiv 2 \pmod{12}, and \lfloor k/12 \rfloor + 1 otherwise. The of cusp forms S_k(\Gamma), consisting of those with a_0 = 0 (vanishing at i\infty), has \dim S_k(\Gamma) = \dim M_k(\Gamma) - 1 for k > 2. These spaces are graded by , and modular forms generate rings like the one for \Gamma spanned by . For a general congruence subgroup \Gamma \subseteq \mathrm{SL}_2(\mathbb{Z}) of finite index, a weak modular form of weight k is a holomorphic function on \mathbb{H} transforming as f(\gamma \tau) = (c\tau + d)^k f(\tau) for \gamma = \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \Gamma. A holomorphic modular form for \Gamma is then a weak modular form that extends holomorphically to the compactified upper half-plane \mathbb{H}^* = \mathbb{H} \cup \mathbb{P}^1(\mathbb{Q}), ensuring holomorphy at all cusps of \Gamma. The cusps are the orbits \Gamma \backslash \mathbb{P}^1(\mathbb{Q}), and holomorphy at a cusp \xi requires that the function f|_{k, \sigma}(\tau) = (c\tau + d)^{-k} f(\sigma \tau), for a matrix \sigma sending \xi to i\infty, has a q-expansion without negative powers. If \Gamma contains parabolic elements like \begin{pmatrix} 1 & 1 \\ 0 & 1 \end{pmatrix}, the q-expansions are well-defined and provide a basis for M_k(\Gamma). The dimension formula for M_k(\Gamma) with even k > 0 is \dim M_k(\Gamma) = (k-1)(g-1) + \frac{k}{4} \nu_2 + \frac{k}{3} \nu_3 + \frac{k}{2} \nu_\infty, where g is the of the modular curve X(\Gamma), and \nu_2, \nu_3, \nu_\infty count elliptic points and cusps of \Gamma. For cusp forms, \dim S_k(\Gamma) = (k-1)(g-1) + \frac{k}{4} \nu_2 + \frac{k}{3} \nu_3 + \left(\frac{k}{2} - 1\right) \nu_\infty when k > 2. These formulas arise from Riemann-Roch theorems applied to line bundles on X(\Gamma). Representative examples include the Eisenstein series E_k(\tau) = 1 - \frac{2k}{B_k} \sum_{n=1}^\infty \sigma_{k-1}(n) q^n for even k \geq 4, which are non-zero modular forms for \Gamma = \mathrm{SL}_2(\mathbb{Z}) generating M_k(\Gamma) alongside lower-weight forms. The discriminant \Delta(\tau) = q \prod_{n=1}^\infty (1 - q^n)^{24} = \sum_{n=1}^\infty \tau(n) q^n, a cusp form of weight 12, satisfies \Delta(\tau) \neq 0 on \mathbb{H} and generates the cusp form ring. For subgroups like \Gamma_0(N), Hecke operators act on M_k(\Gamma_0(N)), preserving the space and enabling decompositions into eigenforms.

Modular Functions

Modular functions are meromorphic functions on the upper half-plane \mathbb{H} = \{ z \in \mathbb{C} \mid \operatorname{Im}(z) > 0 \} that are invariant under the action of a discrete subgroup \Gamma of \mathrm{SL}(2, \mathbb{Z}), such as a , and extend meromorphically to the cusps when \mathbb{H} is compactified to the extended upper half-plane \mathbb{H}^*. They satisfy f(\gamma z) = f(z) for all \gamma = \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \Gamma and z \in \mathbb{H}, reflecting weight-zero automorphy without the multiplier factor present in higher-weight forms. In relation to holomorphic modular forms, which are analytic functions on \mathbb{H} transforming as f(\gamma z) = (cz + d)^{k} f(z) for integer weight k \geq 0 and holomorphic at cusps, modular functions can be viewed as meromorphic modular forms of weight zero. The space of modular functions for \Gamma forms the field of meromorphic functions on the compact \Gamma \backslash \mathbb{H}^*, often denoted X(\Gamma), and is generated by quotients of modular forms of positive weight. Their Fourier expansions at cusps take the form f(z) = \sum_{n=-\infty}^{\infty} a_n q^n where q = e^{2\pi i z} (or a suitable power for other cusps), with coefficients typically in rings of algebraic . A fundamental property is their role in parametrizing isomorphism classes of elliptic curves: for the full modular group \Gamma(1) = \mathrm{SL}(2, \mathbb{Z}), every modular function is a rational function of the j-invariant j(z), defined as j(z) = q^{-1} + 744 + 196884 q + 21493760 q^2 + \cdots, which has a simple pole at the cusp \infty and maps \mathbb{H}/\Gamma(1) biholomorphically onto \mathbb{C}. For general \Gamma, modular functions generate function fields of modular curves and satisfy algebraic relations known as modular equations, such as the relation for level 2: X^3 + Y^3 - X^2 Y^2 + 1488 XY (X + Y) - 162000 (X^2 + Y^2) + 40773375 XY + 8748000000 (X + Y) - 157464000000000 = 0, where X = j(z) and Y = j(2z). In , modular functions underpin by generating Hilbert and ring class fields via values at CM points; for instance, j(\tau) for \tau with complex multiplication by orders in imaginary quadratic fields produces unramified abelian extensions. They also connect to Hecke operators, which act on spaces of modular forms and descend to the function field, and feature in the , linking elliptic curves to cusp forms. Historically, their study traces to Poincaré's work on Fuchsian groups in the and Kronecker's Jugendtraum, with key advancements by Shimura and Deligne in the –1970s establishing their arithmetic significance.

Modular Forms for the Full Modular Group SL(2, ℤ)

Standard Analytic Definition

A modular form of weight k \in 2\mathbb{Z}_{\geq 0} for the full modular group \Gamma = \mathrm{SL}(2, \mathbb{Z}) is defined as a function f: \mathbb{H} \to \mathbb{C} that is holomorphic on the upper half-plane \mathbb{H} = \{\tau \in \mathbb{C} \mid \mathrm{Im}(\tau) > 0\} and satisfies the automorphy condition f(\gamma \tau) = (c\tau + d)^k f(\tau) for all \gamma = \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \Gamma. This transformation law ensures invariance under the action of \Gamma on \mathbb{H} via Möbius transformations \gamma \tau = \frac{a\tau + b}{c\tau + d}. Equivalently, the condition can be expressed using the Petersson slash operator: (f \mid_k \gamma)(\tau) = (c\tau + d)^{-k} f(\gamma \tau) = f(\tau) for all \gamma \in \Gamma. The weight k must be even and non-negative for the space of such forms to be finite-dimensional, though the definition applies more generally. Holomorphy at the cusps, particularly at the single cusp \infty for \Gamma, requires that f admits a Fourier expansion of the form f(\tau) = \sum_{n=0}^\infty a_n q^n, \quad q = e^{2\pi i \tau}, with no negative powers of q, ensuring the expansion converges holomorphically in a neighborhood of q = 0. This q-expansion arises from the stabilizer of \infty in \Gamma, which consists of translations \tau \mapsto \tau + 1. The space of all such functions is denoted M_k(\Gamma) and forms a finite-dimensional complex vector space. If the constant term a_0 = 0 in the q-expansion, then f is a cusp form, belonging to the subspace S_k(\Gamma). This analytic framework, originating in the work of Poincaré and developed by Hecke, captures the essential symmetry and regularity properties central to the theory.

Lattice and Elliptic Curve Perspectives

Modular forms for the full modular group \mathrm{SL}(2, \mathbb{Z}) admit an equivalent characterization as homogeneous functions on lattices in the complex plane \mathbb{C}. A lattice \Lambda \subset \mathbb{C} is a discrete subgroup generated by two linearly independent elements, say \Lambda = \mathbb{Z} \omega_1 + \mathbb{Z} \omega_2 with \mathrm{Im}(\omega_2 / \omega_1) > 0. A function F: \{\Lambda\} \to \mathbb{C} on the set of lattices is homogeneous of weight k if F(\lambda \Lambda) = \lambda^{-k} F(\Lambda) for all \lambda \in \mathbb{C}^\times. Such functions descend to well-defined holomorphic functions f on the upper half-plane \mathbb{H} via the identification \Lambda_z = \mathbb{Z} z + \mathbb{Z} for z \in \mathbb{H}, where f(z) = F(\Lambda_z), and the transformation law f(\gamma z) = (c z + d)^k f(z) for \gamma = \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \mathrm{SL}(2, \mathbb{Z}) follows from the action on lattices by \gamma \cdot \Lambda_z = \Lambda_{\gamma z}. This lattice viewpoint naturally arises in the construction of classical examples like the . The of weight k \geq 4 even is given by G_k(z) = \sum_{(m,n) \neq (0,0)} (m z + n)^{-k}, which sums over the dual lattice \Lambda_z^\vee = \{ w \in \mathbb{C} \mid \langle w, \lambda \rangle \in \mathbb{Z} \ \forall \lambda \in \Lambda_z \} and extends homogeneously as G_k(\lambda \Lambda) = \lambda^{-k} G_k(\Lambda). Its Fourier expansion is G_k(z) = 2 \zeta(k) + \frac{2 (2 \pi i)^k}{(k-1)!} \sum_{n=1}^\infty \sigma_{k-1}(n) q^n, where q = e^{2 \pi i z} and \sigma_{k-1}(n) = \sum_{d \mid n} d^{k-1}, confirming its status as a modular form. Conversely, any modular form on \mathbb{H} pulls back to a on lattices, providing a bridge between analytic and geometric interpretations. From the elliptic curve perspective, lattices parametrize isomorphism classes of elliptic curves over \mathbb{C}. Each lattice \Lambda yields the elliptic curve E_\Lambda = \mathbb{C} / \Lambda with group law induced by complex addition, and two elliptic curves E_{\Lambda} and E_{\Lambda'} are isomorphic if and only if \Lambda' is homothetic to \Lambda, i.e., \Lambda' = \lambda \Lambda for some \lambda \in \mathbb{C}^\times. The moduli space of such elliptic curves up to isomorphism is the quotient \mathrm{SL}(2, \mathbb{Z}) \backslash \mathbb{H}^*, where \mathbb{H}^* = \mathbb{H} \cup \mathbb{Q} \cup \{ i \infty \} compactifies via the action extended to the extended rationals. The absolute invariant j(\Lambda), defined via Weierstrass invariants g_2(\Lambda) = 60 G_4(\Lambda) and g_3(\Lambda) = 140 G_6(\Lambda) as j(\Lambda) = 1728 \frac{g_2(\Lambda)^3}{\Delta(\Lambda)} with discriminant \Delta(\Lambda) = g_2(\Lambda)^3 - 27 g_3(\Lambda)^2, is a modular function of weight 0 that classifies these isomorphism classes uniquely. The Weierstrass \wp-function \wp_\Lambda(z) = \frac{1}{z^2} + \sum_{\omega \in \Lambda \setminus \{0\}} \left( \frac{1}{(z - \omega)^2} - \frac{1}{\omega^2} \right) satisfies the equation \wp'(z)^2 = 4 \wp(z)^3 - g_2(\Lambda) \wp(z) - g_3(\Lambda), embedding the elliptic curve as a cubic in the projective plane. Since g_2 and g_3 are homogeneous of weights 4 and 6, respectively, they correspond to modular forms E_4(z) and E_6(z) (normalized Eisenstein series) on \mathbb{H}, and \Delta(z) = (2\pi)^{12} q \prod_{n=1}^\infty (1 - q^n)^{24} is the unique cusp form of weight 12 up to scalar. This perspective underscores how modular forms encode arithmetic data of elliptic curves, such as their discriminants and j-invariants, facilitating connections to Diophantine geometry.

Fundamental Examples

The Eisenstein series provide the primary non-cusp examples of modular forms for the full modular group \mathrm{SL}(2, \mathbb{Z}). For even integers k \geq 4, the Eisenstein series G_k(\tau) is defined by the lattice sum G_k(\tau) = \sum_{(m,n) \in \mathbb{Z}^2 \setminus \{(0,0)\}} \frac{1}{(m\tau + n)^k}, where \tau lies in the upper half-plane \mathbb{H}. This series converges absolutely for k > 2 and defines a holomorphic function on \mathbb{H} that transforms as G_k\left(\frac{a\tau + b}{c\tau + d}\right) = (c\tau + d)^k G_k(\tau) for all \gamma = \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \mathrm{SL}(2, \mathbb{Z}), making it a modular form of weight k. The normalized version is E_k(\tau) = G_k(\tau) / (2\zeta(k)), where \zeta(k) is the Riemann zeta function value at k, and its q-expansion (with q = e^{2\pi i \tau}) is E_k(\tau) = 1 - \frac{2k}{B_k} \sum_{n=1}^\infty \sigma_{k-1}(n) q^n, with B_k the k-th Bernoulli number and \sigma_{k-1}(n) the sum of the (k-1)-th powers of the divisors of n. The lowest-weight examples are E_4 and E_6, of weights 4 and 6, respectively. Their q-expansions begin E_4(\tau) = 1 + 240 \sum_{n=1}^\infty \sigma_3(n) q^n, \quad E_6(\tau) = 1 - 504 \sum_{n=1}^\infty \sigma_5(n) q^n. These forms are holomorphic everywhere on \mathbb{H} and at the cusp i\infty, with constant term 1, and they generate the ring of all modular forms for \mathrm{SL}(2, \mathbb{Z}) as \mathbb{C}[E_4, E_6]. For instance, E_4^2 spans the space of weight-8 forms, while higher weights follow from symmetric products. Eisenstein series like E_4 and E_6 arise in the theory of elliptic functions, such as the Weierstrass \wp-function, where g_2 = 60 G_4 and g_3 = 140 G_6. A fundamental cusp form example is the modular discriminant \Delta(\tau) of weight 12, the unique (up to scalar) nonzero element in the cusp form space S_{12}(\mathrm{SL}(2, \mathbb{Z})). It is given explicitly by \Delta(\tau) = \frac{E_4(\tau)^3 - E_6(\tau)^2}{1728} = (2\pi)^{12} \eta(\tau)^{24} = q \prod_{n=1}^\infty (1 - q^n)^{24}, where \eta(\tau) = q^{1/24} \prod_{n=1}^\infty (1 - q^n) is the . This form vanishes to order 1 at the cusp i\infty (hence cuspidal) and is holomorphic on \mathbb{H}, with no other zeros in the fundamental domain. The coefficients \tau(n) in its q-expansion \Delta(\tau) = \sum_{n=1}^\infty \tau(n) q^n are the , satisfying multiplicative properties under Hecke operators. \Delta plays a central role in the j-invariant, j(\tau) = E_4^3 / \Delta, which classifies elliptic curves up to isomorphism. Theta series offer another perspective on modular forms, though the classical Jacobi \theta(\tau) = \sum_{n \in \mathbb{Z}} e^{\pi i n^2 \tau} is a modular form of weight $1/2 for the \Gamma_\theta = \Gamma_0(4) \cap \Gamma_1(4) of \mathrm{SL}(2, \mathbb{Z}), transforming as \theta\left(\frac{a\tau + b}{c\tau + d}\right) = \chi(\gamma) (c\tau + d)^{1/2} \theta(\tau) with a character \chi. Powers like \theta(\tau)^8 yield integer-weight forms related to \Delta, specifically \Delta(\tau) = (2\pi)^{-12} \theta(\tau)^8 \cdot \theta\left(\frac{\tau + 1}{4}\right)^8 \cdot \theta\left(\frac{4\tau + 1}{\tau}\right)^8, connecting lattice theta series to the full group structure. For even unimodular lattices of rank 8 (like E_8), the theta series equals E_4, providing explicit lattice-theoretic constructions.

Modular Forms for General Groups

Congruence Subgroups and Modular Curves

Congruence subgroups are specific finite-index subgroups of the modular group \mathrm{SL}_2(\mathbb{Z}) defined by arithmetic conditions. A subgroup \Gamma \leq \mathrm{SL}_2(\mathbb{Z}) is a congruence subgroup if it contains the principal congruence subgroup \Gamma(N) for some positive integer N, where \Gamma(N) is the kernel of the natural reduction map \mathrm{SL}_2(\mathbb{Z}) \to \mathrm{SL}_2(\mathbb{Z}/N\mathbb{Z}), consisting of matrices congruent to the identity modulo N. The integer N is called the level of \Gamma, and the smallest such N defines the minimal level. These subgroups play a central role in the theory of modular forms because they encode level structure in the associated moduli problems for elliptic curves. Prominent examples include the principal congruence subgroups \Gamma(N), the Hecke subgroups \Gamma_0(N) = \{ \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \mathrm{SL}_2(\mathbb{Z}) \mid c \equiv 0 \pmod{N} \}, and \Gamma_1(N) = \{ \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \mathrm{SL}_2(\mathbb{Z}) \mid a \equiv d \equiv 1 \pmod{N}, c \equiv 0 \pmod{N} \}. The index [\mathrm{SL}_2(\mathbb{Z}) : \Gamma(N)] = N^3 \prod_{p \mid N} (1 - 1/p^2) grows with N, reflecting the increasing complexity of the quotient spaces. All these subgroups have finite index in \mathrm{SL}_2(\mathbb{Z}), ensuring that the associated quotient spaces are finite covers of the moduli space of elliptic curves. Modular curves arise as quotients of the upper half-plane \mathbb{H} by the action of . For a \Gamma, the affine modular curve Y(\Gamma) = \Gamma \backslash \mathbb{H} is a obtained by identifying points under the fractional linear transformations induced by \Gamma, and it is Hausdorff in the . To compactify, one adjoins the cusps by extending to the extended upper half-plane \mathbb{H}^* = \mathbb{H} \cup \mathbb{Q} \cup \{\infty\} and forming X(\Gamma) = \Gamma \backslash \mathbb{H}^*, which adds finitely many points corresponding to rational slopes at infinity. This X(\Gamma) is a smooth compact of g(\Gamma), with the number of cusps and elliptic fixed points determining its . For instance, X(\Gamma_0(4)) has three cusps at representatives {{grok:render&&&type=render_inline_citation&&&citation_id=0&&&citation_type=wikipedia}}, [1/2], and [\infty]. These modular curves parametrize s with additional level-N structure. Specifically, Y_0(N) classifies pairs (E, C) where E is an over \mathbb{C} and C \subset E[N] is a cyclic of N, while Y_1(N) parametrizes pairs (E, Q) with Q \in E[N] a N, and Y(N) handles full level-N structures via bases for E[N] compatible with the Weil . The g(X(\Gamma)) can be computed via the formula involving the index, number of cusps v_\infty, and elliptic points: for \Gamma = \Gamma_0(N) or similar, it starts at 0 for small N (e.g., g(X_0(11)) = 1) and grows roughly as (\mu/12) - 1, where \mu = [\mathrm{SL}_2(\mathbb{Z}) : \Gamma]. The connection to modular forms is profound: holomorphic modular forms of weight $2k for \Gamma correspond bijectively to holomorphic k-fold differentials on X(\Gamma), via the map f(\tau) (d\tau)^k \mapsto \omega on the quotient. The dimension of the space of such forms is given by \dim M_{2k}(\Gamma) = (2k-1)(g-1) + k v_\infty + \left\lfloor \frac{k}{2} \right\rfloor \epsilon_2 + \left\lfloor \frac{2k}{3} \right\rfloor \epsilon_3 for k > 0, where \epsilon_2, \epsilon_3 count fixed points of order 2 and 3. This links the analytic properties of modular forms directly to the geometry of the modular curve.

General Definition via Automorphy

A modular form of k \in \mathbb{Z} for a \Gamma \subseteq \mathrm{SL}_2(\mathbb{Z}) is a f: \mathbb{H} \to \mathbb{C} on the upper half-plane \mathbb{H} that satisfies the automorphy condition f(\gamma z) = j(\gamma, z)^k f(z) for all \gamma = \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \Gamma and z \in \mathbb{H}, where the automorphy factor is j(\gamma, z) = cz + d. This transformation law ensures that f is invariant up to the specified factor under the action of \Gamma on \mathbb{H} via transformations \gamma z = \frac{az + b}{cz + d}. The automorphy factor j(\gamma, z) satisfies the cocycle relation j(\gamma_1 \gamma_2, z) = j(\gamma_1, \gamma_2 z) j(\gamma_2, z) for \gamma_1, \gamma_2 \in [\mathrm{SL}](/page/SL)_2(\mathbb{R}), which guarantees the consistency of the transformation law under group composition. For the full \Gamma = \mathrm{SL}_2(\mathbb{Z}), this yields the classical definition, but the framework extends naturally to subgroups such as \Gamma_0(N) = \left\{ \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \mathrm{SL}_2(\mathbb{Z}) \mid c \equiv 0 \pmod{N} \right\} or \Gamma_1(N). To incorporate multiplier systems or Dirichlet characters, the definition generalizes to weakly holomorphic modular forms with a character \chi: \Gamma \to \mathbb{C}^\times, where the condition becomes f(\gamma z) = \chi(\gamma) j(\gamma, z)^k f(z) for \gamma \in \Gamma, ensuring \chi is a compatible with the automorphy factor. The slash operator formalizes this via (f|_k \gamma)(z) = j(\gamma, z)^{-k} f(\gamma z), so the automorphy condition is equivalent to f|_k \gamma = f for all \gamma \in \Gamma. For \Gamma of finite index in \mathrm{SL}_2(\mathbb{Z}), modular forms must also be holomorphic at the cusps of \Gamma \backslash \mathbb{H}, meaning f|_k \sigma has a Fourier expansion with non-negative powers at infinity for a suitable \sigma \in \mathrm{SL}_2(\mathbb{Z}) mapping cusps to \infty. This automorphic perspective unifies the analytic across levels, facilitating the study of spaces M_k(\Gamma) as finite-dimensional \mathbb{C}- spaces, with dimensions computable via Riemann-Roch on the modular X(\Gamma). For non- subgroups, the requires additional like multiplier ideals, but cases suffice for most applications.

Line Bundles and Sheaf Cohomology

In the geometric framework, modular curves provide a natural setting for interpreting modular forms via . The modular curve X_\Gamma associated to a \Gamma \leq \mathrm{[SL](/page/SL)}_2(\mathbb{Z}) is a compact , and the universal elliptic curve \mathcal{E} \to X_\Gamma over it has a relative dualizing sheaf \omega_{\mathcal{E}/X_\Gamma}, often denoted simply as \omega. This \omega, known as the Hodge line bundle, is the of the sheaf of relative differentials \Omega^1_{\mathcal{E}/X_\Gamma}, and its fiber over a point corresponding to an E is isomorphic to H^0(E, \Omega^1_E). Modular forms of weight $2k for \Gamma over \mathbb{C} are precisely the global holomorphic sections of the k-th tensor power \omega^{\otimes k} on the open modular curve Y_\Gamma = X_\Gamma \setminus (cusps and elliptic points), extending holomorphically to X_\Gamma. More formally, the space of modular forms M_{2k}(\Gamma, \mathbb{C}) is isomorphic to H^0(X_\Gamma, \omega^{\otimes k}), where the isomorphism arises from the analytic uniformization X_\Gamma \cong \Gamma \backslash \mathbb{H}^* and the automorphy factor (cz + d)^{2k}. This unifies the classical analytic definition with the algebro-geometric one, allowing modular forms to be viewed as automorphic sections of line bundles on stacks of elliptic curves with level structure. Sheaf cohomology enters crucially in computing the dimensions of these spaces and understanding their properties. The dimension \dim H^0(X_\Gamma, \omega^{\otimes k}) for the corresponding weight $2k is given by (2k-1)(g-1) + k v_\infty + \left\lfloor \frac{k}{2} \right\rfloor \epsilon_2 + \left\lfloor \frac{2k}{3} \right\rfloor \epsilon_3, derived from the Riemann-Roch theorem applied to the line bundle \omega^{\otimes k} and accounting for the geometry of cusps and elliptic points. Higher cohomology groups H^i(X_\Gamma, \omega^{\otimes k}) vanish for i > 0 and sufficiently large k under suitable conditions on \Gamma (e.g., \Gamma( N) for N \geq 3), ensuring that the space of sections is finite-dimensional and computable via topological invariants of X_\Gamma. This cohomological framework also facilitates the study of modular forms over rings of integers, where base change theorems relate analytic and algebraic cohomology. The \omega is ample on X_\Gamma, generating the in many cases, and its powers encode the ring of modular forms. For instance, over the full \Gamma = \mathrm{[SL](/page/SL)}_2(\mathbb{[Z](/page/Z)}), \omega^{\otimes 12} \cong \mathcal{O}_{X(1)}(\mathrm{[pt](/page/pt)}) via the , linking sheaf sections to the modular invariant j. In higher levels, the on the moduli of elliptic curves with \Gamma- provides a , with computations yielding q-expansions and Hecke actions.

Transformation Properties and Consequences

The transformation property of a modular form f of weight k for a \Gamma \leq \mathrm{SL}_2(\mathbb{Z}) requires that f is a on the upper half-plane \mathbb{H} satisfying f(\gamma z) = j(\gamma, z)^k f(z) for all \gamma = \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \Gamma, where the automorphy factor is j(\gamma, z) = cz + d. This condition ensures that f is invariant under the action of \Gamma up to the weight factor, allowing f to descend to a well-defined function on the quotient \Gamma \backslash \mathbb{H}. For non-trivial Dirichlet characters \chi, the transformation can be generalized to include a multiplier: f(\gamma z) = \chi(d) j(\gamma, z)^k f(z), which incorporates nebentypus and enriches the theory by connecting to . The automorphy factor j(\gamma, z) is normalized such that |j(\gamma, z)| = 1 for z \in \mathbb{H}, preserving the holomorphy of f under group actions. This setup extends the classical case for \mathrm{SL}_2(\mathbb{Z}) to arbitrary congruence subgroups, where \Gamma contains \Gamma(N) for some N \geq 1. A key consequence is the existence of Fourier expansions at cusps. For a cusp c represented by \sigma \in \mathrm{SL}_2(\mathbb{Z}), the width h_\Gamma(c) determines the expansion f|_\sigma(z) = \sum_{n \geq 0} a_n q^{n/h_\Gamma(c)} with q = e^{2\pi i z}, where holomorphy at c requires non-negative powers. The transformation property implies that the space M_k(\Gamma) of such forms is finite-dimensional, with dimension bounded by the index [\mathrm{SL}_2(\mathbb{Z}) : \Gamma]. The valence formula quantifies this dimension precisely: for even k \geq 0, \dim M_k(\Gamma) = \frac{k}{12} [\mathrm{SL}_2(\mathbb{Z}) : \Gamma] + \text{corrections from elliptic points and cusps}, derived from the transformation invariance and residue computations on the compactified modular curve X(\Gamma). This formula, originally due to Rademacher, yields explicit dimensions for small levels, such as \dim M_2(\Gamma_0(4)) = 1. Further implications include the invariance of the Petersson inner product \langle f, g \rangle = \int_{\Gamma \backslash \mathbb{H}} |f(z)|^2 y^{k-2} \frac{dx dy}{y^2}, which is preserved under the and enables relations for Hecke eigenforms.

Classification and Special Types

Cusp Forms

Cusp forms constitute a distinguished of the space of modular forms, characterized by their vanishing behavior at the cusps of the associated . For the full \Gamma = \mathrm{SL}(2, \mathbb{Z}), a cusp form of weight k is a modular form f \in M_k(\Gamma) whose expansion at the cusp \infty, given by f(\tau) = \sum_{n=0}^\infty a_n e^{2\pi i n \tau} for \mathrm{Im}(\tau) > 0, has vanishing a_0 = 0. This condition ensures that f extends holomorphically to the cusp \infty with a zero there, reflecting the geometric interpretation as sections of line bundles on the compactified X(1) = \mathbb{H}/\Gamma \cup \{\infty\} that vanish at \infty. The space of cusp forms S_k(\Gamma) forms a finite-dimensional , with dimension given by \dim S_k(\mathrm{SL}(2, \mathbb{Z})) = \dim M_k(\mathrm{SL}(2, \mathbb{Z})) - 1 for even k \geq 4, where \dim M_k(\mathrm{SL}(2, \mathbb{Z})) = \lfloor k/12 \rfloor + 1 if k \not\equiv 2 \pmod{12} and \lfloor k/12 \rfloor otherwise. Explicitly, \dim S_k(\mathrm{SL}(2, \mathbb{Z})) = 0 for $2 \leq k < 12 even, \dim S_{12} = 1, and the dimensions grow asymptotically as k/12. A canonical generator of S_{12}(\mathrm{SL}(2, \mathbb{Z})) is the Ramanujan discriminant function \Delta(\tau) = q \prod_{n=1}^\infty (1 - q^n)^{24} with q = e^{2\pi i \tau}, whose coefficients \tau(n) satisfy the Ramanujan congruence \tau(p) \equiv \sigma_{11}(p) \pmod{691} for primes p. This form plays a pivotal role in the valence formula, which relates the orders of zeros of modular forms to their weight: for f \in M_k(\Gamma), \sum_{z \in \mathbb{H}/\Gamma} \mathrm{ord}_z(f) + \frac{1}{2} \mathrm{ord}_i(f) + \frac{1}{3} \mathrm{ord}_\omega(f) + \mathrm{ord}_\infty(f) = k/12, and cusp forms achieve equality only through their cusp zeros. For a general congruence subgroup \Gamma of finite index in \mathrm{SL}(2, \mathbb{Z}), such as \Gamma_0(N) or \Gamma_1(N), the cusps form a finite set \mathrm{Cusps}(\Gamma) = \mathbb{P}^1(\mathbb{Q}) / \Gamma, and a cusp form f \in M_k(\Gamma, \chi) (with Nebentypus character \chi) is defined by requiring that the constant term vanishes in the Fourier expansion at every cusp. To obtain the expansion at a cusp s = a/c \in \mathbb{Q} \cup \{\infty\}, one applies a suitable Atkin-Lehner or slash operator to translate s to \infty, yielding f|_{k} \sigma (\tau) = \sum_{n \gg 0} a_n(s) e^{2\pi i n \tau / w} for some width w > 0, and cusp forms satisfy a_0(s) = 0 for all s \in \mathrm{Cusps}(\Gamma). The space S_k(\Gamma, \chi) decomposes orthogonally as M_k(\Gamma, \chi) = S_k(\Gamma, \chi) \oplus E_k(\Gamma, \chi), where E_k is the Eisenstein subspace, enabling the under Hecke operators. Cusp forms exhibit strong analytic properties, including bounded growth |a_n| = O(n^{k/2 - 1/2 + \epsilon}) for \epsilon > 0 by Deligne's theorem, resolving Ramanujan's conjecture for \Delta. They are central to the Langlands program, as normalized Hecke eigenforms in S_k(\Gamma_0(N), \chi) (newforms) correspond to irreducible cuspidal automorphic representations of \mathrm{GL}_2(\mathbb{A}_\mathbb{Q}), with associated L-functions satisfying the Riemann hypothesis. For instance, the space S_2(\Gamma_0(11)) is one-dimensional, spanned by a newform whose L-function encodes the class number of \mathbb{Q}(\sqrt{-11}).

Eisenstein Series

Eisenstein series provide classical examples of holomorphic modular forms for the full \mathrm{[SL](/page/SL)}_2(\mathbb{[Z](/page/Z)}), distinguished by their explicit construction as sums and their role in spanning the space of all such forms. For an even k \geq 4, the of weight k is defined by G_k(\tau) = \sum_{(m,n) \in \mathbb{Z}^2 \setminus \{(0,0)\}} \frac{1}{(m + n \tau)^k}, where \tau lies in the upper half-plane \mathbb{H}. This series converges absolutely for k > 2, ensuring holomorphicity on \mathbb{H}. The function G_k transforms as a modular form of weight k: for any \gamma = \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \mathrm{SL}_2(\mathbb{Z}), it satisfies G_k(\gamma \tau) = (c \tau + d)^k G_k(\tau). This automorphy property follows from the invariance of the under the action of \mathrm{SL}_2(\mathbb{Z}) and the homogeneity of the summand. At the cusp \infty, G_k extends holomorphically with constant term $2 \zeta(k), where \zeta is the , confirming it is a modular form but not a cusp form due to the nonzero constant. For odd k, G_k \equiv 0 by antisymmetry. The Fourier expansion of G_k at infinity is G_k(\tau) = 2 \zeta(k) + \frac{2 (-1)^{k/2} (2 \pi)^k}{(k-1)!} \sum_{n=1}^\infty \sigma_{k-1}(n) q^n, where q = e^{2 \pi i \tau} and \sigma_{k-1}(n) = \sum_{d \mid n} d^{k-1} is the sum-of-divisors function. The normalized Eisenstein series E_k(\tau) = G_k(\tau) / (2 \zeta(k)) then has leading term 1 and coefficients involving Bernoulli numbers B_k: E_k(\tau) = 1 - \frac{2k}{B_k} \sum_{n=1}^\infty \sigma_{k-1}(n) q^n. This normalization highlights the connection to arithmetic data, as the coefficients encode divisor sums. Eisenstein series generate the ring of modular forms for \mathrm{SL}_2(\mathbb{Z}): the space M_k(\mathrm{SL}_2(\mathbb{Z})) for even k \geq 4 is spanned by E_4^{k/4} and E_6^{k/6} (suitably adjusted for dimensions), and the full ring is \mathbb{C}[E_4, E_6]. For instance, the cusp form \Delta(\tau) of weight 12, whose zero at infinity gives the q-expansion of the partition function, is \Delta(\tau) = (E_4^3 - E_6^2)/(1728). These forms also underpin Hecke operators, where E_k are eigenforms with eigenvalues \sigma_{k-1}(n).

Newforms and Hecke Eigenforms

Hecke operators T_n act on the space of modular forms M_k(\Gamma_0(N)) of weight k and level N by summing over cosets in a decomposition, preserving the space and commuting with each other. A Hecke eigenform is a nonzero modular form f \in M_k(\Gamma_0(N)) that is a simultaneous eigenvector for all Hecke operators, satisfying T_n f = \lambda_n f for each positive n, where the eigenvalues \lambda_n coincide with the n-th coefficient a_n(f) of f. This concept was introduced by Erich Hecke in his foundational work on modular functions and , where he showed that such eigenforms form a basis for the space of modular forms under the full \mathrm{SL}_2(\mathbb{Z}). The coefficients of a Hecke eigenform are multiplicative, meaning a_{mn}(f) = a_m(f) a_n(f) whenever \gcd(m,n) = 1, which enables the associated [L-function](/page/L-function) L(f,s) = \sum_{n=1}^\infty a_n(f) n^{-s} to admit an Euler product decomposition L(f,s) = \prod_p (1 - a_p(f) p^{-s} + p^{k-1} p^{-2s})^{-1} over primes p. For cusp forms, the eigenvalues satisfy the Ramanujan-Petersson conjecture, bounding |a_p(f)| \leq 2 p^{(k-1)/2}, proven by Deligne in using . Hecke eigenforms thus play a central role in , linking modular forms to [L-functions](/page/L-function) with arithmetic significance, such as those attached to elliptic curves via the . When considering modular forms for congruence subgroups like \Gamma_0(N), the space M_k(\Gamma_0(N)) decomposes into "oldforms" and "newforms." Oldforms arise from forms of lower level d \mid N via operators V_d and U_d, which embed and average Fourier coefficients, respectively; specifically, for g \in M_k(\Gamma_0(d)), the induced form g \mid V_d has coefficients a_n(g \mid V_d) = a_{n/d}(g) if d \mid n and 0 otherwise. This decomposition, orthogonal with respect to the Petersson inner product \langle f, g \rangle = \int_{\Gamma_0(N) \backslash \mathbb{H}} |f(z)|^2 y^{k-2} \, dx \, dy, was established by Atkin and Lehner to isolate forms genuinely associated to level N. A newform is a normalized Hecke eigenform f (with a_1(f) = 1) belonging to the new subspace M_k^{\mathrm{new}}(\Gamma_0(N)), the orthogonal complement to the oldforms. Newforms form an orthonormal basis for the cusp form new subspace under the Petersson product and are eigenvectors for the Atkin-Lehner operators W_Q (for divisors Q \mid N with \gcd(Q, N/Q)=1), which act as involutions interchanging cusps. Their L-functions have no "missing" Euler factors at primes dividing N, ensuring full multiplicativity and enabling precise control over arithmetic properties, such as the conductor in the Langlands program. For example, the discriminant modular form \Delta(z) = q \prod_{n=1}^\infty (1 - q^n)^{24} is the unique newform of weight 12 and level 1.

Algebraic Structure

Ring of Modular Forms

The ring of modular forms for a subgroup \Gamma \subset \mathrm{SL}(2, \mathbb{Z}) of finite index is the M_*(\Gamma) = \bigoplus_{k \geq 0} M_k(\Gamma), where M_k(\Gamma) denotes the \mathbb{C}- of modular forms of weight k for \Gamma. Addition is defined componentwise, and multiplication is induced by the of functions on the upper half-plane, which preserves the space of modular forms since the transformation laws multiply appropriately under the . For the full modular group \Gamma = \mathrm{SL}(2, \mathbb{Z}), this ring has a particularly simple structure: it is freely generated over \mathbb{C} by the E_4 and E_6 of weights 4 and 6, respectively. That is, M_*(\mathrm{SL}(2, \mathbb{Z})) \cong \mathbb{C}[E_4, E_6] as graded rings, where the grading is induced by (with \deg E_4 = 4 and \deg E_6 = 6). Every modular form of k (even, as odd weights vanish) is thus a unique \mathbb{C}-linear combination of monomials E_4^a E_6^b such that $4a + 6b = k. This isomorphism follows from the dimension formula: \dim M_k(\mathrm{SL}(2, \mathbb{Z})) = 0 if k or k < 0, \dim M_0 = 1, \dim M_2 = 0, and for even k > 2, \dim M_k = \lfloor k/12 \rfloor + 1 if k \not\equiv 2 \pmod{12}, \dim M_k = \lfloor k/12 \rfloor if k \equiv 2 \pmod{12}, which matches the number of such monomials, combined with the fact that the span M_k via their explicit expansions and the valence formula. The subspace of cusp forms S_*(\mathrm{SL}(2, \mathbb{Z})) = \bigoplus_{k \geq 0} S_k(\mathrm{SL}(2, \mathbb{Z})) forms a graded ideal in M_*(\mathrm{SL}(2, \mathbb{Z})), generated by the cusp form \Delta of weight 12, known as the . Explicitly, \Delta = \frac{1}{1728}(E_4^3 - E_6^2), and it has a simple Fourier expansion \Delta(\tau) = q \prod_{n=1}^\infty (1 - q^n)^{24} (with q = e^{2\pi i \tau}), ensuring no . Consequently, M_*(\mathrm{SL}(2, \mathbb{Z})) \cong \mathbb{C}[E_4, E_6, \Delta] when including cusp forms explicitly, though the polynomial presentation in E_4 and E_6 already incorporates \Delta via the relation E_4^3 - E_6^2 = 1728 \Delta. The ring is thus Noetherian and finitely generated, with the cusp forms comprising all forms vanishing at the cusp \infty. For general congruence subgroups \Gamma of level N, the ring M_*(\Gamma) remains finitely generated as a \mathbb{C}-algebra, but the presentation is more involved, typically requiring generators in weights up to 6 (or lower if M_3(\Gamma) \neq 0) and relations in weights up to 12. For example, for \Gamma(2), the principal congruence subgroup of level 2, the ring is freely generated by the theta series \Theta_2^4 and \Theta_3^4 (or equivalently, Eisenstein series of weight 2). In general, the structure reflects the geometry of the modular curve X(\Gamma), with the ring canonically isomorphic to the ring of global sections of powers of the canonical bundle on X(\Gamma). These presentations have been explicitly computed for small levels using Gröbner bases and algebro-geometric methods generalizing classical results of Noether and Petri.

Generators and Modularity Theorem

This polynomial presentation highlights the algebraic simplicity of level 1 modular forms, where higher-weight forms arise from products and powers of these basic . For instance, the Eisenstein series E_8 = \frac{1}{45} (E_4^2 + 7 E_6 E_4) and E_{12} = \frac{1}{660} (E_4^3 + 75 E_6^2 E_4 - 81 E_6^3) are explicitly constructed from E_4 and E_6, underscoring the generative role of these weight-4 and weight-6 forms. Over the integers, the ring M_*(\mathrm{SL}(2,\mathbb{Z})) is generated by E_4, E_6, and the \Delta, but the complex coefficients allow the free generation by just two elements. The modularity theorem, formerly known as the Taniyama-Shimura-Weil conjecture, asserts that every elliptic curve E over the rational numbers \mathbb{Q} is modular, meaning there exists a cuspidal newform f of weight 2 and level equal to the conductor N of E such that the L-function of E coincides with the L-function of f. Conjectured in the 1950s by Yutaka Taniyama and further developed by Goro Shimura and André Weil, the theorem establishes a deep correspondence between elliptic curves and modular forms, implying that the Fourier coefficients of f match the coefficients of the L-series of E. The full proof was completed in 2001 by Christophe Breuil, Brian Conrad, Fred Diamond, and Richard Taylor, building on partial results by Andrew Wiles and Wiles with Taylor that covered semistable elliptic curves. This theorem has profound implications for , notably providing a proof of as a , since it links the arithmetic of s to the analytic properties of modular forms via Galois representations. Specifically, for a non-CM elliptic curve E/\mathbb{Q}, the associated 2-dimensional Galois representation \rho_{E,\ell} : \mathrm{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) \to \mathrm{GL}_2(\mathbb{F}_\ell) is irreducible and matches that of the modular form f, ensuring the modularity linkage. The result extends the in this setting, confirming that elliptic curves over \mathbb{Q} parametrize modular forms of weight 2.

Hecke Operators and L-Functions

Hecke operators are linear endomorphisms T_n (for positive integers n) on the space M_k(\Gamma) of modular forms of weight k for a \Gamma, such as \Gamma = \mathrm{SL}_2(\mathbb{Z}) or \Gamma_0(N). They were introduced by Erich Hecke in to study the properties of modular forms through their multiplicative structure on coefficients. For a modular form f(z) = \sum_{m=0}^\infty c(m) q^m with q = e^{2\pi i z}, the action of T_n produces another modular form whose m-th coefficient is \sum_{d \mid \gcd(n,m)} d^{k-1} c\left( \frac{nm}{d^2} \right). This formula reveals the multiplicative nature of the operators, as T_m T_n = \sum_{d \mid \gcd(m,n)} d^{k-1} T_{mn/d^2}, making the Hecke operators \{ T_n \} generate a known as the . The acts diagonally on a basis of simultaneous eigenforms, which are normalized so that the constant term c(1) = 1. A Hecke eigenform f satisfies T_n f = \lambda_n f for eigenvalues \lambda_n, and these eigenvalues coincide with the coefficients \lambda_n = c(n). In the of cusp forms S_k(\Gamma_0(N)), the newforms—those eigenforms orthogonal to forms induced from smaller levels—form an with respect to the Petersson inner product. For prime p \nmid N, the eigenvalues satisfy the Ramanujan bound |\lambda_p| \leq 2 p^{(k-1)/2}, ensuring convergence properties essential for associated analytic objects. To each normalized Hecke eigenform f = \sum_{n=1}^\infty \lambda_n q^n of weight k and level N, one associates an L-function L(f, s) = \sum_{n=1}^\infty \frac{\lambda_n}{n^s}, which converges absolutely for \mathrm{Re}(s) > (k+1)/2. This Dirichlet series admits an Euler product L(f, s) = \prod_p L_p(f, s)^{-1}, where for primes p \nmid N, the local factor is L_p(f, s) = 1 - \lambda_p p^{-s} + p^{k-1-2s}, reflecting the Hecke eigenvalue relations. Hecke proved that L(f, s) extends to a holomorphic function on the entire complex plane, satisfying a functional equation \tilde{L}(f, s) = \epsilon N^{s/2} (2\pi)^{-s} \Gamma(s) L(f, s) = \pm \tilde{L}(f, k - s), where \epsilon is a root number of absolute value 1. These L-functions encode arithmetic data, such as special values at integers linking to Birch and Swinnerton-Dyer conjectures for elliptic curves via modularity.

Historical Context

Origins in Complex Analysis

The study of modular forms traces its origins to the early 19th-century investigation of elliptic functions and integrals within complex analysis. Carl Friedrich Gauss laid foundational groundwork around 1800 through his work on the arithmetic-geometric mean (AGM), where he derived expressions for complete elliptic integrals that implicitly involve modular functions of level 4; specifically, the AGM of 1 and \sqrt{2} relates to the elliptic modulus k = \frac{\sqrt{2}-1}{\sqrt{2}+1}, yielding hypergeometric series that are precursors to modular forms. Carl Gustav Jacobi further advanced this in the 1820s with his theta functions, such as \vartheta_3(z \mid \tau) = \sum_{n=-\infty}^{\infty} e^{2\pi i n^2 \tau + 2\pi i n z}, which exhibit transformation properties under the action of the modular group \mathrm{SL}(2, \mathbb{Z}) on the upper half-plane \mathbb{H}, though initially studied for their role in elliptic function theory rather than modular invariance. Bernhard Riemann's 1857 habilitation lecture, "Theorie der Abel'schen Functionen," provided a geometric by associating elliptic curves with complex tori \mathbb{C}/\Lambda, where \Lambda is a generated by periods \omega_1, \omega_2. He introduced the period matrix and the of the of elliptic curves, parameterized by the tau \tau \in \mathbb{H}, which is invariant under \mathrm{SL}(2, \mathbb{Z}) actions; this perspective linked analytic properties of functions on \mathbb{H} to the of Riemann surfaces, setting the for modular forms as holomorphic sections on these spaces. In the 1880s, Henri Poincaré independently discovered automorphic functions through his work on Fuchsian groups, detailed in his 1882 paper "Sur les fonctions fuchsiennes" published in Acta Mathematica. He constructed series expansions, now known as Poincaré series, such as \sum_{\gamma \in \Gamma_\infty \backslash \Gamma} (Im(\gamma z))^k |cz + d|^{-2k} f(\gamma z) for a discrete group \Gamma \subset \mathrm{SL}(2, \mathbb{R}) acting on \mathbb{H}, showing that certain holomorphic functions remain invariant under \Gamma; modular forms emerged as the special case where \Gamma = \mathrm{SL}(2, \mathbb{Z}), with weight k \geq 2 ensuring holomorphy at the cusps. Concurrently, Felix Klein developed the theory of modular functions in his 1879 paper "Über die Transformationen der elliptischen Funktionen" and subsequent works, emphasizing their role in solving the icosahedral equation and classifying Riemann surfaces for congruence subgroups. Klein's collaboration with Robert Fricke culminated in the multi-volume "Vorlesungen über die Theorie der elliptischen Modulfunktionen" (1890–1892), where they systematically classified modular functions and introduced the term "Modulform" to describe holomorphic functions on the modular curve X(1) = \mathbb{H}/\mathrm{SL}(2, \mathbb{Z}) with specified transformation laws, such as f\left(-\frac{1}{\tau}\right) = \tau^k f(\tau) for integer weight k. This work integrated Riemann's surfaces with Poincaré's automorphic constructions, establishing modular forms as central objects in and foreshadowing their number-theoretic applications.

Key Developments in Number Theory

The study of modular forms gained prominence in through Srinivasa Ramanujan's early 20th-century investigations into functions and related arithmetic series. In his 1916 paper, Ramanujan introduced the discriminant modular form \Delta(z) = \eta(z)^{24}, where \eta(z) is the , and defined its Fourier coefficients \tau(n) as the , establishing key identities such as the \tau(p) \equiv \sigma_{11}(p) \pmod{691} for primes p. These results foreshadowed deep connections between modular forms and multiplicative functions, with Ramanujan's conjectures on the growth of \tau(n) later proven by Deligne in 1974 as a consequence of the . Erich Hecke's work in systematized these ideas by developing a general theory linking modular forms to and L-functions. In his seminal 1936 paper, Hecke defined Hecke operators on spaces of modular forms and showed that eigenforms under these operators have Euler product expansions for their L-functions, analogous to the . This framework, detailed in Hecke's 1938 lectures, established modular forms as a bridge between and algebraic structures, enabling the study of their arithmetic properties through . A pivotal advancement came in the mid-20th century with the Taniyama-Shimura conjecture, proposed orally by in 1955 and formalized by Goro Shimura in 1958, positing that every over the rationals is modular, meaning it corresponds to a weight-2 cusp form. This conjecture, refined by , implied profound arithmetic consequences, including the modularity of s' L-functions. Its partial proof by in 1995, published in his Annals paper, proved that every semistable over the rationals is modular and resolved by linking it to the Frey curve's non-modularity under the conjecture. The full was established in 2001 by Breuil, Conrad, , and , confirming the conjecture for all s and unlocking applications in Galois representations and arithmetic geometry.

Extensions and Generalizations

Maass Forms

Maass forms, also known as Maass cusp forms or non-holomorphic modular forms, are real-analytic functions on the upper half-plane \mathbb{H} that generalize classical holomorphic modular forms by relaxing the holomorphy condition while preserving automorphic properties under the action of subgroups of \mathrm{[SL](/page/SL)}_2(\mathbb{Z}). Introduced by Hans Maass in as a class of non-analytic automorphic functions satisfying certain transformation laws and differential equations, they play a central role in the of automorphic forms and the study of L-functions associated to number-theoretic objects. Unlike holomorphic modular forms, which are eigenfunctions of the Cauchy-Riemann operator, Maass forms are eigenfunctions of the hyperbolic Laplace-Beltrami operator, enabling connections to the geometry of hyperbolic surfaces and random matrix theory. Formally, a Maass form f: \mathbb{H} \to \mathbb{C} for a \Gamma \subset \mathrm{[PSL](/page/PSL)}_2(\mathbb{R}), such as \Gamma_0(N), is a smooth function satisfying the automorphy condition f(\gamma z) = j(\gamma, z)^k f(z) for \gamma \in \Gamma, where j(\gamma, z) is the automorphic factor and k \in \mathbb{R} is the weight (often k=0 for the classical case). It is an of the weight-k Laplacian \Delta_k = -y^2 \left( \frac{\partial^2}{\partial x^2} + \frac{\partial^2}{\partial y^2} \right) + i k y \frac{\partial}{\partial x} with eigenvalue \lambda = s(1-s), where s = \frac{1}{2} + i R and R \in \mathbb{R} is the spectral parameter. For cusp forms, f must decay exponentially at the cusps of \Gamma \backslash \mathbb{H}, ensuring square-integrability in L^2(\Gamma \backslash \mathbb{H}). The expansion of an even Maass form f(z) = f(x + i y) at the cusp \infty takes the form f(z) = \sum_{n \neq 0} \rho_n \sqrt{y} K_{i R}(2 \pi |n| y) \cos(2 \pi n x + \phi_n), where K_{i R} is the modified of the third kind, \rho_n are the coefficients, and \phi_n is a (often $0 or \pi/2 for even/ forms). This expansion reflects the Maass form's behavior as a superposition of hyperbolic waves, contrasting with the q-expansions of holomorphic forms. Maass forms that are simultaneous eigenfunctions of all Hecke operators T_p (for primes p) are called Hecke-Maass forms, with eigenvalues \lambda_p satisfying the Ramanujan-Petersson conjecture \lvert \lambda_p \rvert \leq 2, proven by and in 2002 for \mathrm{GL}_2.[](https://www.math.purdue.edu/~fshahidi/articles/Kim & Shahidi [2002, 57pp]---Functorial products for GL_2 x GL_3 and the symmetric cube for GL_2.pdf) The Atkin-Lehner-Li theory extends the notion of newforms to Maass forms, decomposing the space into new and old components based on the level N. Associated to a Hecke-Maass form is the L-function L(s, f) = \sum_{n=1}^\infty \frac{\lambda_n}{n^{s + 1/2}}, which admits analytic continuation to the complex plane and a functional equation of the form \Lambda(s, f) = \epsilon N^{1/2 - s} \Lambda(1 - s, \tilde{f}), where \Lambda(s, f) = (2\pi)^{-s} \Gamma\left(s + \frac{i R}{2}\right) \Gamma\left(s - \frac{i R}{2}\right) L(s, f) for even forms, \epsilon = \pm i^k is the root number, and \tilde{f} is the form at the dual cusp. This mirrors the properties of holomorphic newform L-functions but incorporates the non-holomorphic spectral parameter R, linking to the continuous spectrum of the Laplacian. Maass forms contribute to the discrete spectrum of L^2(\Gamma_0(N) \backslash \mathbb{H}), with their multiplicities and distribution studied via Selberg's trace formula, providing tools for bounding primes in arithmetic progressions and moments of L-functions. In applications, Maass forms underpin the for \mathrm{GL}_2/\mathbb{Q}, where they correspond to irreducible cuspidal representations, and their coefficients appear in moments of zeta functions, as in the work of Conrey, , and Zirnbauer on analogies. Numerical computations, such as those for the first Maass form on \Gamma_0(1) with R \approx 9.533, confirm eigenvalue spacings predicted by the Gaussian Unitary Ensemble.

Hilbert and Siegel Modular Forms

Hilbert modular forms generalize classical modular forms to totally real number fields. For a totally real algebraic number field F of degree g over \mathbb{Q}, with ring of integers \mathcal{O}_F, the Hilbert modular group is \Gamma_F = \mathrm{SL}(2, \mathcal{O}_F), acting on the product of g upper half-planes H^g. A Hilbert modular form of parallel weight k \in \mathbb{Z} and level \Gamma \subseteq \Gamma_F is a holomorphic function f: H^g \to \mathbb{C} satisfying the automorphy condition f(\gamma z) = \left( \prod_{\sigma: F \hookrightarrow \mathbb{R}} (c^\sigma z_\sigma + d^\sigma) \right)^k f(z) for \gamma \in \Gamma, along with suitable growth conditions at the cusps. These forms admit Fourier expansions f(z) = \sum_{\mu} a_\mu \exp(2\pi i \mathrm{Tr}(\mu z)), where the sum is over fractional ideals \mu in the inverse different, and coefficients a_\mu satisfy multiplicativity under Hecke operators. The theory originated in the early , with foundational work by on invariants of binary quadratic forms over number fields and Otto Blumenthal's 1903 thesis on multi-variable modular functions. Goro Shimura formalized the modern adelic framework in the 1970s, establishing connections to automorphic representations and s. Hilbert modular forms play a central role in arithmetic geometry, associating to each newform a motive whose L-function matches the form's Hecke L-series, and they underpin constructions like Heegner points on abelian varieties over F. Siegel modular forms extend the concept to symplectic groups, providing higher-genus analogues relevant to abelian varieties. For genus g, they are defined on the Siegel upper half-space H_g of g \times g complex symmetric matrices with positive definite imaginary part, transforming under the Siegel modular group \Gamma_g = \mathrm{Sp}(2g, \mathbb{Z}) via f((AZ + B)(CZ + D)^{-1}) = \det(CZ + D)^k f(Z) for weight k, with holomorphy and moderate growth. Cusp forms vanish under the Siegel \Phi-operator, restricting to constant terms in their Fourier expansions \sum_{T \succeq 0} a(T) \exp(2\pi i \mathrm{Tr}(T Z)), where T are positive semi-definite half-integral matrices. Introduced by in 1935 as multi-variable theta functions linked to quadratic forms, the theory advanced through Hecke operators and in works by Maass and Andrianov in the mid-20th century. modular forms of genus g parametrize principally polarized abelian varieties of dimension g, with their theta series representing lattices and L-functions encoding arithmetic data like those of motives attached to abelian varieties. Key results include the generalized Ramanujan conjecture for non-lift forms, bounding Satake parameters on the unit circle, as proven for genus 2 by Weissauer. Both Hilbert and Siegel forms fit into the Langlands program as automorphic representations of reductive groups over number fields, with Hilbert forms corresponding to \mathrm{GL}_2 over totally real fields and Siegel to \mathrm{GSp}_{2g} over \mathbb{Q}, enabling functorial lifts and Galois representation attachments via modularity theorems.

Vector-Valued and Half-Integral Weight Forms

Modular forms of half-integral weight generalize classical modular forms by allowing the weight k to be a half-integer, such as k = m + 1/2 for integer m \geq 0. These forms are defined on the metaplectic double cover of \mathrm{SL}_2(\mathbb{R}), often denoted \widetilde{\mathrm{SL}}_2(\mathbb{R}), which extends the transformation law to account for the square root in the automorphy factor. Specifically, a holomorphic function f: \mathbb{H} \to \mathbb{C} of weight k = \kappa/2 (with \kappa odd positive integer) on a congruence subgroup like \Gamma_0(4N) satisfies f\left|_\kappa \xi (\tau) = f(\tau) for \xi in the appropriate extension group G, where the slash operator incorporates a phase factor \phi(\xi, \tau) with \phi(\xi, \tau)^2 = (c\tau + d)^{-\kappa} and |\phi| = 1. Holomorphy at cusps is ensured by boundedness or polynomial growth after suitable transformations. The theory was established by Shimura in 1973, who constructed spaces of such forms, defined Hecke operators, and proved multiplicativity of Fourier coefficients for eigenforms. A prototypical example is the \theta(\tau) = \sum_{n \in \mathbb{Z}} q^{n^2} (with q = e^{2\pi i \tau}), which is a weight $1/2 cusp form on \Gamma_0(4) with trivial character; powers \theta^\kappa yield weight \kappa/2 forms. Serre and Stark showed that the space of weight $1/2 forms on \Gamma_0(4) is spanned by theta series \sum \psi(n) q^{t n^2} for even Dirichlet characters \psi of conductor dividing $4t, with Hecke eigenvalues \psi(p)(1 + p^{-1}) for primes p. Key structures include the Shimura lift, which maps a weight \kappa/2 eigenform f of level N and character \chi (with N divisible by 4) to a weight \kappa - 1 form of level N/2 and character \chi^2, preserving Hecke eigenvalues up to the Niwa relation. The Shintani lift serves as a partial inverse, associating integral weight forms to half-integral ones via theta kernels. Kohnen introduced the plus space +M_{k}(\Gamma_0(4N), \chi), a subspace where Fourier coefficients a(n) = 0 unless n \equiv 0,1 \pmod{4}, and a(4m + r) = 0 for r \equiv 2,3 \pmod{4} with sign conditions; this space admits a bijection with scalar modular forms via the Shimura correspondence restricted to newforms. Waldspurger's theorem links central L-values of quadratic twists of the Shimura lift to squares of half-integral coefficients, underpinning applications to Birch--Swinnerton-Dyer conjectures. Vector-valued modular forms extend the scalar theory by letting f: \mathbb{H} \to V map to a finite-dimensional V, transforming via a \rho: \Gamma \to \mathrm{GL}(V) as f(\gamma \tau) = (c\tau + d)^k \rho(\gamma) f(\tau) for \gamma = \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \Gamma \subseteq \mathrm{SL}_2(\mathbb{Z}) and weight k \in \mathbb{C}. The space M_k(\Gamma, \rho) consists of holomorphic such functions with growth at cusps, while weakly holomorphic variants M_k^!(\Gamma, \rho) allow poles. Historical roots trace to Poincaré's scalar forms, with Selberg advocating a representation-theoretic framework in the to unify theta series and automorphic representations. Seminal developments include Knopp and Mason's work on finite-dimensionality and Hecke algebras for unitary representations, establishing that M_k(\Gamma, \rho) is a module over the scalar ring M_k(\Gamma). For the principal congruence subgroup, the spaces form free modules of rank \dim V over \mathbb{C}[E_4, E_6], the ring generated by Eisenstein series, when \rho is unitary. Bruinier and Funke developed Hecke operators for forms valued in the Weil representation, connecting to Borcherds products and harmonic weak Maass forms. These forms arise in solutions to modular linear differential equations (MLDEs), where the monodromy group yields \rho, and hypergeometric functions parametrize examples like weight 2 forms for \mathrm{SL}_2(\mathbb{Z}). Half-integral weights often embed as vector-valued cases via the Weil representation on oscillatory theta series, linking the two generalizations.

References

  1. [1]
    [PDF] Modular forms (draft, CTNT 2016)
    Introduction. A modular form is a holomorphic function on the upper half-plane h = {x + iy : x ∈ R,y > 0} = {τ ∈ C : Imτ > 0}.
  2. [2]
    [PDF] The 1-2-3 of modular forms, by JH Bruinier, G. van der Geer, G. Harder
    Mar 23, 2009 · The origins of this definition date back to the first half of the nineteenth century, to the era of Jacobi and Eisenstein. Since then, many ...
  3. [3]
  4. [4]
    [PDF] modular forms lecture 1: introduction and motivating examples
    Modular forms are functions on the upper half plane, defined by a modularity transformation, and have nice analytic and growth conditions.
  5. [5]
    [PDF] 25 Modular forms and L-functions - MIT Mathematics
    May 15, 2017 · ... forms defined below). Definition 25.5. A modular form f is a weak modular form that is holomorphic at the cusps. Equivalently, f is a weak ...
  6. [6]
    [PDF] Modular Forms and Modular Congruences of the Partition Function
    Mar 25, 2019 · The Hecke operator is a key method that allows us to create new modular forms by modifying the q-series coefficients. While Hecke operators can ...
  7. [7]
    [PDF] Modular Functions and Modular Forms
    It is easy to define modular functions and forms, but less easy to say why they are important, especially to number theorists. Thus I will begin with a rather ...
  8. [8]
    [PDF] 14. Elliptic modular forms
    Mar 9, 2021 · Thus, more precisely, modular forms are homogeneous functions on lattices in C. [1.2] From lattices to the upper half-plane We would be ...
  9. [9]
    [PDF] Elliptic Modular Forms and Their Applications
    These notes give a brief introduction to a number of topics in the classical theory of modular forms. Some of theses topics are (planned) to be treated.
  10. [10]
    [PDF] Modular Forms - of /websites
    ... Examples of modular forms: Eisenstein series. Let k be an even integer with k ≥ 4. We define the Eisenstein series of weight k (for SL2(Z)) by. Gk : H −→ C.<|control11|><|separator|>
  11. [11]
    [PDF] Introductory Lectures on SL(2,Z) and modular forms.
    Examples of modular forms: Eisenstein series. Convergence properties and modularity. Poincaré's moment of inspiration was to see that the non-Euclidean ...
  12. [12]
    [PDF] Theta Series as Modular Forms - Zachary Abel
    The group SL2(Z) consisting of. 2 × 2 matrices over Z with determinant 1 acts on H by means of the following map: SL2(Z) ×H→H,. a b. c d ! ,z 7→ az + ...
  13. [13]
    None
    ### Summary of Congruence Subgroups, Modular Curves, and Elliptic Curves
  14. [14]
    [PDF] Modular forms and Modular curves - UChicago Math
    A modular form of weight k is a holomorphic function on the upper half complex plane that transforms under the action of SL2(Z).
  15. [15]
    A First Course in Modular Forms - SpringerLink
    In stock Free deliveryThis book introduces the theory of modular forms with an eye toward the Modularity Theorem: All rational elliptic curves arise from modular forms.
  16. [16]
    Modular Forms - Book - SpringerLink
    It sets out to provide the reader with the basic knowledge of elliptic modular forms necessary to understand the recent developments in number theory.
  17. [17]
    [PDF] 6.1 Automorphic forms Definition 6.1. The automorphy factor j
    Feb 21, 2024 · automorphic forms in A(Γ,χ) are meromorphic at the cusps of Γ, modular forms in M(Γ,χ) are holomorphic at the cusps of Γ, and cusp forms in ...<|control11|><|separator|>
  18. [18]
    [PDF] Math 248B. Modular curves
    Let us now explain the correspondence between: • classical modular forms,. • sections of holomorphic line bundles built from universal elliptic curves over ...
  19. [19]
    [PDF] Congruences Between Modular Forms.
    a meromorphic modular form of weight zero and level 1 for any p. It follows ... modular function into an infinite sum of eigenforms. The key ...
  20. [20]
    [PDF] Modular Forms
    2.5 More examples: the modular form ∆ and the modular function j . ... (c) Conclude that the space of modular forms of weight zero consists exactly of the ...<|control11|><|separator|>
  21. [21]
    Modular Forms
    In this section we define spaces of modular forms of arbitrary level. Definition 1.9 (Congruence Subgroup). A congruence subgroup of SL2(Z) is any subgroup of ...
  22. [22]
    Cusp Form -- from Wolfram MathWorld
    A cusp form is a modular form for which the coefficient c(0)=0 in the Fourier series f(tau)=sum_(n=0)^inftyc(n)e^(2piintau)
  23. [23]
    [PDF] Computing dimensions of spaces of modular forms
    We finally prove a general valence formula for any finite index subgroup Γ which we use to obtain explicit bounds on the dimension. Using these results, we ...
  24. [24]
    [PDF] Introduction to modular forms - Henrik Bachmann
    Nov 20, 2018 · The goal of this course is to prove a dimension formula for modular forms and show that every modular form is actually a polynomial in just E4 ...
  25. [25]
    [PDF] 11.1 Modular forms for congruence subgroups Definition 11.1. For ...
    Mar 18, 2024 · For each positive integer N the principal congruence subgroup of level N is. Γ(N) := {γ ∈ SL2(Z) : γ ≡ 1 mod N}. A subgroup Γ ≤ Γ(1) that ...
  26. [26]
    [PDF] modular forms and the four squares Theorem
    This paper introduces modular forms and aims to obtain a formula for the number of ways to write a positive integer as a sum of four squares.
  27. [27]
    [PDF] 24 Modular forms and L-functions - MIT Mathematics
    Dec 7, 2023 · j(Nτ) is a weak modular form of weight 0 for Γ0(N). For an example of a weak modular form of positive weight, recall the Eisenstein series.
  28. [28]
    Eisenstein Series | SpringerLink
    Cite this chapter. Diamond, F., Shurman, J. (2005). Eisenstein Series. In: A First Course in Modular Forms. Graduate Texts in Mathematics, vol 228. Springer ...
  29. [29]
  30. [30]
    Über Modulfunktionen und die Dirichletschen Reihen mit Eulerscher ...
    Hecke, Die Primzahlen in der Theorie der elliptischen Modulfunktionen, Danske Vidensk. Selsk. Matematisk-fysiske Meddelelser13, 10 (1935), und mein Vortrag ...
  31. [31]
    Über Modulfunktionen und die Dirichletschen Reihen mit Eulerscher ...
    Über Modulfunktionen und die Dirichletschen Reihen mit Eulerscher Produktentwicklung. I. E. Hecke · Mathematische Annalen (1937). Volume: 114, page 1-28; ISSN: ...Missing: Faktoren | Show results with:Faktoren
  32. [32]
    Hecke Operators on ... (m). - EUDML
    LEHNER, J., and ATKIN, AOL. "Hecke Operators on ... (m).." Mathematische Annalen 185 (1970): 134-160. <http://eudml.org/doc/161948>.
  33. [33]
  34. [34]
    [PDF] The ring of modular forms - UvA-DARE (Digital Academic Repository)
    ... [E4, E6] is freely generated over C by the Eisenstein series E4 and E6 of weights 4 and 6. In the 1960s Igusa [11] determined the structure of R2(C): R2(C) ...
  35. [35]
    [PDF] Modular Forms: A Computational Approach William A. Stein (with an ...
    This is a graduate-level textbook about algorithms for computing with mod- ular forms. It is nontraditional in that the primary focus is not on underly-.
  36. [36]
    [PDF] Modular Forms
    (ii) First note that constant functions are modular forms of weight zero. ... is a modular function of weight zero with no poles in H. Choose k ∈ Z such ...
  37. [37]
    AMS :: Journal of the American Mathematical Society
    On the modularity of elliptic curves over Q : Wild 3 -adic exercises. HTML articles powered by AMS MathViewer. by Christophe Breuil, Brian Conrad, Fred ...
  38. [38]
    [PDF] on the modularity of elliptic curves over q: wild 3-adic exercises.
    Introduction. In this paper, building on work of Wiles [Wi] and of Wiles and one of us (R.T.) [TW], we will prove the following two theorems (see §2.2).
  39. [39]
    [PDF] Hecke Operators - UC Davis Math
    If f ∈ Mk, then Tnf ∈ Mk for all n. If f is a cusp form (i.e., the first term of the. Fourier expansion of f is 0), then so is Tnf.
  40. [40]
    [PDF] 24 Modular forms and L-functions - MIT Mathematics
    May 17, 2021 · c d ∈ Γ. Example 24.4. The j-function j(τ) is a weak modular form of weight 0 for SL2(Z), and j(Nτ) is a weak modular form of weight 0 for Γ0(N) ...
  41. [41]
    [PDF] Gauss and the Arithmetic-Geometric Mean
    His work on modular functions and modular forms, and. His work on the arithmetic-geometric mean! David A. Cox (Amherst College). Gauss and the Arithmetic ...
  42. [42]
    [PDF] Riemann and Complex algebraic geometry
    Sep 20, 2024 · In 1857, Riemann published Theorie der Abel'schen Functionen (Theory of abelian functions), an amazing work which poses new themes of research ...
  43. [43]
    Zur Theorie der elliptischen Modulfunktionen
    Wegen der Einzelheiten dieser Theorie sei auf die ausführliche Darstellung von Klein-Fricke, Vorlesungen über die Theorie der elliptischen Modulfunktionen, ...
  44. [44]
    Vorlesungen über die Theorie der elliptischen Modulfunctionen
    Feb 13, 2008 · Vorlesungen über die Theorie der elliptischen Modulfunctionen. by: Klein, Felix, 1849-1925; Fricke, Robert, 1861-1930, ed ... PDF WITH TEXT ...
  45. [45]
    [PDF] Maass form L-functions
    We first recall the definition of Maass form, and then describe the connection with L-functions. A good reference on Maass forms is Iwaniec's book [Iw]. 2.1. ...
  46. [46]
    [PDF] Introduction to Hilbert modular forms
    Aug 1, 2006 · A classical Hilbert modular form of level Γ0(c, n) and weight k is a holomorphic function f : HF → C such that fkk γ = f, for all γ ∈ Γ0(c, n).
  47. [47]
  48. [48]
    [PDF] A short course on Siegel modular forms
    ”Siegel modular forms”, as they are called today, were first introduced by Siegel in a paper of 1935 and nowadays often are given as a first example of ...
  49. [49]
    [PDF] Siegel modular forms: Classical approach and representation theory
    Finally, we introduce the two L-functions associated with Hecke eigenforms – the spin L-function and the standard L-function. 3.1. The Hecke algebra. Let Gn ...<|control11|><|separator|>
  50. [50]
  51. [51]
    [PDF] Notes on modular forms of half-integral weight.
    Oct 2, 2013 · Shimura and Shintani established an extraordinary connection between forms of weight κ/2 and forms of weight κ−1—the “Shimura lift” and the “ ...Missing: developments | Show results with:developments
  52. [52]
  53. [53]
    NEWFORMS OF HALF-INTEGRAL WEIGHT - Project Euclid
    Shimura, On modular forms of half-integral weight, Annals of Math., 97 (1973),. 440-481. [ 9 ] G. Shimura, The critical values of certain zeta functions ...Missing: summary | Show results with:summary
  54. [54]
    [PDF] arXiv:2002.09388v2 [math.RT] 1 Jul 2021
    Jul 1, 2021 · The history of vector-valued modular forms could be traced back ... Vector-valued modular forms (VVMF) are defined as follows. Consider ...
  55. [55]
    [PDF] arXiv:1503.05519v1 [math.NT] 18 Mar 2015
    Mar 18, 2015 · We survey the theory of vector-valued modular forms and their connec- tions with modular differential equations and Fuchsian equations over the ...
  56. [56]
    [PDF] the weil representation and hecke operators for vector valued ...
    The purpose of the present paper is to develop the foundations of a Hecke theory for such vector valued modular forms. The results can be used to associate an L ...