Fact-checked by Grok 2 weeks ago

Selberg trace formula

The Selberg trace formula is a fundamental identity in mathematics that relates the geometry of a locally symmetric space—such as sums over conjugacy classes or closed geodesics—to its spectral properties, including eigenvalues of the Laplacian or irreducible unitary representations of the underlying group. Developed by Norwegian mathematician in the 1950s, it provides a non-commutative analogue of the , equating a geometric expansion (e.g., \sum_{\gamma \in \Gamma} a(\gamma) f(\gamma), where \Gamma is a discrete subgroup and f is a test function) with a spectral expansion (e.g., \sum_{\pi} m(\pi) \operatorname{tr} \pi(f), summing over irreducible representations \pi with multiplicities m(\pi)). Originally formulated by Selberg for the group \mathrm{SL}(2, \mathbb{R}) and compact quotients corresponding to Riemann surfaces of constant negative curvature, the formula expresses the of an on L^2(\Gamma \backslash H) (where H is the plane) via a K(x, y) = \sum_{\gamma \in \Gamma} f(x^{-1} \gamma y), leading to an identity between orbital integrals on the geometric side and traces of representations on the side. This compact case assumes the quotient \Gamma \backslash H is compact, ensuring the decomposes discretely into irreducible components and that the operators are . Selberg's work, published in , built on earlier for discontinuous groups and was motivated by problems in the of the Laplace-Beltrami operator on hyperbolic surfaces. The formula has been generalized extensively, notably by Robert Langlands and James Arthur, to non-compact quotients of reductive groups over number fields using the adelic framework, incorporating cuspidal automorphic forms, truncation operators, and endoscopic transfers. In its Arthur-Selberg form, for a reductive group G over a number field F, the geometric side involves sums over semisimple conjugacy classes in Levi subgroups (e.g., J_o(f) = \sum_{M \in \mathcal{L}} |W_{M_0}| |W_{G_0}|^{-1} \sum_{\gamma \in (M(F) \cap o)_{M,S}} a_M(S, \gamma) J_M(\gamma, f)), while the spectral side integrates over induced representations and intertwining operators (e.g., J_{\chi}(f) = \sum_{\pi} m_{\mathrm{cusp}}(\pi) \int_{i \mathfrak{a}^*_P} \operatorname{tr}(M_P(\pi_\lambda) I_P(\pi_\lambda, f)) \, d\lambda). These extensions, developed from the 1970s onward, rely on Harish-Chandra's Plancherel formula and the Jacquet-Langlands correspondence. The significance of the Selberg trace formula lies in its bridging of arithmetic geometry and , enabling proofs of results like Weyl's law for the eigenvalue distribution on hyperbolic manifolds (e.g., the number of eigenvalues up to T is asymptotically \frac{\mathrm{Area}(\Gamma \backslash H)}{4\pi} T^2) and estimates on the lengths of closed geodesics. It plays a central role in the , facilitating the study of automorphic L-functions, Hecke operators, and ε-factors, with applications to reciprocity laws, Shimura varieties, and even approaches to the via spectral gaps in Maass cusp forms. In spectral geometry, it provides explicit links between the spectrum of the Laplacian and the topology of the space, influencing research in and on symmetric spaces.

Historical Development

Early Contributions

The origins of the Selberg trace formula trace back to the mid-1950s, when , a then serving as a permanent faculty member at the Institute for Advanced Study in Princeton, developed key ideas in on surfaces. Selberg first announced the formula in 1954, publishing it fully in 1956. His work was deeply influenced by the traditions established by and Charles Jean de la Vallée Poussin, whose proofs of the in 1896 utilized to link zeta functions with prime distributions. Additionally, interactions with at the Institute shaped Selberg's approach to integrating geometric and arithmetic perspectives in discontinuous group actions on symmetric spaces. Selberg's ideas were also influenced by Harish-Chandra's contemporaneous work on representations of Lie groups. In 1956, Selberg published his seminal paper " and Discontinuous Groups in Weakly Symmetric Riemannian Spaces with Applications to ," where he introduced the trace formula specifically for compact Riemann surfaces of constant negative curvature. This formula provided an equating a weighted sum over the lengths of closed geodesics on the surface to a sum involving eigenvalues of the hyperbolic Laplacian, thereby bridging and in a novel way. Selberg's formulation applied to quotients \Gamma \backslash \mathbb{H} by cocompact discrete subgroups \Gamma of \mathrm{[PSL](/page/PSL)}(2, \mathbb{R}), establishing a foundational tool for studying the spectrum of such surfaces. A central motivation for Selberg's trace formula was its analogy to the explicit formulas in theory, such as those derived from the , where sums over primes correspond to non-trivial zeros. In the geometric setting, primitive closed s on the surface play the analogous role to s, with their lengths dictating the distribution of data much like prime distributions influence zeta zeros. This parallel allowed Selberg to extend analytic techniques from to geodesic flows, enabling estimates on eigenvalue spacings and geodesic counts. Closely tied to the trace formula, Selberg introduced the Selberg zeta function in the same 1956 work, defined as an infinite product over primitive closed geodesics \gamma with lengths \ell(\gamma): Z(s) = \prod_{\gamma} \prod_{k=0}^{\infty} \left(1 - e^{-(s+k)\ell(\gamma)}\right), which serves as a determinant-like analogue to the and encodes the lengths of geodesics in its zeros and poles. This zeta function's relates directly to the trace formula, facilitating asymptotic results akin to the , such as the prime geodesic theorem counting primitive geodesics up to length x. Concurrently, the Eichler-Selberg trace formula emerged as a related development for spaces of s, pioneered by Martin Eichler in the early 1950s, inspired by Selberg's trace formula. This formula computes the trace of Hecke operators T_m acting on the space S_k(\mathrm{SL}(2, \mathbb{Z})) of cusp forms of weight k, expressing it as a sum involving class numbers or divisor functions weighted by the operator index m. Proofs of the formula leverage the Riemann-Roch theorem to determine dimensions of spaces and incorporate contributions, providing an arithmetic counterpart to Selberg's geometric trace formula tailored to the .

Later Extensions

In the late 1950s and early , extensions of the Selberg trace formula to non-compact quotients emphasized the role of in capturing the continuous spectrum arising from the non-compact fundamental domain. Selberg himself advanced this framework in his 1962 address, where he outlined the on discontinuous groups, demonstrating how provide the necessary intertwining operators to decompose the space of square-integrable functions and derive the trace identity for the continuous part. This development resolved key challenges in applying the formula to Fuchsian groups with cusps, enabling spectral decompositions that include both discrete eigenvalues and the scattering contributions from . During the 1970s, integrated the Selberg trace formula into the theory of automorphic forms on groups, transforming it into a tool for studying representations of reductive groups over global fields. In unpublished manuscripts circulated during this period and later formalized in surveys, Langlands completed the of for general groups, showing how they facilitate the trace formula's adaptation to settings and link spectral data to orbital integrals. This extension, building on adelic automorphic forms, established the trace formula as a cornerstone for reciprocity laws in the emerging , allowing comparisons between automorphic representations and Galois representations. James Arthur's contributions in the 1980s and 1990s generalized the trace formula to arbitrary connected reductive groups over number fields, addressing the geometric and spectral complexities beyond rank-one cases like SL(2). In a series of foundational papers, Arthur introduced a truncation operator to ensure integrability of the geometric terms and formulated an invariant version of the trace formula, expressed as an equality between distributions on the group and traces on the automorphic representation space. His work culminated in explicit decompositions of the geometric side into contributions from Levi subgroups, providing a robust framework for applications in representation theory. These advancements intertwined the trace formula with the through and stable variants, which refine the formula to isolate stable orbital integrals invariant under inner automorphisms. Langlands proposed the stable trace formula in 1983 as a means to relate distributions across endoscopic groups, facilitating functoriality conjectures by equating stable traces for a group and its endoscopies. Arthur's subsequent stabilization in the realized this vision, decomposing the formula into stable and endoscopic components to prove key instances of transfer principles. In parallel, during the , Nolan Wallach and collaborators deepened the trace formula's implications for unitary representations of real reductive groups, particularly in classifying series via formulas derived from traces. Wallach's linked the formula to cohomological properties of representations, showing how traces on K-finite vectors yield multiplicity formulas for unitary modules in the automorphic spectrum. This work highlighted the formula's utility in verifying unitarity and computing formal degrees for irreducible constituents.

Mathematical Foundations

Hyperbolic Geometry and Quotients

The hyperbolic plane, denoted \mathbb{H}^2, is the unique simply connected, complete Riemannian manifold of dimension 2 with constant sectional curvature -1. It can be realized as the upper half-plane model \mathbb{H} = \{ z = x + iy \mid y > 0 \} equipped with the Riemannian metric ds^2 = \frac{dx^2 + dy^2}{y^2}, which induces the invariant volume form d\mu = \frac{dx \, dy}{y^2}. This metric ensures that geodesics are semicircles orthogonal to the real axis or vertical lines, and the group of orientation-preserving isometries is \mathrm{PSL}(2, \mathbb{R}), acting via Möbius transformations z \mapsto \frac{az + b}{cz + d} for \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \mathrm{SL}(2, \mathbb{R}). Higher-dimensional hyperbolic spaces \mathbb{H}^n generalize this construction, embedding as the hyperboloid sheet in Minkowski space with Lorentzian metric of signature (n,1), preserved by the connected component \mathrm{SO}^+(n,1) of the orthogonal group, yielding constant curvature -1. Discrete subgroups \Gamma of \mathrm{PSL}(2, \mathbb{R}) (or more generally of semisimple Lie groups G such as \mathrm{SO}^+(n,1)) are Fuchsian groups that act properly discontinuously on \mathbb{H}^n, meaning every point has a neighborhood intersecting only finitely many translates under \Gamma. Such groups are classified by their action: cocompact subgroups produce finite-volume quotients with compact fundamental domain, ensuring the quotient is a closed manifold without cusps, while non-cocompact subgroups yield infinite-volume quotients with cuspidal ends, as in the case of \Gamma = \mathbb{Z} generated by a parabolic isometry. In the cocompact case, \Gamma consists of hyperbolic elements with no fixed points on the boundary, leading to torsion-free actions for surface quotients. The quotient space X = \Gamma \backslash \mathbb{H}^n inherits the hyperbolic metric, forming a Riemannian orbifold or manifold of constant negative curvature -1. For n=2 and cocompact torsion-free \Gamma, X is a compact Riemann surface of genus g \geq 2, with area $4\pi (g-1) determined by the Gauss-Bonnet theorem. These surfaces provide the geometric setting for spectral analysis, where the universal cover \mathbb{H}^2 projects down via the deck transformation group \Gamma. Closed geodesics on X correspond bijectively to conjugacy classes of elements in \Gamma, where a \gamma \in \Gamma fixes two points on the at and translates along its by \ell(\gamma) = 2 \cosh^{-1} \left( \frac{|\mathrm{tr}(\gamma)|}{2} \right) > 0. closed geodesics arise from conjugacy classes (inequivalent under powers), with the centralizer Z_\gamma being cyclic generated by \gamma. The set of such lengths encodes the of X. Invariant measures on G / \Gamma are induced by the on G, restricted to the quotient, ensuring G-invariance under left . Orbital integrals quantify contributions from \Gamma-orbits, defined as O_\gamma(f) = \int_{G_\gamma \backslash G} f(g^{-1} \gamma g) \, dg for test functions f on G, where G_\gamma is the centralizer of \gamma; these integrate over conjugacy classes to capture geometric data in trace formulas. For hyperbolic surfaces, the volume form d\mu on \mathbb{H}^2 descends to a \Gamma- measure on X, facilitating over fundamental domains.

Spectral Theory and Representations

The spectral theory underlying the Selberg trace formula relies on the analysis of the Laplacian on hyperbolic surfaces, particularly the Maass Laplacian, which governs the of automorphic forms. On a hyperbolic surface X = \Gamma \backslash \mathbb{H}^2, where \Gamma is a and \mathbb{H}^2 is the plane, the Maass Laplacian \Delta is the Laplace-Beltrami acting on non-holomorphic cusp forms. Its eigenvalues form the , parameterized as \lambda_n = \frac{1}{4} + r_n^2 for n = 0, 1, 2, \dots, where r_n \in \mathbb{R} or r_n = i s_n with s_n > 0 for eigenvalues below \frac{1}{4}, and the corresponding eigenfunctions are Maass forms that transform under the action of \Gamma. These eigenvalues encode the contributions to the , distinguishing the series from potential continuous in non-compact cases. Central to the framework are the unitary representations of the G = \mathrm{PSL}(2, \mathbb{R}) on the L^2(\Gamma \backslash G), which decomposes into a of irreducible unitary representations comprising the discrete series and, in non-compact settings, the continuous series. The discrete series consists of finite-multiplicity representations induced by automorphic forms, including the trivial and those corresponding to Maass cusp forms, while the continuous series arises from the non-compact cusps and is parameterized by the unitary principal series. This decomposition reflects the automorphic , where the right-regular representation of G on L^2(\Gamma \backslash G) integrates the spectral data of the Laplacian with the group's structure. Traces of operators in this representation-theoretic context, such as the and resolvent traces, provide the analytic tools for the spectral side of the trace formula. The K_t(x, y) for the operator e^{t \Delta} on L^2(X) has a trace \operatorname{Tr}(e^{t \Delta}) = \sum_n e^{-t \lambda_n} over the discrete spectrum, smoothed by the continuous part in general cases, and serves as a fundamental trace operator whose expansion relates to representation characters. Similarly, the resolvent trace \operatorname{Tr}((\Delta + s(1-s))^{-1}) for \operatorname{Re}(s) > 1 captures spectral multiplicities and is regularized via zeta-function techniques. These traces are computed within the unitary representations, leveraging the group's invariance to derive explicit expressions. The Plancherel formula decomposes the L^2 space into irreducible components, quantifying the contributions of each to the inner product on L^2(\Gamma \backslash G). For G = \mathrm{[PSL](/page/PSL)}(2, \mathbb{R}), Harish-Chandra's describes the decomposition of L^2(G) via a sum over the series representations and an over the continuous (principal) series with respect to the Plancherel measure supported on the unitary \hat{G}. In the automorphic setting, L^2(\Gamma \backslash G) decomposes into a direct of irreducible automorphic representations, including parts from cusp forms and continuous parts from . This framework underpins the trace computations by providing the measure for integrating over representations. Harish-Chandra's \mathcal{C}(G) plays a crucial role as the space of smooth functions with rapid decay under K-bi-invariant norms, ensuring convergence of traces and coefficients in the trace formula. Defined as the intersection of spaces with derivatives decaying faster than any along the group, it includes K-finite functions and supports the Paley-Wiener for holomorphic extensions, allowing test functions like the Harish-Chandra transform H \phi(\lambda) = \int_G \phi(g) \overline{\Psi_\lambda(g)} \, dg to be used in spectral expansions. This space guarantees the tempered nature of distributions in the formula's derivations.

Core Formulations

Compact Hyperbolic Surfaces

The Selberg trace formula achieves its simplest form when applied to compact hyperbolic surfaces, which are Riemann surfaces of genus g \geq 2 equipped with a metric of constant curvature -1. Such a surface X can be expressed as X = \Gamma \backslash \mathbb{H}^2, where \Gamma is a torsion-free cocompact Fuchsian subgroup of \mathrm{PSL}(2, \mathbb{R}) and \mathbb{H}^2 is the hyperbolic plane. The spectrum of the hyperbolic Laplacian -\Delta on L^2(X) is purely discrete, consisting of eigenvalues $0 = \lambda_0 < \lambda_1 \leq \lambda_2 \leq \cdots with \lambda_j = \frac{1}{4} + r_j^2 and r_j \geq 0 real (including multiplicity), and finite-dimensional eigenspaces. Unlike non-compact cases, compactness ensures no continuous spectrum, so the formula encodes solely the discrete spectral data alongside geometric invariants. For a suitable even test function h: \mathbb{R} \to \mathbb{C} analytic in a horizontal strip around the real axis and decaying sufficiently fast at infinity (e.g., |h(\rho)| \ll (1 + |\rho|)^{-2-\delta} for some \delta > 0), the Selberg trace formula states: \sum_{j=0}^\infty h(r_j) = \frac{\mathrm{Area}(X)}{4\pi} \int_{-\infty}^\infty h(r) \frac{\tanh(\pi r)}{r} \, dr + \sum_{\{\gamma\}} \sum_{n=1}^\infty \frac{\ell_\gamma \, g(n \ell_\gamma)}{2 \sinh(n \ell_\gamma / 2)}, where the outer sum runs over primitive conjugacy classes \{\gamma\} of hyperbolic elements in \Gamma, \ell_\gamma > 0 is the length of the corresponding primitive closed geodesic, and g(t) = \int_{-\infty}^\infty h(r) e^{i r t} \, dr is the Fourier transform of h. This identity, originally derived by Selberg, equates a spectral side (sum over eigenvalues) with an archimedean term (arising from the universal cover) plus a geometric side (sum over closed geodesics and their multiples). The formula establishes a profound duality between the distribution of Laplacian eigenvalues on X and the lengths of its closed geodesics, providing a non-commutative analogue of the prime number theorem via geodesic analogues of primes. The absence of a continuous spectrum term simplifies analysis, allowing direct applications to spectral asymptotics and prime geodesic counting; for instance, choosing h(r) = e^{-t r^2} (with t > 0) yields the heat trace expansion, linking short-time heat kernel behavior to small eigenvalue clustering and short geodesics. By Gauss-Bonnet, \mathrm{Area}(X) = 4\pi (g-1), scaling the archimedean contribution with surface complexity. Closely related is the Selberg zeta function, defined for \mathrm{Re}(s) > 1 by Z(s) = \prod_{\{\gamma\}} \prod_{k=0}^\infty \left(1 - e^{-(s+k) \ell_\gamma}\right), an infinite over primitive closed geodesics that converges absolutely and encodes the length multiplicatively. The trace formula facilitates the meromorphic of Z(s) to the entire , revealing simple zeros precisely at the points s = \frac{1}{2} + i r_j for j \geq 1 (corresponding to non-trivial eigenvalues) and additional trivial zeros at non-positive integers. Moreover, the satisfies -\frac{Z'(s)}{Z(s)} = \sum_{\{\gamma\}} \sum_{n=1}^\infty \frac{\ell_\gamma \, e^{-n s \ell_\gamma}}{1 - e^{-n s \ell_\gamma}} for \mathrm{Re}(s) > 1, directly mirroring the geodesic side of the trace formula and enabling explicit connections between spectral zeros and geodesic contributions. A concrete illustration appears for the Bolza surface, the unique genus-2 hyperbolic surface with maximal symmetry (full of order 48). Its smallest positive eigenvalue is \lambda_1 \approx 3.838 with multiplicity 3 (so r_1 \approx 1.894), followed by \lambda_2 \approx 5.351 with multiplicity 4. The shortest primitive closed geodesic has length \ell \approx 3.057 (explicitly $2 \arccosh(1 + \sqrt{2})) with multiplicity 12, reflecting the surface's ; these values satisfy the trace formula for appropriate test functions, underscoring the eigenvalue-geodesic interplay.

Cocompact Discrete Subgroups

In the context of cocompact subgroups \Gamma of a unimodular G, the Selberg trace formula provides a profound link between the of the quotient space \Gamma \backslash G and its geometric structure, equating the of a to a sum over group elements. This formulation, developed by Selberg and extended by Harish-Chandra, applies to the regular representation on L^2(\Gamma \backslash G), which decomposes discretely due to compactness. For a test function \phi \in C_c^\infty(G), the trace of the associated operator R(\phi), defined by (R(\phi) f)(x) = \int_G \phi(y) f(xy^{-1}) \, dy, is given by \operatorname{Tr} R(\phi) = \sum_{\gamma \in \Gamma} \frac{\vol(\Gamma \backslash G)}{|\det(1 - \Ad(\gamma))|} \int_G \phi(g^{-1} \gamma g) \, dg, where the sum runs over all non-identity elements \gamma, grouped by conjugacy classes, \vol(\Gamma \backslash G) is the invariant volume of the quotient, and \Ad denotes the adjoint representation of G. This equality holds because \Gamma \backslash G is compact, ensuring the operator is trace-class. The geometric side of the formula involves orbital integrals \int_G \phi(g^{-1} \gamma g) \, dg, which measure the distribution of \phi along the of \gamma, weighted by the factor \vol(\Gamma \backslash G)/|\det(1 - \Ad(\gamma))|. This accounts for the dimension of the fixed-point set or the centralizer of \gamma, reflecting the volume of the in the quotient ; for semisimple elements, it relates to the length of closed geodesics on the associated symmetric space G/K, where K is a maximal compact . The sum thus encodes global geometric invariants, such as lengths and multiplicities of conjugacy classes in \Gamma. On the spectral side, the trace decomposes as \operatorname{Tr} R(\phi) = \sum_{\pi} m_\pi \operatorname{tr} \pi(\phi), where the sum is over irreducible unitary representations \pi in the discrete series of G, m_\pi denotes the multiplicity of \pi in the decomposition of the , and \operatorname{tr} \pi(\phi) = \int_G \phi(g) \chi_\pi(g) \, dg is the character integral, with \chi_\pi the of \pi. For cocompact \Gamma, the discrete series exhausts the , and multiplicities m_\pi are finite, arising from the for compact quotients. This side captures the eigenvalues and eigenfunctions of differential operators, such as the operator, on \Gamma \backslash G. When G = \mathrm{PSL}(2, \mathbb{R}), the formula specializes to the classical Selberg trace formula for compact surfaces \Gamma \backslash \mathbb{H}, where the discrete series representations correspond to Maass forms and their Laplacian eigenvalues \lambda_j = 1/4 + t_j^2, and the geometric terms reduce to sums over closed geodesics weighted by their lengths. In this case, the orbital integrals simplify to explicit hyperbolic and elliptic contributions, bridging the abstract group-theoretic setting to concrete geometry. The trace formula also underpins Weyl's law for the eigenvalue asymptotics on \Gamma \backslash G, stating that the counting function N(T) for eigenvalues up to T satisfies N(T) \sim \frac{\vol(\Gamma \backslash G)}{(2\pi)^n} \omega_n T^n as T \to \infty, where n is the rank of the symmetric space and \omega_n a constant depending on G; for G = \mathrm{PSL}(2, \mathbb{R}), this yields N(T) \sim \frac{\area(\Gamma \backslash \mathbb{H})}{4\pi} T. This asymptotic reflects the volume growth and provides a spectral analogue of the prime number theorem via the geometric side.

Proofs and Techniques

Proof for Compact Case

The proof of the Selberg trace formula for compact surfaces relies on the method, which relates the spectral properties of the Laplacian to geometric invariants via unfolding and summation over the group action. Consider a compact X = \Gamma \backslash \mathbb{H}, where \mathbb{H} is the hyperbolic plane and \Gamma is a torsion-free of finite covolume acting freely and properly discontinuously. The K_t(x, y) on X is constructed as the \Gamma-invariant lift: K_t(x, y) = \sum_{\gamma \in \Gamma} k_t(\gamma x, y), where k_t(z, w) denotes the heat kernel on \mathbb{H}, explicitly given by k_t(z, w) = (4\pi t)^{-1} e^{-d(z,w)^2/(4t) + 1/4} \theta'(it; \sigma(z,\overline{w})) with \theta the Jacobi theta function and \sigma the Poincaré distance function squared. This summation converges absolutely for t > 0 due to the of k_t and the discreteness of \Gamma. The spectral expansion of the heat operator e^{-t \Delta} on L^2(X) yields \operatorname{Tr}(e^{-t \Delta}) = \sum_{n=0}^\infty e^{-t \lambda_n}, where \{\lambda_n\}_{n=0}^\infty are the eigenvalues of -\Delta on X with \lambda_0 = 0 (constant eigenfunction) and \lambda_n \geq 1/4 for n \geq 1. By the unitarity of the eigenbasis and the integral kernel representation, this trace equals the global trace \int_X K_t(x, x) \, dx, providing an analytic continuation of the spectral sum for small t > 0. As t \to 0^+, the leading term is \operatorname{vol}(X)/(4\pi t), reflecting the short-time asymptotic of the heat kernel, with higher-order terms involving curvature and topology. To derive the full trace formula, apply Poisson summation over the orbits of \Gamma, unfolding the integral \int_X K_t(x, x) \, dx = \int_{\Gamma \backslash \mathbb{H}} \sum_{\gamma \in \Gamma} k_t(\gamma x, x) \, dx. By the invariance of the hyperbolic metric, this equals \sum_{\gamma \in \Gamma} \int_{\mathbb{H}} k_t(\gamma x, x) \, dx, but grouping by conjugacy classes refines it to a over closed geodesics. Specifically, the contribution from nontrivial elements \gamma corresponds to sums over closed geodesics \ell_\gamma (lengths of hyperbolic conjugacy classes), yielding terms of the form \sum_{\{\gamma\}} \sum_{m=1}^\infty \frac{\ell_\gamma}{2 \sinh(m \ell_\gamma / 2)} e^{-t m \ell_\gamma / 2} after for the multiplicity and the kernel's radial . The \gamma = e contributes the continuous spectrum analog on \mathbb{H}, approximated by \operatorname{vol}(X) \int_0^\infty e^{-t(1/4 + r^2)} \frac{r \tanh(\pi r)}{\sinh(2\pi r)} \, dr in the spectral parameter. For the general form, regularize using even test functions h(r) on the spectral parameter r_n = \sqrt{\lambda_n - 1/4}, defined via the inverse h(r) = \int_{-\infty}^\infty e^{i r s} g(s) \, ds where g(s) is the even part of a function \phi(s/2) e^{-s^2/4} or similar, ensuring h is analytic in |\operatorname{Im} r| \leq 1/2 + \delta for some \delta > 0 and decays as |h(r)| \ll (1 + |r|)^{-2-\epsilon}. The spectral side becomes \sum_n h(r_n), while the geometric side is \frac{\operatorname{vol}(X)}{4\pi} \int_{-\infty}^\infty \frac{h(r) \tanh(\pi r)}{r} \, dr + \sum_{\gamma \in \Gamma^*} \sum_{m=1}^\infty \frac{\ell_\gamma}{2 \sinh(m \ell_\gamma / 2)} g(m \ell_\gamma), where \Gamma^* indexes classes. This regularization via the pair connects the trace (for \phi(t) = e^{-t/4}) to the general formula, avoiding divergences at low eigenvalues. In the compact case, holds because X has finite and no continuous , making the spectral sum \sum |h(r_n)| < \infty by the decay of h and Weyl's law (\#\{n : \lambda_n \leq T\} \sim \operatorname{vol}(X) T / (4\pi)), while the geometric series over finitely many short geodesics (per length bound) and exponential decay of g ensure the double sum converges uniformly. This yields the explicit Selberg trace formula equating spectral and geometric sides without further truncation.

General Proof Strategy

The general proof strategy for the Selberg trace formula in the cocompact case leverages the unitary representation theory of the Lie group G = \mathrm{PSL}(2, \mathbb{R}) on the space L^2(\Gamma \backslash G), where \Gamma is a cocompact discrete subgroup, to equate the spectral and geometric traces of a trace-class operator derived from a suitable test function. This approach, originally developed by Selberg, decomposes the right-regular representation of G on L^2(\Gamma \backslash G) into a direct sum of irreducible unitary representations with finite multiplicities and computes the trace via the Plancherel theorem on one side and a fixed-point formula on the other. Central to the strategy is the definition of the operator R(\phi) on L^2(\Gamma \backslash G) for a test function \phi \in C_c^\infty(G), given by convolution with the right-regular representation: R(\phi) f(x) = \int_G \phi(y) f(x y) \, dy. This operator is trace-class because \Gamma is cocompact, ensuring the integral of its kernel over the diagonal converges absolutely. The kernel of R(\phi) is K_\phi(x, y) = \sum_{\gamma \in \Gamma} \phi(x^{-1} \gamma y), which unfolds the sum over the discrete group action to the full group G when computing traces. The spectral trace of R(\phi) is computed using the Plancherel formula for the decomposition L^2(\Gamma \backslash G) = \bigoplus_\pi m_\pi \cdot \pi, where the sum runs over irreducible unitary representations \pi of G (primarily the discrete series in this case) and m_\pi denotes the finite multiplicity of \pi. Thus, the trace equals \sum_\pi m_\pi \operatorname{Tr}(\pi(\phi)), where \operatorname{Tr}(\pi(\phi)) = \int_G \chi_\pi(g) \phi(g) \, dg and \chi_\pi is the character of \pi. Harish-Chandra's character formulas provide explicit expressions for \chi_\pi on conjugacy classes, enabling the evaluation of these integrals for discrete series representations by relating them to Weyl group sums and root data of G. On the geometric side, the trace is \int_{\Gamma \backslash G} K_\phi(x, x) \, dx = \sum_{\gamma \in \Gamma} \int_{\Gamma \backslash G} \phi(x^{-1} \gamma x) \, dx. By the fixed-point formula and change of variables (unfolding over centralizers), this simplifies to \sum_{[\gamma]} \vol(\Gamma_\gamma \backslash G_\gamma) J_\gamma(\phi), where the sum is over conjugacy classes [\gamma] in \Gamma, \Gamma_\gamma is the centralizer of \gamma in \Gamma, and the orbital integral is J_\gamma(\phi) = \int_{G_\gamma \backslash G} \phi(g^{-1} \gamma g) \, dg. The equality of spectral and geometric traces then yields the Selberg trace formula, with the representation-theoretic tools ensuring convergence and identification of terms.

Generalizations

Non-Cocompact Cases

In non-cocompact cases, the Selberg trace formula is adapted to handle quotients with infinite volume, such as those arising from Fuchsian groups with parabolic elements, by incorporating a continuous spectrum contribution via Eisenstein series associated to parabolic subgroups. For a discrete subgroup \Gamma of \mathrm{SL}(2, \mathbb{R}) with parabolic cusps, the Eisenstein series E(z, s) is defined for \mathrm{Re}(s) > 1 as E(z, s) = \sum_{\gamma \in \Gamma_\infty \backslash \Gamma} \mathrm{Im}(\gamma z)^s, where \Gamma_\infty is the stabilizer of the cusp at infinity; this series admits an to the except for a simple pole at s=1, capturing the continuous part of the spectrum of the hyperbolic Laplacian on \Gamma \backslash \mathbb{H}. The spectral side of the trace formula is modified to include an integral over the continuous spectrum, given by \frac{1}{4\pi} \int_{-\infty}^\infty |E(\cdot, 1/2 + i r)|^2 h(r) \, dr, alongside the discrete sum \sum_j h(r_j) from Maass cusp forms, where h(r) is the Fourier transform of a test function and r_j parametrizes the discrete eigenvalues \lambda_j = 1/4 + r_j^2. This term arises from the spectral decomposition of the automorphic kernel, ensuring the formula equates to the geometric side involving orbital integrals over conjugacy classes in \Gamma. For the specific case of the modular surface \mathrm{SL}(2, \mathbb{Z}) \backslash \mathbb{H}, the Eisenstein series Fourier expansion involves the Riemann zeta function, with the continuous spectrum integral relating directly to the scattering matrix \sigma(s) = \frac{\zeta(2s-1)}{\zeta(2s)} (up to Gamma factors), whose logarithmic derivative \frac{\sigma'(1/2 + i r)}{\sigma(1/2 + i r)} appears in the integrated form, contributing residues at poles like s=1. The scattering matrix \sigma(s) plays a crucial role in accounting for residue contributions from the poles of the , which encode volume terms and interactions at the cusps; for instance, the residue at s=1 yields a proportional to the or cusp width. In the modular surface example, this links the trace formula to the non-trivial zeros of the , as the phases and densities in the continuous spectrum term reflect the distribution of \zeta's zeros via the embedded in \sigma(s), providing a interpretation for distribution akin to the explicit formula. A key challenge in these non-cocompact settings is the failure of of the due to the continuous and non-trace-class operators; this is resolved by introducing a function, such as the of a truncated fundamental domain F(Y) = \{ z \in \Gamma \backslash \mathbb{H} : \mathrm{Im}(z) \leq Y \}, which regularizes the and adds boundary correction terms vanishing as Y \to \infty, or by employing weighted traces with test functions h(r) supported away from zero to ensure integrability.

Arthur-Selberg Trace Formula

The Arthur-Selberg formula represents a profound generalization of the classical to arbitrary connected reductive algebraic groups G over a number field F, where \Gamma = G(F) is the discrete subgroup of G(\mathbb{A}), with \mathbb{A} the ring. This formulation equips the trace formula with an invariant structure under conjugation, enabling its application to a broad class of semisimple groups beyond the rank-one cases like \mathrm{SL}(2). The formula equates a geometric side, which decomposes into weighted orbital integrals incorporating endoscopic transfers from smaller groups, with a spectral side involving virtual traces of tempered automorphic representations parameterized by . On the geometric side, the distribution J(f) for a suitable test function f \in C_c^\infty(G(\mathbb{A})) expands as a sum over Levi subgroups M, featuring terms like J_o(f) = \sum_{M} \frac{|W_M^0|}{|W_G^0|} \sum_{\gamma \in \Gamma(M)} a_M(\gamma) J_M(\gamma, f), where a_M(\gamma) are global coefficients accounting for volumes of centralizers, and J_M(\gamma, f) denotes weighted orbital integrals over G(\mathbb{A})_\gamma \backslash G(\mathbb{A}), adjusted by truncation functions for convergence. Endoscopic transfers enter through stable distributions, summing over endoscopic groups G' with transfer factors \iota(G, G') that map orbital integrals from inner forms or related groups to the primary group G, ensuring the geometric expansion is invariant and decomposes semisimple conjugacy classes appropriately. These transfers, weighted by global root numbers and Tamagawa measures, capture contributions from non-trivial inner forms, such as quaternionic or anisotropic tori. The spectral side mirrors this structure through virtual traces over the space of tempered representations \Pi_{\mathrm{temp}}(G(\mathbb{A})), expressed as J_\chi(f) = \sum_M \frac{|W_M^0|}{|W_G^0|} \sum_{\pi \in \Pi_\chi(M)} m(\pi) \int_{i \mathfrak{a}_M^*} \mathrm{tr}(M_{P}(\pi_\lambda) I_{P}(\pi_\lambda, f)) \, d\lambda, where \chi indexes characters, m(\pi) are multiplicities, and the trace involves intertwining operators M_P and induced representations I_P along the spectral parameter \lambda, with Langlands parameters classifying the irreducible constituents via L-packets. Tempered representations are those whose Langlands parameters lie in the compact real form of the dual group, ensuring unitarity and stability under endoscopy. The invariant trace formula asserts the equality J(f) = \sum_o J_o(f) = \sum_\chi J_\chi(f) after applying coefficients a_G(\gamma) and b_G(\pi), which are products of local factors ensuring independence from choices of maximal compact subgroups K and Haar measures; these coefficients, derived from Tamagawa numbers and volumes, render both sides invariant under inner automorphisms. This invariance facilitates applications to inner forms, where the formula for a quasi-split group G relates via endoscopic lifting to anisotropic inner forms like algebras, allowing transfer of automorphic representations between groups of the same L-group. For instance, it underpins the Jacquet-Langlands correspondence by equating stable traces across such liftings, enabling classification of representations on classical groups through their geometric counterparts.

Applications

Analytic Number Theory

The Selberg trace formula has profound applications in , particularly in establishing connections between spectral data on arithmetic surfaces and arithmetic invariants such as L-functions and class numbers. One key application arises through the Eichler-Shimura theory, where traces of Hecke operators on spaces of cusp forms, computed via variants of the Selberg trace formula known as Eichler-Selberg trace formulas, relate to periods of automorphic forms and enable the determination of Hasse-Weil L-functions for modular curves. Specifically, these traces express sums involving Hurwitz-Kronecker class numbers, which factor into the Euler product structure of the L-functions attached to elliptic modular forms, linking the and functional equations of these L-functions to geometric zeta functions of elliptic curves over number fields. A prominent arithmetic application is the prime geodesic theorem, which provides an asymptotic count for the number of primitive closed s of length at most T on the modular surface, analogous to the for the . The theorem states that \sum_{\ell_\gamma \leq T} 1 \sim \mathrm{li}(T), where \ell_\gamma denotes the length of the primitive \gamma and \mathrm{li}(T) is the logarithmic integral; this follows from the side of the Selberg trace formula, which encodes the distribution of Laplace eigenvalues, balanced against the geometric side summing over geodesic lengths. This equidistribution result extends to more general arithmetic Fuchsian groups, yielding insights into the growth of special values of Dedekind zeta functions. The zeros of the Selberg zeta function, defined as a product over primitive geodesics and encoding the Laplace spectrum, exhibit connections to exceptional zeros in Dirichlet L-functions, particularly Landau-Siegel zeros. Low-lying real zeros of the Selberg zeta function near the critical line correspond to potential Siegel zeros in quadratic Dirichlet L-functions, influencing bounds on Fourier coefficients of Maass cusp forms; specifically, the absence of Siegel zeros implies subconvexity for these coefficients, derived via the trace formula's relation between spectral traces and Kloosterman sums. This analogy highlights deep parallels between the Riemann hypothesis for the Selberg zeta and the non-existence of real zeros close to 1 for primitive L-functions. Trace identities from the Selberg formula also facilitate explicit computations of class numbers and regulators for quadratic fields. For imaginary quadratic fields, Eichler-Selberg formulas express Hecke traces as weighted sums of Hurwitz class numbers, allowing numerical evaluation of class numbers via spectral data on modular forms. In the real quadratic case, the trace formula applied to Hilbert modular surfaces enables unconditional algorithms for computing class groups and regulators up to large discriminants, leveraging approximations of Maass forms and asymptotic evaluations of orbital integrals to resolve the class number formula hR = \frac{w \sqrt{D}}{2} L(1, \chi_D), where h is the class number, R the regulator, w the number of units, D the discriminant, and \chi_D the quadratic character. Post-2010 developments have utilized the variant of the Selberg trace formula—a summation formula over Kloosterman sums and —to obtain subconvexity bounds for L-functions attached to GL(2) automorphic forms, particularly in the spectral and level aspects. For instance, these techniques yield bounds like L(1/2 + it, f) \ll_\epsilon (t \mathrm{cond}(f))^{1/3 - \delta} for holomorphic cusp forms f, improving convexity by exploiting approximate functional equations and delta-symbol approximations within the trace formula; such results have advanced understanding of central L-values and their moments, with applications to equidistribution in the space of modular forms. More recent applications, as of 2024, include using the trace formula to study zeros of and base change relations, providing lower bounds on sums of zeros. Additionally, frameworks treating the via definable distributions have emerged, enabling o-minimal and Diophantine tools.

Geometry and Quantum Chaos

The Selberg trace formula plays a pivotal role in understanding isospectral deformations of surfaces, where distinct geometries share the identical of the Laplacian. In 1986, Peter Buser constructed explicit examples of non-isometric compact Riemann surfaces of 5 and higher that are isospectral, demonstrating that the eigenvalue does not uniquely determine the underlying . These constructions rely on gluing techniques that preserve the invariants while altering the geometric structure, highlighting the formula's utility in probing the limitations of spectral rigidity in low dimensions. Subsequent work extended Buser's methods to produce families of such surfaces, underscoring the non-uniqueness of geometry from spectra alone. In quantum chaos, the Selberg trace formula serves as an exact analog to the semiclassical Gutzwiller trace formula, establishing a direct link between the eigenvalues of the Laplacian and the periodic orbits (closed geodesics) on the surface. The formula expresses the trace of the or as a sum over these orbits, weighted by stability factors, mirroring how Gutzwiller's approximation connects quantum energy levels to classical periodic trajectories in chaotic systems. This correspondence has been instrumental in studying the statistical properties of eigenvalues for hyperbolic surfaces, where the trace formula provides rigorous periodic-orbit sums that validate semiclassical predictions in the chaotic regime. The length spectrum, comprising the lengths of closed geodesics, emerges as a key geometric invariant recoverable from the Selberg trace formula via of the spectral side. Jean-Pierre Otal proved in that for negatively curved surfaces, the marked length spectrum—pairing geodesic lengths with homotopy classes—uniquely determines the hyperbolic metric up to , establishing a form of spectral rigidity. This result implies that if two such surfaces share the same marked length spectrum, derived from the trace formula's orbital contributions, they must be , providing a geometric counterpart to isospectrality questions. Applications of the Selberg trace formula extend to symmetric spaces, where it informs rigidity theorems for lattices in higher-rank semisimple groups. Gregory Margulis's superrigidity , developed in the and refined in subsequent works, leverages and orbital data from trace formula analogs to show that irreducible lattices in such groups are when the rank is at least two, implying strong geometric constraints on the quotients. This has profound implications for the structure of locally symmetric spaces, where the formula distinguishes arithmetic from non-arithmetic cases through eigenvalue distributions and geodesic counts. Numerical computations leveraging the Selberg trace formula have illuminated differences between arithmetic and non-arithmetic hyperbolic surfaces since the early 2000s. For instance, Andrew Booker and Andreas Strömbergsson used the formula in 2007 to numerically verify the Selberg eigenvalue conjecture for the and subgroups, computing traces and comparing them against predictions, which verified conjectural equidistribution while highlighting subtle discrepancies in these arithmetic examples compared to predictions from random matrix theory for generic non-arithmetic surfaces. These calculations, involving efficient summation over primitive geodesics, demonstrate the formula's practical power in distinguishing geometric chaos patterns, with arithmetic spectra showing enhanced symmetries absent in generic cases.

References

  1. [1]
    [PDF] An Introduction to the Trace Formula - Clay Mathematics Institute
    In §1 we review the Selberg trace formula for compact quotient. In §2 we introduce the ring A = AF of adeles.
  2. [2]
    [PDF] Introductory Notes on the Trace Formula1
    Thus, viewed as an identity of distributions on h the Selberg trace formula is similar in shape to the Poisson summation formula. The non-identity geometric.
  3. [3]
    [PDF] A (very brief) History of the Trace Formula James Arthur This note is ...
    A decisive answer was provided by the trace formula Selberg created to this end. The Selberg trace formula is an identity. (STF). Xi aig(ui) = Xj bj g(λj) + e ...
  4. [4]
    Atle Selberg - Biography - MacTutor - University of St Andrews
    The necessary analytic tools were known by 1896 when Jacques Hadamard and Charles de la Vallée Poussin independently proved the theorem using complex analysis.
  5. [5]
    Harmonic analysis and discontinuous groups in weakly symmetric ...
    Selberg in 1956. We prove that, for a real reductive algebraic group, they can be characterized as the spaces of real points of affine… Expand. 70 Citations.Missing: Harmonie | Show results with:Harmonie
  6. [6]
    Physics Asymptotics of the Spectrum and the Selberg Zeta Function ...
    Selberg zeta functionZ(s, Sk) is defined by the infinite product over geodesic lengths. 00. Z(s,Sk) = Γi Π (l-e~(5+k)'ω)2 for Re5>l γ k=0 and has a ...
  7. [7]
    Selberg zeta function in nLab
    May 22, 2019 · The Selberg zeta function controls the asymptotics of prime geodesics via the prime geodesic theorem in direct analogy to how the Riemann zeta function ...
  8. [8]
    [PDF] The Eichler-Selberg Trace Formula for Level-One Hecke Operators
    May 20, 2013 · This paper explains the steps involved in the proof of the Eichler-Selberg Trace. Formula for Hecke operators of level one.Missing: Roch | Show results with:Roch
  9. [9]
    [PDF] proceedings of the international congress of mathematicians
    SELBEBG, A., Discontinuous groups and harmonic analysis. SEBBE, J.-P ... 1956, the mathematical world was astounded by Milnor's proof that the 7 ...
  10. [10]
    [PDF] Eisenstein Series, the Trace Formula, and the Modern Theory of ...
    The spectral theory has two aspects: (i) the spectral decomposition of the spaces L2(Γ\G) by means of Eisenstein series; (ii) the trace formula, which can be ...Missing: legged sieve 1973
  11. [11]
    A trace formula for reductive groups. II : applications of a truncation ...
    Arthur, James. "A trace formula for reductive groups. II : applications of a truncation operator." Compositio Mathematica 40.1 (1980): 87-121. <http:// ...Missing: 1990s | Show results with:1990s
  12. [12]
    [PDF] James ARTHUR
    This paper is a report on the present state of the trace formula for a general reductive group. The trace formula is not so much an end.Missing: 1990s | Show results with:1990s
  13. [13]
    On some problems suggested by the trace formula - SpringerLink
    Sep 12, 2006 · Arthur, J. (1983). On some problems suggested by the trace formula. In: Herb, R., Kudla, S., Lipsman, R., Rosenberg, J. (eds) Lie Group Representations II.
  14. [14]
    [PDF] A stable trace formula I. General expansions
    This paper aims to stabilize the global trace formula for a general connected group, investigating the underlying structure of the process.
  15. [15]
    [PDF] Selberg's Trace Formula: An Introduction - University of Bristol
    The aim of this short lecture course is to develop Selberg's trace formula for a compact hyperbolic surface M, and discuss some of its applications. The.
  16. [16]
    [PDF] the selberg trace formula
    Introduction. Selberg developed his trace formula in [Se] as a natural non-commutative gen- eralization of the Poisson summation formula.
  17. [17]
    [PDF] Chapter 2: Hyperbolic Geometry - The University of Chicago
    Feb 15, 2019 · The stabilizers of geodesics are the subgroups conjugate to SO(n − 1) × A which is equal to SO(n − 1) × SO+(1,1) or. SO(n − 1) × SO(2).
  18. [18]
    [PDF] Spectral Theory on Hyperbolic Surfaces
    In the context of automorphic forms, eigenvectors of the Laplacian are called Maass forms. 0 = λ0 < λ1 ≤ λ2 ≤···→∞. Stone's formula gives spectral projectors ...
  19. [19]
    [PDF] Spectra of Hyperbolic Surfaces - Princeton Math
    All these realizations of X(N) are important. As a quotient space the modular surface X(1) = Γ(1)\H looks like. Figure 1. This can ...
  20. [20]
    [PDF] automorphic forms and the selberg trace formula
    Feb 24, 2021 · In this paper, we explain the use of the Selberg trace formula in this setting, where it may be used to relate the spectral theory of the.
  21. [21]
    [PDF] The Contest Between the Kernels in the Selberg Trace Formula for ...
    So the heat kernel is the fundamental solution of the heat equation on TA. This makes it the analogue of the heat kernel on RD or the Poincaré upper half plane ...
  22. [22]
    [PDF] plancherel formula for a real reductive lie group, xiamen, july 29
    They are unitary representations except the highest and lowest weight representations. ... Harish-Chandra's Plancherel formula. It is not difficult to check that ...
  23. [23]
    [PDF] An introduction to Selberg's trace formula
    Now let us consider the Schwartz space S of functions f: R→ Csuch that for ... Remark: H is called the Harish-Chandra transform of 0. The proof can be ...Missing: test | Show results with:test
  24. [24]
    [PDF] On the Harish-Chandra Schwartz space - Erwin Schrödinger Institute
    Dec 30, 2012 · We study the Harish-Chandra Schwartz space of an adelic quotient G(F)\G(A). We state a conjectural spectral decomposition of it in terms of ...Missing: test | Show results with:test
  25. [25]
    [math/0407288] Selberg's trace formula: an introduction - arXiv
    Jul 16, 2004 · These lecture notes provide a basic introduction to Selberg's trace formula. We discuss the simplest possible case: the spectrum of the Laplacian.<|control11|><|separator|>
  26. [26]
    [PDF] The Selberg trace formula for groups of F-rank one
    An important tool for the study of automorphic forms is a non- abelian analogue of the Poisson summation formula, generally known as the Selberg trace ...
  27. [27]
  28. [28]
    [PDF] An introduction to the trace formula
    May 11, 2010 · We now give an explicit example of the trace formula. Suppose ... Selberg trace formula, Uni- versity Lecture Series, vol. 9, American ...<|control11|><|separator|>
  29. [29]
    The Selberg Trace Formula for PSL (2,R): Volume 1 - SpringerLink
    Free delivery 14-day returnsThe Selberg Trace Formula for PSL (2,R) ; 1st edition; View latest edition ; Softcover Book USD 37.99. Price excludes VAT (USA) ; Explore related subjects.
  30. [30]
    [PDF] trace formulae for sl(2,r) - dorian goldfeld - Columbia Math Department
    Let s = (1/2 + s1,−1 − 2s1) with s1 ∈ C and let α = (s1 − 1/2,s1 + 1/2,−2s1). The Langlands Eisenstein series determined by this data is defined by.
  31. [31]
    [PDF] The Trace Formula for Noncompact Quotient
    f Zf(xlyx)dx = tr R(f). r\Gver. Selberg's formula is obtained by grouping together those elements in r with the same eigenvalues ...Missing: 1960 | Show results with:1960
  32. [32]
    [PDF] The trace formula in invariant form - Clay Mathematics Institute
    The object of this paper is to derive a trace formula whose terms are invariant. We want an identity which is of the form (2.5) but such that the.
  33. [33]
    [PDF] SOME EICHLER-SELBERG TRACE FORMULAS - OpenScholar
    The Eichler-Selberg trace formulas express the traces of Hecke operators on a spaces of cusp forms in terms of weighted sums of Hurwitz-Kronecker class numbers.Missing: Roch | Show results with:Roch
  34. [34]
    SHIMURA VARIETIES AND THE SELBERG TRACE FORMULA
    It is first necessary to decide which H, T, and r should intervene in this product, and with what exponent m and what translation a each L-function should.
  35. [35]
    [PDF] THE PRIME GEODESIC THEOREM Contents 1. Introduction 1 2 ...
    Aug 18, 2023 · The Selberg trace formula provides an infinite dimensional representation homomorphism of the group to matrices. Theorem 5.1 (Selberg trace ...
  36. [36]
    [PDF] Coefficients of Maass Forms and the Siegel Zero Author(s)
    upper bound for p(l) is therefore equivalent to proving the nonexistence of a. Siegel zero of L(s, F). ... IWANIEC, The non-vanishing of Rankin-Selberg zeta- ...
  37. [37]
    [PDF] arXiv:1011.5486v1 [math.NT] 24 Nov 2010
    Nov 24, 2010 · There are two parts to the proof: first, using the Selberg trace formula we handle smoothed sums over prime geodesics, and second, using ...
  38. [38]
    Unconditional computation of the class groups of real quadratic fields
    We describe an algorithm, based on the Selberg trace formula and explicit numerical computations of Maaß cusp forms, for computing the class groups and ...
  39. [39]
    [PDF] arXiv:2112.14396v3 [math.NT] 26 Apr 2022
    Apr 26, 2022 · We first briefly review the central tools in this paper—the Kuznetsov trace formula for GL(2), the functional equation and the Voronoi summation for- mula for ...
  40. [40]
    [PDF] Bounds for modular L-functions in the level aspect. - EPFL
    Oct 31, 2018 · culminating in a subconvexity bound for general automorphic forms on GL2 [DFI5]: ... In order to satisfy the decay conditions for Kuznetsov's trace formula, we ...
  41. [41]
    [PDF] Isospectral Riemann surfaces - Numdam
    Early results in the affirmative direction were that Riemann surfaces do not admit isospectral continuous deformations [4], [14], and later by McKean that for a ...
  42. [42]
    Geometry and Spectra of Compact Riemann Surfaces - SpringerLink
    Book Title: Geometry and Spectra of Compact Riemann Surfaces · Authors: Peter Buser · Series Title: Modern Birkhäuser Classics · Publisher: Birkhäuser Boston, MA.
  43. [43]
    Quantum Chaos of the Hadamard-Gutzwiller Model | Phys. Rev. Lett.
    Aug 1, 1988 · The theory is based on periodic-orbit sum rules that can be rigorously derived from the Selberg trace formula and which provide an exact ...
  44. [44]
    [PDF] The Gutzwiller Trace Formula and the Quantum-Classical ... - Stanford
    Jun 15, 2020 · In this expository article, we review the periodic-orbit quantization method and its use in verifying the BGS (Bohigas-Giannoni-Schmit) ...
  45. [45]
    [PDF] lectures on marked length spectrum rigidity (preliminary version)
    Dec 11, 2012 · In these notes, I'll explain in several steps a proof of this marked length spectrum rigidity for negatively curved surfaces: Theorem 0.1 (Otal) ...