Fact-checked by Grok 2 weeks ago

Monodromy

In mathematics, particularly in and , monodromy refers to the transformation that occurs when branches of a multi-valued , such as the roots of a , are analytically continued along a closed in the that encircles singularities or branch points, resulting in a of those branches. This phenomenon, which highlights the path-dependent nature of such continuations, is formalized through the monodromy group, a subgroup of the generated by these permutations, capturing the topological structure of the function's . The concept of monodromy emerged in the mid-19th century amid foundational developments in , building on Bernhard Riemann's 1851 introduction of Riemann surfaces to resolve multi-valuedness in functions like the logarithm or . formalized the idea in his 1868 lectures, where he proved the : if a element defined on a disk can be analytically continued along every piecewise smooth path within a simply connected domain without encountering singularities, then these continuations define a single-valued on the entire domain. This theorem, later generalized by in 1883 and fully published in 1922, underscores that multi-valuedness arises from the domain's non-trivial topology, specifically its . Beyond its role in analytic continuation, monodromy has profound implications across mathematics. In algebraic geometry, monodromy groups determine properties of polynomial equations, such as solvability by radicals via the Abel-Ruffini theorem, where insoluble cases correspond to non-solvable groups like the alternating group A_5. In the theory of linear differential equations, the monodromy group describes how solutions permute around singular points on the , aiding in the classification of equations and their global behavior. More broadly, in topology and dynamical systems, iterated monodromy groups serve as algebraic invariants for rational maps on the , influencing studies from to .

Fundamental Concepts

Definition

In complex analysis, functions are classified as single-valued if analytic continuation along any closed path in the domain returns the function to its original value at the base point, whereas multivalued functions exhibit path-dependent behavior, returning to a different branch or value due to the presence of branch points or singularities in the domain. This distinction arises naturally when dealing with solutions to algebraic equations or integrals of holomorphic forms, where the topology of the punctured plane or surface influences the possible values. Monodromy quantifies this path dependence by describing the transformations induced on the set of function values—or more generally, on the fiber of a vector bundle—when continuing along loops in the fundamental group of the domain. Formally, consider a multivalued f defined on a punctured domain U \subset \mathbb{C} (or more generally, a with punctures), with a base point z_0 \in U. The over z_0 is the finite or infinite set of possible values of f at z_0, reflecting the branches of the function. The \pi_1(U, z_0) acts on this through : for a \gamma based at z_0, continuing f along \gamma induces an of the , known as the monodromy transformation associated to \gamma. This action defines a \rho: \pi_1(U, z_0) \to \mathrm{Aut}(F_{z_0}), where F_{z_0} denotes the , turning the monodromy into a from the to the group of automorphisms of the . The basic notation for monodromy operators reflects this group structure: for loops \gamma, \delta \in \pi_1(U, z_0), the operator M_\gamma satisfies M_\gamma \circ M_\delta = M_{\gamma \delta}, where \gamma \delta denotes the of paths (with the such that \delta follows \gamma). This ensures that the monodromy respects the of the base space, capturing how encircling singularities permutes or linearly transforms the branches of f. In the context of vector-valued functions or solutions to systems, the automorphisms lie in \mathrm{GL}(n, \mathbb{C}) for dimension n. The concept of monodromy was introduced by in 1857, in his study of hypergeometric functions, where he analyzed the linear transformations undergone by solutions upon encircling singular points.

Illustrative Example

A classic illustrative example of monodromy arises in the analytic continuation of the principal function \sqrt{z} around its at z = 0. Consider the principal branch defined for z = re^{i\theta} with r > 0 and -\pi < \theta \leq \pi, given by \sqrt{z} = \sqrt{r} e^{i\theta/2}, which assigns positive values to positive real numbers. Starting at a point z = r > 0 on the positive real axis, where \sqrt{r} is the positive real , we analytically continue this function along a closed path encircling the origin counterclockwise. To compute this explicitly, parameterize the path as \gamma(t) = r e^{2\pi i t} for $0 \leq t \leq 1, a circle of radius r centered at the origin. The analytic continuation of \sqrt{z} along this path begins with \sqrt{\gamma(0)} = \sqrt{r} e^{i \cdot 0} = \sqrt{r}. As t increases, the argument of \gamma(t) is $2\pi t, so the continued value is \sqrt{\gamma(t)} = \sqrt{r} e^{i \pi t}. At t = 1, this yields \sqrt{r} e^{i \pi} = -\sqrt{r}, the negative of the starting value. Thus, completing the loop transforms the function value to its negative, demonstrating the monodromy effect of switching branches. This behavior is visualized using branch cuts, typically a ray from the along the negative real , which prevents crossing between in the principal domain. A full counterclockwise around z = 0 crosses the cut effectively, resulting in negation of the value, while a half-loop (e.g., from the positive real axis to the negative real axis without encircling the ) remains within one branch and produces no such . In contrast, a clockwise full would again negate the value, but two full loops in either direction return to the original . This example generalizes to the n-th root function z^{1/n}, which has n branches and a branch point at z = 0. Analytic continuation around a full counterclockwise loop rotates the value in the fiber by e^{2\pi i / n}, cycling through the branches, with n loops required to return to the starting value.

Complex Analysis Context

Analytic Continuation and Multivaluedness

Analytic continuation extends the domain of an analytic function f defined on an open set D \subset \mathbb{C} by constructing a sequence of analytic functions that agree on overlapping regions along a specified path \gamma: [0,1] \to \Omega, where \Omega is a domain containing D. The maximal analytic continuation along \gamma starts with a function element (f_0, D_0) near \gamma(0) and proceeds by successively extending to disks centered at points along \gamma, ensuring overlaps where the functions coincide. This process yields a continued function f_\gamma analytic near \gamma(1), but the result depends on the homotopy class of \gamma in \Omega: if two paths \gamma_0 and \gamma_1 from the same starting point are homotopic with fixed endpoints, their continuations agree near the endpoint, as established by the monodromy theorem. The continuation along \gamma satisfies the differential relation \frac{df}{dz} = f'(z) in local coordinates, where f' is the of the original , preserving analytic properties through overlaps; however, the global value of the continued at the endpoint depends on the initial value and the class of the path, potentially differing for non-homotopic paths. For instance, the \sqrt{z} illustrates this, as continuation around the yields the negative . Multivalued functions, such as the , arise when along non-contractible loops in the domain yields distinct values, rendering the function single-valued only on a of the punctured domain. In this framework, the function lifts to a single-valued on the universal cover, and the monodromy describes the deck transformations—automorphisms of the cover that permute the sheets corresponding to loops in the base space, isomorphic to the . These transformations capture how encircling certain points shifts the function values among its branches. In the punctured plane \mathbb{C} \setminus \{0\} or on the with isolated singularities, such as or , non-trivial monodromy emerges when around these points induce permutations of branches via continuation. For example, an like that of \log z at z=0 causes a shift by $2\pi i upon traversing a , while a may lead to trivial monodromy for meromorphic functions but non-trivial effects in branched settings; this reflects the topological obstruction to single-valued extension across the .

Riemann Surfaces and Branch Points

Riemann surfaces provide a geometric framework to resolve the multivaluedness arising from monodromy in , constructed as covering spaces over the or to make analytic continuations single-valued. Specifically, for a like the \log z, the is built as the universal cover of the punctured plane \mathbb{C} \setminus \{0\}, realized as a helical surface where each "sheet" corresponds to a branch of the logarithm, and the covering map is the \exp: \mathbb{C} \to \mathbb{C}^*. This construction ensures that paths encircling the origin lift to non-closed paths on the surface, eliminating the ambiguity in the function's value upon continuation around loops. Branch points on the base domain classify the local behavior of these coverings and the associated monodromy. Algebraic branch points occur where the monodromy has finite order, as in the nth root function z^{1/n}, leading to a finite-sheeted branched cover with n sheets ramified at the origin; here, the local monodromy is represented by a cycle of length n in the permutation of the sheets. In contrast, logarithmic branch points, exemplified by \log z at z=0, exhibit infinite-order monodromy, resulting in an infinite-sheeted covering with no ramification but an essential singularity in the covering space. These distinctions arise from the Puiseux expansions near the points: algebraic branches admit finite power series solutions, while logarithmic ones involve terms like \log z. On the Riemann surface, the original multivalued function becomes single-valued and holomorphic, except possibly at ramification points, normalizing the analytic structure. The deck transformation group of this covering, consisting of automorphisms that permute the sheets while preserving the fibers, is isomorphic to the monodromy group generated by the local actions around branch points; for the logarithmic example, this group is \mathbb{Z}, reflecting the additive shifts in the argument. This isomorphism links the topological covering properties directly to the algebraic structure of the monodromy representation.

Differential Equations Applications

Linear Ordinary Differential Equations

In the complex domain, consider a linear system of ordinary differential equations given by Y' = A(z) Y, where Y is an n-dimensional and A(z) is an n \times n with meromorphic entries on a connected open subset U \subseteq \mathbb{C}. The singularities of the system occur at the poles of A(z), and the space of solutions forms a over U, the parameter space excluding these singularities. Analytic continuation of a fundamental matrix along a closed \gamma in U, based at a point p \in U, induces a linear on the space. Specifically, if Y denotes a local , then after continuation along \gamma, the continued satisfies Y(\gamma(1)) = M_\gamma Y(\gamma(0)), where M_\gamma \in \mathrm{GL}(n, \mathbb{C}) is the monodromy matrix associated to \gamma. This defines a monodromy \pi_1(U, p) \to \mathrm{GL}(n, \mathbb{C}), with the image forming the monodromy group. Key properties of the monodromy matrices arise from the topology and structure of the . For a regular singular point, the eigenvalues of M_\gamma are e^{2\pi i \alpha_j}, where the \alpha_j are the indicial roots determining the local behavior near the ; consequently, the of M_\gamma encodes the of these terms, reflecting the of the . Moreover, for a Fuchsian on the with finitely many singularities (including at infinity), the product of the monodromy matrices corresponding to loops around each is the , mirroring the for meromorphic forms. A example illustrates these concepts for a scalar second-order , y'' + \frac{1}{z} y' = 0, with a regular singular point at z = 0. The general solution is y(z) = c_1 + c_2 \log z, where the constant solution is single-valued and the logarithmic term acquires an increment of $2\pi i upon encircling the . In form, with basis solutions \{1, \log z\}, the monodromy for a positive around z = 0 is M_\gamma = \begin{pmatrix} 1 & 2\pi i \\ 0 & 1 \end{pmatrix}, whose eigenvalues are both 1 (indicial roots 0, 0) and 2, consistent with the trivial index at this singularity.

Fuchsian Systems and Regular Singularities

Fuchsian systems are a special class of linear ordinary differential equations on the where all singularities are regular. Specifically, a system of the form \frac{dY}{dz} = A(z) Y, with Y a and A(z) a of rational functions, is Fuchsian if A(z) has poles of order at most 1 at finite singular points and of order at most 2 at infinity. This condition ensures that the only singularities are regular singular points, allowing for controlled analytic behavior near these points. Near a regular singular point, say at z = 0, local solutions can be constructed using the , which assumes a of the form Y(z) = z^\rho \sum_{k=0}^\infty a_k z^k, where \rho is determined by the indicial equation \det(\rho I - \operatorname{Res}_0 A) = 0. The roots \rho of this equation give the exponents governing the leading-order behavior, and the series coefficients a_k are recursively determined, yielding a fundamental set of solutions in a punctured neighborhood of the . These local solutions form the basis for understanding the global around the singularities. The monodromy matrix around a regular has eigenvalues of modulus one, e^{2\pi i \alpha_j}, where the \alpha_j are the indicial modulo the integers. When the differences between indicial are integers, the matrix may have Jordan blocks corresponding to logarithmic terms. The Jordan form of M is determined by the differences between the eigenvalues of the residue matrix at the , reflecting the logarithmic terms that may arise in the solutions when indicial differ by integers. In cases where the indicial are rational, such as in rigid local systems or algebraic , the monodromy is quasi-unipotent. Globally, the Riemann-Hilbert problem for Fuchsian systems establishes a between the monodromy representation—given by the collection of these local monodromy matrices—and the of the differential system. This , under suitable irreducibility conditions, allows reconstruction of the Fuchsian system from prescribed monodromy data with regular singularities at specified points. In the context of general linear ODEs, the monodromy matrices for Fuchsian systems inherit these properties while restricting to finite-order poles in the coefficients.

Topological and Geometric Frameworks

Monodromy Representations

In the topological framework, the monodromy representation arises as a \rho: \pi_1(X, x_0) \to \Aut(F), where X is a base space such as the minus a of points, x_0 \in X is a base point, \pi_1(X, x_0) is its , and F is the fiber over x_0 consisting of local sections (such as solutions to a defined on X). This representation encodes the action of loops in \pi_1(X, x_0) on F via along the loop, yielding an of F that is independent of the choice of continuation path within the class. For linear equations with rational coefficients on X, the fiber F is a finite-dimensional complex , and \Aut(F) \cong \GL(n, \mathbb{C}) for appropriate n, making \rho a linear . A monodromy representation \rho is faithful if it is injective, meaning the image \rho(\pi_1(X, x_0)) is isomorphic to \pi_1(X, x_0) and fully captures the topological structure of X through the on F; otherwise, it is unfaithful, with a proper of the embedded in \Aut(F). often holds in generic settings, such as for the universal family of algebraic curves, where the representation distinguishes distinct classes. In contrast, unfaithful representations may arise when the action factors through a , losing information about certain loops. The choice of base point x_0 affects the representation up to conjugation in \Aut(F): if x_1 is another base point connected by a path \gamma, then \rho_{x_1}(\delta) = M_\gamma^{-1} \rho_{x_0}(\delta) M_\gamma for \delta \in \pi_1(X, x_1), where M_\gamma \in \Aut(F) is the monodromy along \gamma, ensuring the image group is well-defined up to isomorphism. A canonical example occurs on the punctured plane X = \mathbb{C} \setminus \{0\}, where \pi_1(X, x_0) \cong \mathbb{Z} is generated by a loop \gamma encircling the origin once. The monodromy representation sends the generator [\gamma] to powers of a single matrix M \in \GL(n, \mathbb{C}), such as \rho(k[\gamma]) = M^k, reflecting the cyclic action; for the branched cover given by f(z) = \sqrt{z}, this yields M as a transposition in S_2 \cong \Aut(F), interchanging the two branches upon looping around the puncture.

Monodromy Group and Groupoid

The monodromy group arises as the G = \rho(\pi_1(X, x_0)) of the monodromy \rho: \pi_1(X, x_0) \to \mathrm{GL}_n(\mathbb{C}), where X is the complement of singular points in the or , and \rho encodes the of solutions to a along loops based at x_0. This group is generated by the monodromy matrices corresponding to loops encircling individual singularities, with relations imposed by higher classes in \pi_1(X, x_0). For Fuchsian systems, where all singularities are , the monodromy group is finitely generated by these local loops. Schlesinger's theorem further specifies that the Picard-Vessiot group, the smallest Zariski-closed containing G, coincides with the generated by the monodromy when singularities are . These properties highlight the group's role in capturing global topological constraints on local analytic behaviors. The monodromy groupoid extends this framework to account for varying base points, providing a richer structure that tracks path dependencies across the entire X rather than fixing a single base. Defined for a topological or groupoid G as the of covers of its stars (source fibers), the monodromy groupoid \Pi G projects onto G and inherits a structure under suitable smoothness conditions, such as paracompact objects and path-connected stars. This construction embodies the monodromy principle, allowing local morphisms near identities to lift globally, thus generalizing the representation-theoretic view to a category of paths and continuations. In the context of foliations, the monodromy groupoid formalizes within integrable distributions on a manifold, where leaves serve as paths for of local sections. Here, the arises from the equivalence relation of leafwise paths, yielding a whose arrows represent transverse maps between nearby leaves. This structure captures the full dependency of continuations on base points within the , distinguishing it from the coarser holonomy by incorporating higher data.

Advanced Connections

Differential Galois Theory

provides an algebraic framework for studying the solvability of linear equations, extending classical to differential fields. For a linear homogeneous L(y) = 0 over a differential field k equipped with a \delta, the Picard-Vessiot extension is the smallest differential field extension K/k generated by a fundamental set of solutions to L(y) = 0, along with constants if necessary to ensure of the constants of k. The \mathrm{Gal}(K/k) consists of the differential automorphisms of K that fix k pointwise, acting linearly on the solution space and forming a over the constants of k. This group encodes the algebraic and differential dependencies among the solutions, analogous to how the in algebraic theory describes root relations. In the context of analytic solutions on Riemann surfaces, the monodromy group arises from the representation of the of the punctured plane (or more generally, the base space minus singularities) on the solution space via . This monodromy group embeds as a of the differential , since analytic continuations preserve the differential relations among solutions. Specifically, for equations with regular singular points, the monodromy group is Zariski dense in the differential by Schlesinger's density theorem, meaning its Zariski closure coincides with the full . When the monodromy group generates the entire differential (up to Zariski closure), and the latter is solvable, the equation admits solutions by quadratures, i.e., expressible in terms of integrals, exponentials, and algebraic functions over the base field—Liouvillian solutions. The Kovacic algorithm offers a computational to determine whether a second-order over a of rational functions has Liouvillian solutions, by classifying the possible Galois groups into six cases based on their structure (reducible, imprimitive, or specific connected components). In particular, cases involving unipotent monodromy correspond to scenarios where the has a unipotent radical, indicating solutions with logarithmic terms, which are Liouvillian but non-elementary. The algorithm checks for the existence of such structures without computing the full extension, providing explicit criteria like the order of poles in the companion form. A representative example is Bessel's equation x^2 y'' + x y' + (x^2 - \nu^2) y = 0, which has a regular singular point at x=0. For non-integer \nu, the monodromy around x=0 is diagonal with eigenvalues e^{\pm 2\pi i \nu}, generating a non-solvable SL_2(\mathbb{C}) differential , implying no Liouvillian solutions and thus non-elementary functions like the J_\nu(x). For \nu, the monodromy becomes unipotent, introducing logarithmic terms in solutions like the Weber function, but the remains non-solvable unless \nu - 1/2 is an , in which case elementary solutions exist.

Riemann-Hilbert Correspondence

The Riemann-Hilbert correspondence establishes a profound duality between the topological data encoded in monodromy representations of the fundamental group and the analytic data of holomorphic connections on vector bundles with regular singularities. In its classical formulation on the Riemann sphere punctured at finitely many points a_1, \dots, a_k \in \mathbb{C}, it asserts that there is a bijection between isomorphism classes of representations \rho: \pi_1(\mathbb{C} \setminus \{a_1, \dots, a_k\}, x_0) \to \mathrm{GL}(n, \mathbb{C}) where the local monodromies around each puncture (and at infinity) are quasi-unipotent—meaning their eigenvalues are roots of unity—and the isomorphism classes of rank-n holomorphic vector bundles on the compactification \mathbb{P}^1 equipped with flat meromorphic connections having regular singularities at \{a_1, \dots, a_k, \infty\}. This result, originally motivated by Hilbert's 21st problem on the existence of linear differential equations with prescribed singularities and monodromy group, resolves the inverse problem affirmatively under these conditions, with the direct problem (extracting monodromy from a connection) always holding by definition via analytic continuation of fundamental solutions. A sketch of the construction proceeds by first realizing the locally near each . For a puncture at a_j, the local monodromy \rho(\gamma_j) around a small \gamma_j admits a logarithm M_j = \log \rho(\gamma_j) such that \exp(2\pi i M_j) = \rho(\gamma_j), possible since \rho(\gamma_j) is quasi-unipotent. One then builds a -n holomorphic vector bundle E on \mathbb{P}^1 by gluing local trivializations, leveraging the Birkhoff-Grothendieck theorem which decomposes any on \mathbb{P}^1 as a of line bundles. The \nabla = d + A(z) dz is defined with residue matrices A(a_j) satisfying \exp(2\pi i A(a_j)) = \rho(\gamma_j), ensuring the global monodromy matches \rho while maintaining regular singularities (simple poles in A(z)). This yields a unique Fuchsian system up to , as the residues determine the connection locally and the bundle globally. Deligne's theorem extends this bijection to arbitrary smooth algebraic varieties over \mathbb{C}, equating bounded complex local systems (with quasi-unipotent monodromies) to algebraic vector bundles with flat connections of moderate growth, without restricting to the sphere; for higher-genus surfaces, the construction involves étale covers and cohomology to handle non-trivial topology. In the irregular singularity case, partial analogs exist via the Birkhoff-Grothendieck theorem for formal meromorphic bundles on the formal disk, allowing decomposition into formal line bundles and construction of formal fundamental solutions, though global holomorphic realizations remain challenging and are resolved only under additional slope conditions or for specific ranks. Applications abound in , where the correspondence enables reconstruction of stable parabolic bundles from monodromy data, facilitating the study of moduli spaces of connections and their relation to varieties of representations. In modern integrable systems, it underpins isomonodromic deformation theory for equations like the , where solutions are parameterized by monodromy data through asymptotic Riemann-Hilbert problems on contours, yielding explicit uniformization maps between Painlevé fibers and monodromy manifolds.

References

  1. [1]
    [PDF] a survey on the monodromy groups of algebraic functions
    Monodromy studies how objects 'run round' a singularity, originating from the switching of single-valued functions when going around a point.
  2. [2]
    [PDF] arXiv:1507.00711v1 [math.AG] 2 Jul 2015
    Jul 2, 2015 · Abstract. We discuss the history of the monodromy theorem, starting from Weierstraß, and the concept of monodromy group.
  3. [3]
    [PDF] Riemann Surfaces - Berkeley Math
    Then, ˜U is homeomorphic to U. Proof of the Monodromy Theorem: Let (R, r, π, F) be the Riemann surface of f, π(r) = a. Let ˜U be the component of π−1(U) ...
  4. [4]
    [PDF] monodromy groups of parameterized linear differential equations ...
    To define the monodromy group one starts by removing the set S = {a1,...,as} of singular points (possibly including infinity) of (1.1) from the Riemann sphere P1 ...
  5. [5]
    [PDF] iterated monodromy groups
    1. Introduction. Iterated monodromy groups are algebraic invariants of topological dynamical systems (e.g., rational functions acting on the Riemann sphere). ...
  6. [6]
    [PDF] Lectures on Riemann surfaces - staff.math.su.se
    Riemann surfaces originated in complex analysis as a means of dealing with the problem of multi-valued functions. Such multi-valued functions occur because ...
  7. [7]
    [PDF] Beiträge zur Theorie der durch die Gauss'sche Reihe F(α,β,γ,x ...
    Die Gauss'sche Reihe F(α,β,γ,x), als Function ihres vierten Elements x betrachtet, stellt diese Function nur dar, so lange der Modul von x die Einheit.Missing: Funktionen | Show results with:Funktionen
  8. [8]
  9. [9]
    [PDF] 2.4 BRANCHES OF FUNCTIONS
    The point z = 0, common to all branch cuts for the multivalued square root function, is called a branch point. The mapping w = fα (z) and its branch cut are ...
  10. [10]
    Analytic Continuation -- from Wolfram MathWorld
    For example, consider analytic continuation of the square root function f(z)=sqrt(z) . Although this function is not globally well-defined (since every ...
  11. [11]
    Multivalued Function -- from Wolfram MathWorld
    ### Summary on nth Roots and Monodromy/Continuation Around Branch Points
  12. [12]
    [PDF] Complex Analysis I, Christopher Bishop 2024 - Stony Brook University
    If g can be analytically continued along all curves in f(Ω) and if f(Ω) is simply- connected, then by the monodromy theorem there is a function G which is.<|control11|><|separator|>
  13. [13]
    [PDF] IX.3. Monodromy Theorem.
    Apr 9, 2017 · Note. Before stating and proving the Monodromy Theorem, we need two lemmas and a definition. The first lemma concerns radius of convergence.
  14. [14]
    [PDF] ANALYTIC CONTINUATION - UCI Mathematics
    ANALYTIC CONTINUATION ALONG A PATH. 49. Definition 3.4. Consider γ. ∗. : [a, b] → D with γ. ∗. (a) = w0. Call it a lift. (relative to g) of γ (based at w0) if g ...
  15. [15]
    [PDF] A concise course in complex analysis and Riemann surfaces ...
    The monodromy theorem of course implies that on simply connected Riemann sur- faces M any function element (f,D) that can be analytically continued ...
  16. [16]
    [PDF] Riemann Surfaces
    Jun 17, 2018 · Riemann surfaces were originally conceived in complex analysis in order to deal with multivalued functions. The analytic continuation of a given ...
  17. [17]
    [PDF] Differential equations and monodromy - the MPIM Archive
    The monodromy of the equation acts on the space of solutions Y ∗ by the formula Y ∗(τ + 1) = Y ∗(τ)M where M ∈ GLn(C).
  18. [18]
    [PDF] Complex analytic ordinary differential equations
    Aug 9, 2017 · This linear map from. Cn to itself is called the monodromy of the path. It follows from Proposition 1.2 that it depends only on its homotopy ...
  19. [19]
    [PDF] arXiv:0905.1436v1 [math.CA] 9 May 2009
    May 9, 2009 · Inversely, points (a, u, v) corresponding to Fuchsian equations with the same monodromy lie on the integral manifold of the system Gn(θ). Using ...
  20. [20]
    [PDF] Fuchsian Differential Equations
    Aug 10, 2020 · Definition. A homogeneous linear ODE is called Fuchsian if all of its coefficients are rational functions and all of its (finitely many).
  21. [21]
    Fuchsian equations—differential equations' singularities - IOP Science
    Fuchs wrote down an equation that was later called the indicial equation, the roots of which determine the behavior of solutions near a regular singular point.
  22. [22]
    The method of Frobenius to Fuchsian partial differential equations
    The method of Frobenius, Fuchsian partial differential equation, regular singularity, characteristic exponent, characteristic index. The research was partially ...
  23. [23]
    [PDF] arXiv:1108.1847v2 [math.CA] 8 Sep 2011
    Sep 8, 2011 · We say that a regular system (1) has quasiunipotent monodromy (or simply is quasiunipotent), if the monodromy operators Mj of all small loops ...
  24. [24]
    [PDF] Nilpotent connections and the monodromy theorem - Math (Princeton)
    — If p is any place of K/A; {not necessarily rational), at which (W, V) has a regular singular point, we say that the local monodromy at p is quasi-unipotent ( ...
  25. [25]
    [PDF] The Riemann-Hilbert problem
    'Prove that there always exists a linear differential equation of Fuchsian type with given singular points and with a given monodromy group.' A tradition has ...
  26. [26]
    [PDF] The Riemann-Hilbert Problem
    Jul 1, 2024 · In his 1857 paper Riemann suggested to study the problem of constructing a system of functions with regular singularities that has the ...
  27. [27]
    The faithfulness of the monodromy representations associated with ...
    We consider the faithfulness of the monodromy representation associated with the universal family of n-pointed algebraic curves of genus g (2−2g−n<0).
  28. [28]
    [PDF] differential equations associated with nonarithmetic fuchsian groups
    We describe globally nilpotent differential operators of rank 2 defined over a number field whose monodromy group is a nonarithmetic Fuchsian group. We show.
  29. [29]
    [PDF] The monodromy groupoid of a Lie groupoid - Ronald Brown
    local conditions of covering space theory, the monodromy groupoid IIG. is the universal covering group, while if G is the groupoid X x X,
  30. [30]
  31. [31]
    [PDF] 1 Introduction to the Galois Theory of Linear Differential Equations
    Jan 10, 2008 · The Fundamental Theorem also follows from a deeper fact giving geo- metric meaning to the Picard-Vessiot ring and relating this to the Galois.<|control11|><|separator|>
  32. [32]
    [PDF] Lecture 3: Introduction to Galois theory of linear differential equations
    2) A differential subfield k F K is a Picard-Vessiot extension of k if and ... monodromy group is a subset of the differential Galois group over the ...
  33. [33]
    [PDF] 8 Picard–Vessiot theory
    As in Galois theory, one can form the differential Galois group of an extension k ⊂ Kof differential fields as the group of automorphisms of the differential ...
  34. [34]
    AMS eBooks: Graduate Studies in Mathematics
    Regular singular points and the local Riemann-Hilbert correspondence; Chapter 10. Local Riemann-Hilbert correspondence as an equivalence of categories ...