Fact-checked by Grok 2 weeks ago

Uniqueness theorem

In mathematics, a uniqueness theorem asserts that a particular mathematical object satisfying given properties is the only one of its kind, or that all such objects are equivalent under those properties. Such theorems are fundamental across various fields, ensuring that solutions or structures derived from axioms or equations do not admit multiple distinct realizations. One of the most prominent applications of uniqueness theorems occurs in the theory of ordinary differential equations (ODEs), where it addresses the initial value problem (IVP) defined by y' = f(t, y) with y(t_0) = y_0. Specifically, the theorem states that if f(t, y) and its partial derivative \frac{\partial f}{\partial y}(t, y) are continuous on an open rectangle containing the point (t_0, y_0), then the IVP has a unique solution on some open interval around t_0. This condition, often involving Lipschitz continuity in y, prevents the existence of multiple solutions intersecting at the initial point, as integral curves cannot cross under these hypotheses. The theorem's proof typically relies on Picard-Lindelöf iteration or fixed-point arguments in Banach spaces, establishing both and locally. It is crucial for practical applications, such as validating general solutions to linear ODEs like x'' + 8x' + 7x = 0, where arbitrary constants are uniquely determined by initial conditions. However, uniqueness may fail if the continuity conditions are violated, as seen in examples like y' = 2\sqrt{|y|} with y(0) = 0, which admits infinitely many solutions. Beyond ODEs, uniqueness theorems appear in linear algebra, guaranteeing a unique solution to Ax = b when A is invertible, and in other areas like partial differential equations or dynamical systems, where they ensure solitary equilibrium points or boundary value solutions. These results underpin much of applied mathematics, from physics modeling to numerical simulations, by providing reliability in solution existence and computation.

Definition and Principles

Core Definition

A uniqueness theorem in asserts that a satisfying given properties or conditions is unique, meaning there is at most one such object, or that all such objects are equivalent up to representation by the same model. In practical terms, for problems such as solving an subject to or conditions, the theorem guarantees that if a exists, it is the only one fulfilling those conditions. This ensures predictability and in mathematical modeling, preventing ambiguity in outcomes. Formally, consider a well-posed P, with S denoting its . A uniqueness theorem states that if S is non-empty, then |S| \leq 1, implying the is singular when it exists. This general formulation applies across various domains, from algebraic systems to analytic problems, emphasizing that the conditions uniquely determine the object. The logical structure of uniqueness proofs often proceeds by or direct comparison: assume two s u and v exist, then demonstrate that u = v by showing their difference satisfies an with only the trivial . This contrapositive approach—negating the existence of multiple s—underpins many such theorems, relying on properties like or to establish coincidence. In some contexts, uniqueness is understood in an essential sense, where solutions are deemed the same if they differ merely by trivial transformations that preserve the underlying structure, such as reparameterization. For instance, in the study of periodic functions, two solutions with the same period and form are essentially unique if one is obtained from the other by a phase shift, as the shift does not alter the fundamental oscillatory behavior. This notion formalizes "sameness" via an , allowing flexibility in representation without compromising the theorem's intent.

Conditions for Uniqueness

The uniqueness theorem in requires specific conditions on the functions and domains involved to ensure that solutions to associated problems, such as or problems, are unique. A fundamental requirement is the of the defining function, often denoted as f(t, y), combined with the with respect to the dependent variable y. This means there exists a constant L > 0 such that |f(t, y_1) - f(t, y_2)| \leq L |y_1 - y_2| for all t in the and y_1, y_2 in a suitable . The Lipschitz condition prevents the function from varying too rapidly, which could otherwise allow multiple solutions to emerge. In addition to functional regularity, the problem setup must satisfy well-posedness criteria as defined by Hadamard, which encompass of a , , and continuous dependence on or data (). For value problems, appropriate conditions—such as y(t_0) = y_0—must be specified within a bounded where the remains continuous and bounded, ensuring the is well-defined. These criteria collectively guarantee that small perturbations in data do not lead to drastically different solutions, reinforcing the theorem's applicability. Proofs of uniqueness under these conditions frequently employ Gronwall's inequality to bound the difference between two potential solutions. Suppose u(t) and v(t) are solutions satisfying |u'(t) - v'(t)| \leq K |u(t) - v(t)| with u(t_0) = v(t_0); Gronwall's inequality then implies |u(t) - v(t)| \leq 0 \cdot e^{K|t - t_0|}, yielding u(t) = v(t). This technique leverages the exponential growth bound to show that any deviation remains zero, directly tying the Lipschitz constant K to the of the solution. As an illustrative example, consider uniqueness for integral equations of the form y(t) = g(t) + \int_a^t K(t, s, y(s)) \, ds in a . If the kernel operator induces a —meaning its norm is less than 1—then the ensures a unique fixed point, hence a unique solution. This approach extends the condition to operator settings, providing a robust framework for abstract problems.

Historical Context

Origins in Analysis

The foundations of uniqueness theorems in trace back to the , when pioneers like Leonhard Euler and developed methods for solving differential equations arising from physical and geometrical problems, often implicitly assuming that solutions were unique under given initial conditions without providing formal proofs. Euler, in works such as his 1768–1770 Institutionum calculi integralis, constructed series solutions and geometric interpretations for ordinary differential equations, treating uniqueness as a natural outcome of the analytical continuation of solutions, particularly in contexts like the where multiple solutions would disrupt predictive consistency. Similarly, Lagrange, in his 1770 contributions to the and implicit functions, advanced variational principles to derive , presupposing a single extremal path or solution for optimization problems in , as seen in his Théorie des fonctions analytiques (1797), where he emphasized algebraic resolution without rigorous uniqueness arguments. This implicit reliance on uniqueness gained explicit motivation in the early through Augustin-Louis Cauchy's rigorous reformulation of , where he introduced initial value problems for ordinary equations and first conjectured based on continuity assumptions. In his 1820–1830 Exercises d'analyse, Cauchy proved existence via integral equations and successive approximations, arguing that continuous dependence on initial ensured a unique solution in a neighborhood, marking the shift from solving to analytical foundations; he extended this to systems in 1842, using majorants to confirm analytic . A primary impetus for these conjectures stemmed from , where astronomers like Kepler had empirically observed unique planetary orbits governed by inverse-square laws, rendering multiple solutions incompatible with predictable motions—thus, modeling these via equations necessitated to align theory with astronomical , as Cauchy noted in applying his methods to orbital perturbations. A pivotal milestone occurred in the 1880s with Henri Poincaré's stability analysis of dynamical systems, which built on Cauchy's ideas to explore qualitative in nonlinear contexts, particularly for periodic orbits in . In his prize memoir Les Méthodes nouvelles de la mécanique céleste and subsequent 1892–1899 volumes, Poincaré introduced recurrence theorems and stability criteria for fixed points, demonstrating that under certain integrability conditions, solutions remained uniquely determined over long times despite perturbations, laying essential groundwork for modern uniqueness in non-integrable systems without relying solely on explicit solvability.

Key Developments in the 19th and 20th Centuries

In the late , foundational work on for ordinary differential equations (ODEs) emerged alongside efforts to establish . introduced a key condition in the , stipulating that the right-hand side function must be continuous with respect to the dependent variable to guarantee a unique solution to the . This condition provided a sufficient criterion for but required stronger regularity than mere . In contrast, Giuseppe Peano's 1886 theorem demonstrated the of solutions under only the assumption of of the right-hand side function, highlighting cases where multiple solutions could arise without the restriction. Early 20th-century refinements built on these foundations to link existence and uniqueness more tightly. Émile Picard's 1890 work refined Peano's approach through successive approximations, culminating in the , which asserts local existence and uniqueness for ODE initial value problems when the right-hand side is continuous and locally continuous in the dependent variable. Complementing this, William F. Osgood's 1898 criterion extended uniqueness guarantees to cases where the condition fails, replacing it with an integrability requirement on the —specifically, that the of the of the modulus diverges—to ensure a unique solution. Mid-century advancements addressed partial differential equations (PDEs). In 1901, Erik Ivar Fredholm Holmgren established a uniqueness theorem for linear PDEs with analytic coefficients, proving that if a vanishes on a non-characteristic analytic , it must vanish identically in a neighborhood, including in characteristic strips adjacent to the surface. This result strengthened earlier principles and applied broadly to hyperbolic and elliptic equations. Later developments in the focused on relaxing smoothness assumptions for ODEs. extended the framework to non-smooth right-hand sides by introducing conditions where the function is measurable in the independent variable, continuous in the dependent variable, and satisfies a local integrability bound, enabling existence theorems for generalized solutions while preserving under additional one-sided Lipschitz-like constraints. These extensions broadened applicability to irregular systems in and physics.

Mathematical Applications

Ordinary Differential Equations

In the context of ordinary differential equations (ODEs), the uniqueness theorem addresses the existence of a single solution to an of the form y' = f(t, y), y(t_0) = y_0, where f is a defined on a in \mathbb{R}^2. The provides a fundamental result guaranteeing both existence and uniqueness under suitable conditions. Specifically, if f is continuous in t and continuous in y on a |t - t_0| \leq a, |y - y_0| \leq b, with Lipschitz constant K, and if M = \max |f(t, y)| on this rectangle satisfies h = \min(a, b/M) > 0, then there exists a unique solution on the interval |t - t_0| \leq h. This theorem establishes local uniqueness, meaning the solution is unique in a neighborhood of the initial point t_0. For global uniqueness over larger intervals, the solution can be extended via theorems, which assert that under the same hypotheses, the solution can be prolonged to a maximal of ( \alpha, \beta ), where \alpha < t_0 < \beta, and uniqueness holds throughout this interval unless the solution approaches the boundary of the domain in finite time. A key insight into the role of the Lipschitz condition arises from examples where it fails, leading to non-uniqueness. Consider Peano's example: the initial value problem y' = 3 y^{2/3}, y(0) = 0. Here, f(t, y) = 3 y^{2/3} is continuous but not near y = 0, as the partial derivative \partial f / \partial y = 2 y^{-1/3} is unbounded. This equation admits infinitely many solutions, including the trivial solution y(t) = 0 for all t, and non-trivial solutions such as y(t) = 0 for t \leq 0 and y(t) = t^3 for t > 0, or similar piecewise constructions for arbitrary starting times. The proof of uniqueness in the often relies on analyzing the difference between two potential solutions \phi(t) and \psi(t), both satisfying the integral equation form y(t) = y_0 + \int_{t_0}^t f(s, y(s)) \, ds. Subtracting these yields |\phi(t) - \psi(t)| \leq K \int_{t_0}^t |\phi(s) - \psi(s)| \, ds, where K is the Lipschitz constant. Applying Gronwall's inequality to this differential inequality implies |\phi(t) - \psi(t)| \leq 0 for t in the interval, hence \phi(t) = \psi(t).

Partial Differential Equations

In partial differential equations (PDEs), uniqueness theorems establish that solutions to boundary or initial value problems are determined solely by the given data under appropriate conditions on the equation type, domain, and coefficients. For elliptic PDEs, such as \nabla^2 u = f in a bounded domain \Omega \subset \mathbb{R}^n, uniqueness holds for Dirichlet boundary conditions u = g on \partial \Omega, where g is continuous, ensuring a unique solution in C^2(\Omega) \cap C(\overline{\Omega}). This follows from the , which implies that the difference of two solutions is and attains its maximum on the , forcing it to vanish identically if zero there. For conditions \frac{\partial u}{\partial n} = h on \partial \Omega, where h satisfies the condition \int_\Omega f \, dV = \oint_{\partial \Omega} h \, dS, the solution to is unique up to an additive constant. This non-uniqueness arises because constant functions satisfy the homogeneous Neumann problem, and existence requires the condition to ensure solvability. The proof relies on energy methods: if u_1 and u_2 are solutions, their \phi = u_1 - u_2 satisfies \nabla^2 \phi = 0 with \frac{\partial \phi}{\partial n} = 0 on \partial \Omega; applying Green's first identity yields \int_\Omega |\nabla \phi|^2 \, dV = \oint_{\partial \Omega} \phi \frac{\partial \phi}{\partial n} \, dS = 0, implying \nabla \phi = 0 and thus \phi constant. A similar applies to the homogeneous for , where \int_\Omega |\nabla u|^2 \, dV = \oint_{\partial \Omega} u \frac{\partial u}{\partial n} \, dS = 0 if u = 0 on \partial \Omega, confirming by showing u = 0. For parabolic PDEs, such as the u_t - \Delta u = 0 in \Omega \times (0,T) with u(x,0) = g(x) and homogeneous Dirichlet boundaries, follows from energy estimates on the difference of solutions. In the , testing with the solution itself and using of the leads to \frac{1}{2} \frac{d}{dt} \|u\|^2_{L^2(\Omega)} + \beta \|u\|^2_{H^1_0(\Omega)} \leq \gamma \|u\|^2_{L^2(\Omega)}, and Gronwall's inequality with zero initial data implies u \equiv 0, establishing in appropriate Sobolev spaces. For hyperbolic PDEs, Holmgren's theorem provides uniqueness under analytic assumptions: for a linear hyperbolic PDE with analytic coefficients in a domain where Cauchy data vanish on an open non-characteristic subset of the initial , the solution vanishes identically in the domain of dependence. This result, originally for two variables and extended to higher dimensions, strengthens the Cauchy-Kovalevskaya theorem by applying to continuous solutions with continuous first derivatives, relying on to propagate the zero data.

Applications in Physics and Engineering

Electromagnetism

In , uniqueness theorems ensure that solutions to are uniquely determined by specified sources and boundary conditions, which is crucial for predicting electromagnetic fields in bounded domains. For the static case in , the governing equations are ∇·D = ρ and ∇×E = 0, where D is the , ρ is the , and E is the . Given the charge distribution ρ and appropriate boundary conditions on a closed surface enclosing the region, the solution for E (and thus the electric potential φ) is unique. This follows from the fact that if two solutions exist, their difference satisfies the homogeneous equations with the same boundaries, leading to a zero field everywhere by energy considerations or integral identities. The electrostatic potential φ satisfies Δφ = -ρ/ε, where ε is the , subject to Dirichlet boundary conditions (specified φ on the boundary). of this in a bounded is established using Green's first identity: for two solutions φ₁ and φ₂, ∫_V (φ₁ Δφ₂ - φ₂ Δφ₁) dV = ∫_S (φ₁ ∇φ₂ - φ₂ ∇φ₁) · dS, which simplifies to zero for the homogeneous case under Dirichlet conditions, implying φ₁ = φ₂. Similarly, for Neumann boundaries (specified normal derivative), holds up to an additive constant. For time-harmonic fields, assuming e^{jωt} dependence, Maxwell's equations in a bounded domain with no sources inside yield a unique solution when tangential E and normal B (or equivalent impedance boundaries) are specified on the surface. The uniqueness theorem states that if the fields satisfy the source-free time-harmonic equations ∇×E = jωB, ∇×H = -jωD + J (with J=0 inside), ∇·D=0, ∇·B=0, and boundary conditions, then E=0 and H=0 throughout the volume, provided the medium has some loss (Im(ε)>0 or Im(μ)>0) to prevent resonances. This uniqueness is proved using a form of the Poynting theorem for complex fields: integrating (E* · ∇×H - H · ∇×E*) over the volume gives ∫_V (jω |B|^2 + jω |D|^2 + E* · J - H* · M) dV = ∫_S (E* × H) · dS, where the surface vanishes for the homogeneous case with specified tangential fields, and the volume terms imply zero without sources, forcing the fields to zero. For lossless media, uniqueness requires additional conditions like excluding zero-frequency modes. These theorems underpin numerical methods like finite element analysis for electromagnetic simulations.

Boundary Value Problems

In engineering contexts beyond electromagnetism, uniqueness theorems for boundary value problems ensure reliable predictions in thermal and fluid systems, where multiple solutions could lead to erroneous designs or simulations. A prominent example is the heat equation, modeling transient diffusion processes like heat conduction in materials. The equation is \frac{\partial u}{\partial t} - \kappa \Delta u = 0 in a bounded domain \Omega \subset \mathbb{R}^n for t > 0, with initial condition u(\mathbf{x}, 0) = u_0(\mathbf{x}) and suitable boundary conditions (e.g., Dirichlet or Neumann) on \partial \Omega. Under these conditions, solutions are unique in appropriate function spaces, such as C^{2,1}(\overline{\Omega} \times [0, T]). This uniqueness is established via the maximum principle, which implies that any two solutions differ by a function attaining its extrema only on the initial or boundary data, hence must be zero; alternatively, energy methods integrate \int_\Omega (u_1 - u_2)^2 d\mathbf{x} over time, leveraging the equation's structure to show the difference decays to zero. In fluid mechanics, the Stokes equations approximate incompressible flow at low Reynolds numbers, relevant for viscous-dominated regimes like lubrication or slow sedimentation. The system reads -\mu \Delta \mathbf{v} + \nabla p = \mathbf{f}, \quad \nabla \cdot \mathbf{v} = 0 in \Omega, with no-slip boundary conditions \mathbf{v} = 0 on \partial \Omega, where \mathbf{v} is the velocity field, p the pressure, \mu > 0 the viscosity, and \mathbf{f} a body force. Uniqueness of the solution (\mathbf{v}, p) \in [H_0^1(\Omega)]^d \times L^2(\Omega)/\mathbb{R} (for dimension d) follows from the mixed variational formulation: find (\mathbf{v}, p) such that \int_\Omega \mu \nabla \mathbf{v} : \nabla \mathbf{w} \, d\mathbf{x} - \int_\Omega p \nabla \cdot \mathbf{w} \, d\mathbf{x} + \int_\Omega q \nabla \cdot \mathbf{v} \, d\mathbf{x} = \int_\Omega \mathbf{f} \cdot \mathbf{w} \, d\mathbf{x} for test functions \mathbf{w} \in [H_0^1(\Omega)]^d, q \in L^2(\Omega). The bilinear form is coercive on the divergence-free subspace, yielding uniqueness via the Lax-Milgram theorem. An illustrative case from acoustics is the for the , describing time-harmonic wave propagation in media like air or water. The equation is \Delta u + k^2 u = 0 in an exterior domain \Omega (e.g., around an obstacle), with boundary condition u = g on \partial \Omega and at infinity, where k > 0 is the . Uniqueness holds provided k^2 is not a Dirichlet eigenvalue of -\Delta on \Omega, avoiding resonances where spurious modes appear; at non-resonant frequencies, the solution is unique in Sobolev spaces H^1_{\rm loc}(\Omega). This ensures stable predictions for sound waves off structures. For steady-state , common in modeling constant or solute transport, the equation takes divergence form -\nabla \cdot (k \nabla u) = f in \Omega, where k > 0 is a (possibly variable), f a source term, with mixed boundary conditions including flux specifications like \mathbf{n} \cdot (k \nabla u) = h on parts of \partial \Omega. Assuming k is bounded and elliptic (i.e., \xi^T (k \xi) \geq \alpha |\xi|^2 for \alpha > 0), the u \in H^1(\Omega) is unique, proved by showing the associated \int_\Omega k \nabla u \cdot \nabla v \, d\mathbf{x} is coercive and continuous, invoking Lax-Milgram for and uniqueness. This form accommodates heterogeneous media, crucial for composites.

Importance and Limitations

Role in Proofs and Modeling

Uniqueness theorems play a crucial role in mathematical proofs by establishing that a solution, once shown to exist, is the only one satisfying the given conditions, thereby completing the analysis of a problem. For instance, in the context of fixed-point problems, the guarantees both the existence and of a fixed point for contractions in complete metric spaces, providing a powerful tool for proving the solvability of equations in and beyond. This uniqueness aspect ensures that proofs are not only about affirmation of existence but also about exclusivity, preventing multiple interpretations of the result. In scientific modeling, uniqueness theorems ensure that predictions from differential equations are determinate, avoiding ambiguity in simulating physical phenomena. For example, in , uniqueness of solutions to initial value problems for second-order ordinary equations modeling particle trajectories under Newtonian forces guarantees a single path for given initial conditions, which is essential for reliable forecasting in dynamics. Without , models could yield multiple possible outcomes, undermining their predictive power in engineering and physics applications. From a computational perspective, uniqueness theorems justify the reliability of numerical solvers for equations by assuring that iterative methods, such as those based on Picard iteration, converge to the true solution rather than to one of potentially many. This is particularly vital in solving and partial differential equations, where algorithms like Runge-Kutta methods rely on the underlying problem's well-posedness to validate their approximations. Uniqueness forms one of the three pillars of well-posedness, as defined by Hadamard, alongside existence and continuous dependence on initial data, ensuring that mathematical models are robust and suitable for practical use in analysis and simulation. This framework highlights how uniqueness contributes to the stability and interpretability of solutions in both theoretical and applied contexts.

Cases of Non-Uniqueness

In ordinary differential equations, uniqueness fails when the right-hand side function fails to satisfy the Lipschitz condition, allowing multiple solutions to emanate from the same initial condition. A classic counterexample is the initial value problem y' = \sqrt{|y|}, y(0) = 0, where the function f(y) = \sqrt{|y|} is continuous but not Lipschitz continuous at y = 0 since its derivative f'(y) = \frac{1}{2\sqrt{|y|}} blows up there. This yields at least two solutions: the trivial solution y(t) = 0 for all t, and the non-trivial solution y(t) = 0 for t \leq 0 and y(t) = \left( \frac{t}{2} \right)^2 for t \geq 0, both satisfying the equation and initial condition. More generally, infinitely many solutions exist by piecing together the trivial and non-trivial branches at arbitrary points, illustrating how the absence of the Lipschitz condition permits "funneling" of solutions backward from the initial point. Ill-posed problems in partial differential equations, such as the for , demonstrate non-, violating Hadamard's criteria for well-posedness (, , and continuous dependence on data). This problem seeks the solution u in the half-plane x > 0 given the boundary values u(0, y) and normal derivative \frac{\partial u}{\partial x}(0, y) on x = 0, governed by \Delta u = 0. As highlighted by Hadamard, for zero boundary data, there exist infinitely many non-trivial solutions, and the problem generally lacks unless the data are analytic. Small perturbations in the data can lead to non- or dramatically different solutions, as the elliptic nature allows rapid growth away from the boundary, rendering practical solutions non-unique without exact data. This instability, first highlighted by Hadamard, underscores the challenges in overdetermined boundary value problems for elliptic equations. In , uniqueness theorems often hold only modulo automorphisms, leading to multiple isomorphic models for the same object. For elliptic curves over an of characteristic not 2 or 3, two curves are isomorphic if and only if their j-invariants coincide, but the \mathrm{Aut}(E) acts non-trivially, typically as \mathbb{Z}/2\mathbb{Z} generated by the inversion map (x, y) \mapsto (x, -y), allowing distinct Weierstrass equations to represent the same curve up to . Special cases with larger groups occur for j=0 (order 6, hexagonal symmetry) or j=1728 (order 4, square symmetry), where the exhibits non-uniqueness beyond isomorphism classes, as automorphisms permute points and complicate the identification of canonical forms. This topological ambiguity arises from the rigid analytic structure of elliptic curves as genus-1 Riemann surfaces with a specified , where the full solution set includes orbits under the automorphism action. Euler's equations for the free rotation of a with zero exhibit a of , highlighting non-uniqueness in the dynamical . The equations in principal axes are I_1 \dot{\omega}_1 = (I_2 - I_3) \omega_2 \omega_3, and cyclic permutations, with no external torques. For zero \mathbf{L} = \mathbf{I} \boldsymbol{\omega} = \mathbf{0}, this implies \boldsymbol{\omega} = \mathbf{0}, satisfying the equations as an regardless of the body's initial in space. Thus, any constant with vanishing constitutes a , forming a parameterized by the SO(3) space, where the permit arbitrarily many static without specifying uniquely.

References

  1. [1]
    Uniqueness Theorem -- from Wolfram MathWorld
    A uniqueness theorem states that a mathematical object is unique, meaning only one object fulfills given properties, or all objects are equivalent.
  2. [2]
  3. [3]
    [PDF] Proof of existence and uniqueness theorem - MIT OpenCourseWare
    The existence and uniqueness theorem for differential equations is a key technical result. For example, when we solve an equation like 𝑥″ + 8𝑥′ + 7𝑥 = 0, we ...
  4. [4]
    [PDF] The Existence and Uniqueness Theorem for ODE's - Rose-Hulman
    Proof of Existence-Uniqueness Theorem: First recast the differen- tial equation as an integral equation. Note that if y(t) is continuous for t0 ≤ t ≤ t0 + ...
  5. [5]
    The Definitive Glossary of Higher Mathematical Jargon - Math Vault
    In general, an assertion of essential uniqueness presupposes some definition of “sameness”, which is often formalized using some equivalence relation.
  6. [6]
    [PDF] I. An existence and uniqueness theorem for differential equations
    If in Picard's theorem one drops the Lipschitz condition then there may be more than one solution, thus the uniqueness assertion in the theorem is not longer ...
  7. [7]
    [PDF] Lecture 1. Introduction to well- and ill-posed problems.
    Nov 19, 2009 · Clearly, the problem (1.1) is well-posed in the sense of Hadamard if and only if there ... Solution of the integral equation of the first kind.
  8. [8]
    [PDF] The Gronwall inequality - University of South Carolina
    1. Introduction. The Gronwall inequality as given here estimates the difference of solutions to two differential equations y. 0.
  9. [9]
    [PDF] Chapter 3: The Contraction Mapping Theorem
    The contraction mapping theorem states that a strict contraction on a complete metric space has a unique fixed point. The contraction mapping theorem is ...
  10. [10]
    [PDF] The History of Differential Equations, 1670–1950
    Most 18th-century developments consolidated the Leibnizian tradition, extend- ing its multi-variate form, thus leading to partial differential equations.
  11. [11]
    [PDF] THE EXISTENCE THEOREMS IN ORDINARY DIFFERENTIAL ...
    A third method for establishing the existence of solutions of ordinary differential equations, probably known to Cauchy, was first published by J ...
  12. [12]
    History of dynamical systems - Scholarpedia
    Oct 21, 2011 · This article provides a brief, and perhaps idiosyncratic, introductory review of the early history of the subject, from approximately 1885 through 1965.Missing: uniqueness 1880s
  13. [13]
    Rudolf Lipschitz (1832 - 1903) - Biography - University of St Andrews
    Rudolf Lipschitz is remembered for the "Lipschitz condition", an inequality that guarantees a unique solution to the differential equation y' = f (x, y).Missing: 1870s | Show results with:1870s
  14. [14]
    Peano's 1886 existence theorem on first-order scalar differential ...
    Feb 11, 2016 · In 1886 Giuseppe Peano presents the first proof of the existence of a solution of an initial value problem y ′ = f ( x , y ) , y ( a ) = b ...Missing: paper | Show results with:paper
  15. [15]
    MATHEMATICA TUTORIAL: Existence
    Emile Picard. The theorem above is usually referred to as Picard's theorem (or sometimes Picard--Lindelöf theorem or the method of successive approximations) ...
  16. [16]
    [PDF] The Osgood Criterion and Finite-Time Cosmological Singularities
    Apr 27, 2016 · The Osgood criterion [1] is a classical criterion, due to W.F. Osgood in 1898 ... for ordinary differential equations to admit unique solutions ...
  17. [17]
    [PDF] Remarks on Holmgren's uniqueness theorem - Numdam
    Holmgren's uniqueness theorem states that a solution of a linear differential equation with (real) analytic coefficients must vanish in a full.
  18. [18]
    (PDF) On Carathéodory's conditions for the initial value problem
    Aug 6, 2025 · Existence and uniqueness theorems under classical and Carathéodory conditions are discussed extensively in Coddington and Levinson [32]. See ...
  19. [19]
    [PDF] Picard's Existence and Uniqueness Theorem
    One of the most important theorems in Ordinary Differential Equations is Picard's. Existence and Uniqueness Theorem for first-order ordinary differential ...Missing: 1890 | Show results with:1890
  20. [20]
    [PDF] Ordinary Differential Equations
    Theorem (Gronwall's Inequality - differential form). Let I = [t0,t1]. Suppose a : I →. R and b : I → R are continuous, and suppose u : I → R is in C1(I) ...
  21. [21]
    [PDF] Peano's Existence Theorem revisited - arXiv
    Feb 6, 2012 · We owe the following example precisely to Peano. Peano's example of a problem with infinitely many solutions. The scalar problem y′ = 3y2/3, y(0) ...
  22. [22]
    [PDF] 2.3 The Existence and Uniqueness Theorem.
    The second uniqueness proof is a classic method of proving uniqueness. The differential inequality is a Grönwall's Inequality. Here is a slightly more general ...
  23. [23]
    [PDF] Chapter 2: Laplace's equation - UC Davis Mathematics
    The maximum principle gives a uniqueness result for the Dirichlet problem for the Poisson equation. Theorem 2.18. Suppose that Ω is a bounded, connected ...
  24. [24]
    [PDF] Uniqueness of solutions to the Laplace and Poisson equations
    For the case of Dirichlet boundary conditions or mixed boundary conditions, the solution to Poisson's equation always exists and is unique. Finally, for the ...Missing: maximum | Show results with:maximum
  25. [25]
    Uniqueness of Solution to Poisson's Equation - Jean-Sébastien Caux
    Feb 27, 2024 · For Neumann boundary conditions, the solution is unique apart from an unimportant additive constant. We can thus finally state the. Uniqueness ...
  26. [26]
    [PDF] Chapter 6: Parabolic equations - UC Davis Mathematics
    The theory of parabolic PDEs closely follows that of elliptic PDEs and, like elliptic PDEs, parabolic PDEs have strong smoothing properties. For example,.
  27. [27]
    Remarks on Holmgren's uniqueness theorem
    Förh., 58 (1901), 91-103. | JFM. [6] L. Hörmander, Uniqueness ... Hörmander, A uniqueness theorem for second order hyperbolic differential equations, Comm.Missing: original | Show results with:original
  28. [28]
    [PDF] Section 2: Electrostatics
    Now let us show the uniqueness of the solution of Poisson's equation, 2. 0. /ρ ε. ∇ Φ = -. , inside a volume V subject to either Dirichlet or Neumann boundary ...
  29. [29]
    [PDF] Green's Identities, Uniqueness, Dirichlet and Neumann Green's ...
    Thus φ1 = φ2 and the solution is unique. For Neumann boundary conditions, φ1 and φ2 can only differ by only an arbitrary constant. Since E = -vφ, the electric ...
  30. [30]
    [PDF] Uniqueness Theorem
    Uniqueness Theorem. Page 1. Uniqueness Theorem. Consider the symmetric, time-harmonic form of Maxwell's equations given by v × E = −Mt. (1a). Mt = jωB + Mi. (1b).
  31. [31]
    [PDF] discussion of the heat equation - UChicago Math
    The uniqueness is proved in two ways- energy method and maximum principle. The former gives physical interpretation of the heat equation while the latter has ...
  32. [32]
    [PDF] The Stokes Equations
    Choosing v = w shows that the supremum is not smaller than 1. □. Theorem 3.5. Existence and uniqueness of a solution of the Stokes equations. Let Ω be a ...
  33. [33]
    [PDF] Scattering of time-harmonic acoustic waves: Helmholtz equation ...
    Mar 1, 2021 · We describe some boundary value problems (BVPs) and focus on one of them, the exterior Dirichlet problem. We show how to reformulate this as a ...
  34. [34]
    [PDF] Numerical Methods for Elliptic Partial Differential Equations
    elliptic equation of the form. −div (k ∇u) = f. Sometimes, instead of div W = f and W = −k ∇u we get the same elliptic equation from a different starting ...
  35. [35]
    [PDF] banach's fixed point theorem and applications
    It states conditions sufficient for the existence and uniqueness of a fixed point, which we will see is a point that is mapped to itself.
  36. [36]
    The Banach Fixed Point Theorem: selected topics from its hundred ...
    Jul 9, 2024 · The Banach theorem is simple in its formulation, the fixed point is always unique and it is obtained by an explicit calculation. Its ...
  37. [37]
    [PDF] numerical methods for solving ordinary differential equations
    ○ Basic theory for ODE problems: well-posedness: existence, uniqueness, stability with respect to small perturbations;. ○ Basic concepts for ODE solvers ...<|control11|><|separator|>
  38. [38]
    [PDF] Well-Posed Problems - UNL Math
    According to Hadamard, a problem is well-posed (or correctly-set) if a. it has a solution, b. the solution is unique, c. the solution depends continuously ...
  39. [39]
    [PDF] Ordinary Differential Equations - Michigan State University
    Apr 1, 2015 · The techniques were developed in the eighteenth and nineteenth centuries and the equations include linear equations, separable equations ...
  40. [40]
    [PDF] PRINCETON UNIVERSITY BULLETIN.
    XIII. SUR LES PROBLÈMES AUX DERIVEES. PARTIELLES ET LEUR SIGNIFICA. TION PHYSIQUE. PAR M. JACQUES HADAMARD. “ La physique ne nous donne pas seulement l'occa.
  41. [41]
    [PDF] Joseph H. Silverman - The Arithmetic of Elliptic Curves
    The past two decades have witnessed tremendous progress in the study of elliptic curves. Among the many highlights are the proof by Merel [170] of uniform bound ...
  42. [42]
    [PDF] Euler's Equations - 3D Rigid Body Dynamics - MIT OpenCourseWare
    We now turn to the task of deriving the general equations of motion for a three-dimensional rigid body. These equations are referred to as Euler's equations ...Missing: equilibria uniqueness