Fact-checked by Grok 2 weeks ago

Fixed-point iteration

Fixed-point iteration is a fundamental method in for computing fixed points of a g, defined as points p where g(p) = p. This technique is widely used to approximate solutions to nonlinear equations of the form f(x) = 0 by reformulating them into an equivalent fixed-point problem, such as g(x) = x - f(x), where the roots of f correspond to the fixed points of g. The algorithm begins with an initial approximation x_0 and generates a sequence of iterates via x_{n+1} = g(x_n) until the values stabilize within a specified or a maximum number of iterations is reached. This iterative process relies on the function g being well-behaved, and practical implementations often include stopping criteria based on the difference between consecutive iterates, such as |x_{n+1} - x_n| < \epsilon. Convergence of the method to a unique fixed point is ensured if g is continuous on a closed interval [a, b], maps [a, b] into itself, and satisfies |g'(x)| \leq k < 1 for some constant k and all x \in [a, b]. Under these conditions, the error decreases linearly, with bounds like |x_n - p| \leq k^n |x_0 - p|, making the method reliable for sufficiently contractive mappings. The choice of g is critical, as different rearrangements of the original equation can affect the convergence rate and stability.

Fundamentals

Definition

Fixed-point iteration is a fundamental numerical method in applied mathematics used to approximate solutions to equations by identifying fixed points of a suitably chosen function. Given a function g: D \to D, where D is a subset of a metric space (X, d), a fixed point x^* \in D satisfies g(x^*) = x^*. Here, the metric space (X, d) consists of a set X with a distance function d: X \times X \to [0, \infty) that obeys the properties of non-negativity (with d(x, y) = 0 if and only if x = y), symmetry (d(x, y) = d(y, x)), and the triangle inequality (d(x, z) \leq d(x, y) + d(y, z)). This setup generalizes the concept beyond real numbers, allowing application in abstract spaces while assuming familiarity with basic functions and sequences. The iteration process starts with an initial guess x_0 \in D and constructs a sequence \{x_n\}_{n=0}^\infty via the recursive formula x_{n+1} = g(x_n), \quad n = 0, 1, 2, \dots The objective is for this sequence to converge to a fixed point x^*, providing an approximation to the solution of the original problem. Fixed points often arise from reformulating equations; for instance, to solve f(x) = 0, one may define g(x) = x - f(x), so that roots of f correspond to fixed points of g. The origins of fixed-point iteration trace to the method of successive approximations, first introduced by Joseph Liouville in 1837 and systematically advanced by Émile Picard around 1890, particularly for establishing existence of solutions to differential equations through iterative schemes. As a dedicated numerical technique for root-finding, it emerged prominently in the early 20th century, bolstered by theoretical advancements such as L. E. J. Brouwer's 1912 theorem guaranteeing the existence of fixed points for continuous self-maps on compact convex subsets of Euclidean space.

Basic Examples

To illustrate fixed-point iteration, consider the general process of generating a sequence where each term is obtained by applying a function g to the previous one, such that the limit satisfies the desired fixed-point equation if it converges. A classic introductory example is the Babylonian method for computing square roots, which reformulates the equation x^2 = 2 (or more generally x^2 = s) as the fixed-point problem x = g(x) with g(x) = \frac{1}{2} \left( x + \frac{2}{x} \right). This method, known since ancient Babylonian times around 1500 BC, provides rapid approximations through iteration. Starting with an initial guess x_0 = 1, x_1 = \frac{1}{2} \left( 1 + \frac{2}{1} \right) = 1.5, x_2 = \frac{1}{2} \left( 1.5 + \frac{2}{1.5} \right) \approx 1.4167, x_3 = \frac{1}{2} \left( 1.4167 + \frac{2}{1.4167} \right) \approx 1.4142. The sequence approaches \sqrt{2} \approx 1.414213562, demonstrating how the iterates refine the solution. For a linear equation ax + b = 0, the fixed-point form can be constructed by selecting a suitable g(x) to promote convergence, such as through relaxation techniques. Consider the equation $5x - 10 = 0, which has solution x = 2. One reformulation is x = g(x) with g(x) = 0.5x + 1, derived from x_{n+1} = x_n - \omega (5x_n - 10) using relaxation parameter \omega = 0.1. Starting from x_0 = 0, x_1 = 0.5 \cdot 0 + 1 = 1, x_2 = 0.5 \cdot 1 + 1 = 1.5, x_3 = 0.5 \cdot 1.5 + 1 = 1.75, x_4 = 0.5 \cdot 1.75 + 1 = 1.875. The terms steadily approach 2, illustrating the iterative refinement for linear problems. The choice of g significantly affects the iteration for a given equation. For instance, the transcendental equation x = \cos x (in radians) can be solved directly via x_{n+1} = \cos x_n, starting from x_0 = 0, yielding iterates that converge to the fixed point known as the Dottie number, approximately 0.739085. Alternative g functions for the same equation might involve trigonometric identities or approximations, potentially altering the sequence's path. To build intuition for these iterations, plotting the sequence values against n reveals the approach to the fixed point, while a basic cobweb diagram—constructed by drawing vertical lines from points on y = g(x) to y = x and horizontal lines back to y = g(x)—visually traces the trajectory of the iterates.

Convergence Analysis

Attracting and Repelling Fixed Points

In fixed-point iteration, fixed points of a function g are classified as attracting, repelling, or neutral based on the local behavior of iterates near the fixed point x^*, where g(x^*) = x^*, assuming g is differentiable at x^*. This classification relies on the magnitude of the derivative g'(x^*) to determine stability. An attracting fixed point occurs when |g'(x^*)| < 1. In this case, iterates starting sufficiently close to x^* converge to it, as small perturbations diminish over iterations. To derive this condition, consider the error e_n = x_n - x^* in the iteration x_{n+1} = g(x_n). Using the first-order expansion around x^*, g(x_n) = g(x^*) + g'(x^*)(x_n - x^*) + o(|x_n - x^*|), which simplifies to x_{n+1} - x^* \approx g'(x^*)(x_n - x^*), or e_{n+1} \approx g'(x^*) e_n. Iterating this yields e_{n} \approx [g'(x^*)]^n e_0, a geometric series that converges to zero as n \to \infty if |g'(x^*)| < 1. Conversely, a repelling fixed point arises when |g'(x^*)| > 1. Here, the same Taylor approximation shows that errors amplify, |e_{n+1}| \approx |g'(x^*)| |e_n|, causing iterates to diverge from x^* unless starting exactly at it. This instability makes repelling fixed points unsuitable for reliable in fixed-point iteration. When |g'(x^*)| = 1, the fixed point is , and the behavior is more subtle, depending on higher-order terms in the Taylor or the specific form of g. For instance, it may be weakly attracting from one side and repelling from the other, or exhibit slower without clear attraction or repulsion. Analysis requires examining the sign of g'(x^*) and subsequent derivatives to resolve these cases. For attracting fixed points, the convergence is linear, with asymptotic error constant \mu = |g'(x^*)|, meaning the relative error reduces by a factor of \mu < 1 per iteration. The number of accurate decimal digits gained per step is approximately -\log_{10} \mu, providing a practical measure of efficiency. Smaller \mu yields faster convergence, though values too close to 1 result in sluggish progress.

Banach Fixed-Point Theorem

The Banach fixed-point theorem, also known as the contraction mapping theorem, states that if (X, d) is a complete metric space and g: X \to X is a contraction mapping—meaning there exists a constant k with $0 \leq k < 1 such that d(g(x), g(y)) \leq k \, d(x, y) for all x, y \in X—then g has a unique fixed point x^* \in X (i.e., g(x^*) = x^*), and the sequence defined by x_{n+1} = g(x_n) for any initial x_0 \in X converges to x^* as n \to \infty. This result provides a powerful guarantee of global convergence for fixed-point iterations under the contraction condition, distinguishing it from local convergence analyses. The theorem was formulated by Stefan Banach in his 1922 paper "Sur les opérations dans les ensembles abstraits et leur application aux équations intégrales," published in Fundamenta Mathematicae, building on earlier ideas of contractions in the context of integral equations by and others. Banach's work, stemming from his 1920 doctoral dissertation, extended these concepts to abstract complete metric spaces, providing both existence and uniqueness with a constructive iterative proof, a significant advancement over prior existential fixed-point results like those of (1886) and (1909–1912). To outline the proof, consider the sequence \{x_n\} generated by x_{n+1} = g(x_n) starting from arbitrary x_0 \in X. The contraction property implies d(x_{n+1}, x_n) \leq k^n d(x_1, x_0) for all n \geq 1, so the partial sums \sum_{i=m}^{n-1} d(x_{i+1}, x_i) \leq d(x_m, x_n) \leq k^m d(x_1, x_0)/(1-k) for n > m, showing \{x_n\} is Cauchy; by of X, it converges to some x^* \in X. of g (as a contraction) yields g(x^*) = x^*. For uniqueness, suppose y^* \neq x^* is another fixed point; then d(x^*, y^*) = d(g(x^*), g(y^*)) \leq k \, d(x^*, y^*), implying d(x^*, y^*) (1 - k) \leq 0, so x^* = y^* since k < 1. A key application is the Picard–Lindelöf theorem on existence and uniqueness of solutions to initial value problems for ordinary differential equations. For y' = f(t, y) with y(t_0) = y_0, where f is continuous and Lipschitz in y on a suitable rectangle, the integral operator T(y)(t) = y_0 + \int_{t_0}^t f(s, y(s)) \, ds on a closed ball in the space of continuous functions is a contraction for small time intervals, yielding a unique fixed point as the solution by the Banach theorem.

General Convergence Conditions

A sufficient condition for the convergence of fixed-point iteration x_{n+1} = g(x_n) to a fixed point x^* is that g is differentiable and |g'(x)| \leq k < 1 for all x in a closed interval [a, b] containing x^* and the initial guess x_0. This follows from the mean value theorem, which implies that the error satisfies |x_{n+1} - x^*| \leq k |x_n - x^*|, ensuring linear convergence with rate k from any starting point in the interval. For local convergence near x^*, it suffices that |g'(x)| \leq k < 1 in a small neighborhood around x^*, provided x_0 lies within this region. If g is monotonically increasing or decreasing on the interval and maps it into itself with |g'(x)| < 1, the iterates converge monotonically to x^* without overshooting. For example, for g(x) = (x^3 + 2)/7 on [0, 1], which is increasing and satisfies the derivative condition, the sequence approaches the fixed point from below. Convergence fails if |g'(x^*)| > 1, leading to as errors amplify. For instance, with g(x) = 3 + 2 \sin x, the fixed point is approximately x^* \approx 3.094, but |g'(x^*)| \approx 1.998 > 1, causing the iterates to diverge from any initial guess. If g'(x^*) < -1, the sequence may oscillate with increasing amplitude. Even when |g'(x^*)| = 1, convergence is not guaranteed and often fails due to insufficient error reduction. Higher-order convergence occurs if g'(x^*) = 0, resulting in at least rates where the e_{n+1} \approx C e_n^2 for some constant C > 0. This is derived from the expansion around x^*, where the linear term vanishes and the quadratic term dominates. To verify these conditions practically, one can analytically compute or bound |g'(x)| over the relevant interval to check if it is less than 1. Numerically, iterate the method from test initial points and monitor the or |x_{n+1} - x_n| for stabilization, while plotting g'(x) to assess the behavior near suspected fixed points.

Iterative Methods

Construction of Iterative Methods

Fixed-point iteration is a fundamental technique for solving nonlinear equations of the form f(x) = 0, where the goal is to find a x^* such that f(x^*) = 0. The general strategy involves reformulating the equation as x = g(x), where g is a suitably chosen whose fixed point coincides with the root of f. This rearrangement transforms the root-finding problem into an iterative process defined by x_{n+1} = g(x_n), starting from an initial guess x_0 sufficiently close to x^*. A critical aspect of constructing an effective g is selecting a form that promotes . For local convergence near the fixed point x^*, the must satisfy |g'(x)| < 1 throughout a neighborhood of x^*, ensuring that each iteration reduces the error. Additionally, the choice of g should avoid inducing periodic cycles, where the sequence oscillates between points without approaching x^*, which can occur if |g'(x)| \geq 1 at certain points or if the mapping is not contractive. For polynomial equations, rearrangement techniques focus on isolating the variable x in a way that emphasizes dominant terms, such as the highest-degree or linear components, to define g(x). For instance, given a polynomial f(x) = a_k x^k + \cdots + a_1 x + a_0 = 0, one may solve for x by dividing through by the leading coefficient and expressing lower-order terms as a function of the higher ones, or vice versa, while verifying the contraction condition on g'. This approach balances computational simplicity with the need for |g'(x)| < 1 near the anticipated root. In the multivariable case, fixed-point iteration extends naturally to systems \mathbf{F}(\mathbf{x}) = \mathbf{0} where \mathbf{x} \in \mathbb{R}^n. The construction rewrites the system as \mathbf{x} = \mathbf{g}(\mathbf{x}), with \mathbf{g}: \mathbb{R}^n \to \mathbb{R}^n, and proceeds by vector iteration \mathbf{x}_{n+1} = \mathbf{g}(\mathbf{x}_n). Convergence requires that the spectral radius of the D\mathbf{g}(\mathbf{x}) be less than 1 in a neighborhood of the fixed point, often checked using a suitable vector norm. A common pitfall in constructing g is selecting a rearrangement where |g'(x^*)| > 1, leading to divergence as errors amplify with each step. For example, consider the quadratic equation f(x) = x^2 - 4x + 3.5 = 0, with roots approximately 1.293 and 2.707. Rearranging as x = g(x) = x - f(x) = -x^2 + 5x - 3.5 yields fixed points at these roots, but |g'(1.293)| \approx 2.414 > 1, so iterations starting near 1.293 diverge, while those near 2.707 converge since |g'(2.707)| \approx 0.414 < 1. This illustrates how the same equation can yield both attracting and repelling fixed points depending on the local derivative behavior.

Specific Iterative Method Examples

One prominent example of a fixed-point iteration for root-finding is the , which reformulates the root-finding problem f(x) = 0 as finding a fixed point of the function g(x) = x - \frac{f(x)}{f'(x)}. This iteration generates a sequence x_{k+1} = g(x_k), and under the condition that f'(x^*) \neq 0 at the root x^*, it exhibits quadratic convergence, meaning the error typically squares with each iteration once sufficiently close to the root. Another derivative-free approach is the secant method, which approximates the derivative using finite differences and can be viewed as a fixed-point iteration g(x_n, x_{n-1}) = x_n - f(x_n) \frac{x_n - x_{n-1}}{f(x_n) - f(x_{n-1})}, requiring two initial guesses and updating with previous iterates. This method achieves superlinear convergence of order approximately 1.618 (the golden ratio), offering efficiency without explicit derivative computations, though it may require more function evaluations per iteration than Newton-Raphson. For solving systems of linear equations Ax = b, where A = L + D + U with D diagonal, L strictly lower triangular, and U strictly upper triangular, the defines the iteration x^{(k+1)} = D^{-1} (b - (L + U) x^{(k)}), treating it as a fixed-point problem with the iteration matrix -D^{-1}(L + U). The modifies this by incorporating newly computed components immediately, yielding x^{(k+1)} = D^{-1} (b - L x^{(k+1)} - U x^{(k)}), which often converges faster for diagonally dominant matrices due to reduced information lag. Iteration counts for these methods vary by problem conditioning; for instance, Newton-Raphson typically requires 5-10 iterations for machine precision on well-behaved nonlinear functions, while secant may need 8-15 due to its slightly slower order. Error estimates follow from convergence theory: for Newton-Raphson, the asymptotic error constant involves \frac{|f''(x^*)|}{2|f'(x^*)|}, enabling quadratic reduction; secant errors scale with a constant related to the second derivative. For Jacobi and Gauss-Seidel, convergence occurs if the spectral radius of the iteration matrix is less than 1, with error norms decreasing geometrically at rate equal to that radius, often yielding 20-50 iterations for sparse, diagonally dominant systems.
MethodProsCons
Newton-RaphsonQuadratic convergence; efficient near rootRequires derivative computation; sensitive to poor initial guesses
SecantNo derivatives needed; superlinear orderNeeds two initial points; irregular convergence behavior
JacobiSimple parallelizable implementationSlower convergence; requires storing previous iterate fully
Gauss-SeidelFaster than Jacobi for many matricesSequential updates limit parallelism; may diverge if not diagonally dominant

Convergence Acceleration

Fixed-point iterations often exhibit linear convergence with a rate determined by the derivative of the iteration function at the fixed point, leading to slow progress when this rate is close to 1 in magnitude. Convergence acceleration techniques modify the iteration to achieve faster convergence, typically quadratic or superlinear, without requiring higher-order derivatives. These methods are particularly valuable for practical computations where the base iteration converges too slowly, enhancing efficiency in applications like solving nonlinear equations or optimizing systems. One classical approach is Aitken's δ²-process, an extrapolation method designed to accelerate linearly convergent sequences generated by fixed-point iterations. For a sequence {x_n} from x_{n+1} = g(x_n), the accelerated estimate is given by x^* \approx x_n - \frac{(x_{n+1} - x_n)^2}{x_{n+2} - 2x_{n+1} + x_n}, which assumes the error follows a geometric progression and eliminates the linear term. This process, introduced by Aitken in 1926, transforms linear convergence into quadratic under suitable conditions, such as when the iteration function yields asymptotically linear errors. Steffensen's method extends this idea to a self-accelerating fixed-point iteration, avoiding the need for separate sequence computations by incorporating Aitken-like extrapolation within the iteration step. Starting with x_n, it computes y = g(x_n) and z = g(y), then sets x_{n+1} = y - \frac{(y - x_n)^2}{z - 2y + x_n}. This variant achieves quadratic convergence for fixed points where the first derivative of g is nonzero, making it derivative-free and comparable in efficiency to for scalar problems. The method was originally proposed by Steffensen in 1933 as a modification for iterative root-finding, but it directly applies to fixed-point formulations. Relaxation methods introduce a weighting parameter ω to blend the current iterate with the next fixed-point approximation, yielding x_{n+1} = (1 - ω) x_n + ω g(x_n). The optimal ω minimizes the spectral radius of the iteration operator, often chosen such that |1 - ω + ω g'(x^)| < |g'(x^)|, accelerating convergence for rates near 1. For linear problems, this corresponds to successive over-relaxation (SOR), where 1 < ω < 2 can significantly speed up ; in nonlinear settings, it stabilizes and hastens fixed-point schemes. This technique, generalized from early work on elliptic PDE solvers, is widely used when the base method is contractive but sluggish. For nonlinear systems, Anderson acceleration minimizes a residual norm over a limited history of previous iterates, combining them linearly to form the next approximation. Specifically, it solves for coefficients γ_k that minimize ||g(x_n) - x_n - \sum γ_k (g(x_{n-k}) - x_{n-k})||, then updates x_{n+1} = (1 - \sum γ_k) g(x_n) + \sum γ_k g(x_{n-k}). This root-finding perspective on fixed-point residuals promotes superlinear convergence, even for non-quadratic bases, and is effective for high-dimensional problems like electronic structure calculations. Introduced by Anderson in 1965 for , it has gained renewed interest for its robustness in optimization and physics simulations. These acceleration techniques are most beneficial when the fixed-point iteration converges linearly with |g'(x^*)| close to but less than 1, as in nearly neutral fixed points, where standard iterations require many steps for precision. Selection depends on dimensionality and available history: Aitken or Steffensen for scalars, relaxation for simple adjustments, and Anderson for vector systems.

Applications

Chaos Game

The chaos game is an algorithm that generates fractals as attractors of contractive iterated function systems (IFS) by applying fixed-point iteration in a probabilistic manner. It begins with an arbitrary starting point in the plane and iteratively transforms it using randomly selected affine contraction mappings from the IFS, resulting in a sequence of points that densely fills the unique invariant set, or attractor, of the system. This approach leverages the contractive property of the mappings, ensuring convergence to the fixed point of the Hutchinson operator associated with the IFS, as guaranteed by the Banach fixed-point theorem for complete metric spaces. A classic example is the Sierpinski gasket, generated by an IFS consisting of three similarity transformations, each scaling by a factor of 1/2 and mapping to one vertex of an equilateral triangle. The algorithm proceeds as follows: select an initial point \mathbf{x}_0; at each step n, choose an index i randomly according to given probabilities (often uniform), and set \mathbf{x}_{n+1} = M_i(\mathbf{x}_n), where each M_i is an affine contraction of the form M_i(\mathbf{x}) = A_i \mathbf{x} + \mathbf{b}_i with \|A_i\| < 1. Plotting the points \mathbf{x}_n after a sufficient number of iterations reveals the gasket, a self-similar fractal with Hausdorff dimension \log 3 / \log 2 \approx 1.585. Despite the random selection introducing apparent chaos in the point trajectories, the sequence converges almost surely to a dense approximation of the attractor because the contractions pull points toward the invariant set regardless of the initial condition or sequence of choices. The Barnsley fern, another prominent example, uses four such mappings with probabilities approximately 0.01, 0.85, 0.07, and 0.07 to produce a visually realistic fern leaf, modeling natural shapes through self-similarity. The dragon curve can similarly be approximated via an IFS with two rotations and scalings, yielding a space-filling curve with intricate folding patterns. Popularized by Michael Barnsley in the 1980s, the chaos game has applications in computer graphics for efficiently rendering complex fractals without explicit geometric construction, particularly for modeling organic forms like plants and coastlines. In practice, to avoid transient effects from the initial point, the first several thousand iterates are typically discarded before plotting, ensuring the output faithfully represents the attractor.

Attractors in Dynamical Systems

In the context of discrete dynamical systems generated by fixed-point iterations x_{n+1} = f(x_n), an attractor is a closed set A in the phase space such that a neighborhood of A is mapped into itself under iterations of f, and the orbits of points starting sufficiently close to A converge to A as n \to \infty. This generalizes the concept of attracting fixed points, where A consists of a single point x^* satisfying f(x^*) = x^* and nearby orbits approach x^*; more broadly, attractors can include periodic orbits, which are finite cycles \{p_1, p_2, \dots, p_k\} where f(p_i) = p_{i+1} and f(p_k) = p_1, with iterations converging to the cycle for initial conditions in a suitable basin. The basin of attraction of an A, denoted B(A), is the open set of all initial points whose orbits under f converge to A. In fixed-point iteration, the structure of basins depends on the map f; for contractive maps, the entire space may serve as the basin of a unique fixed point, but non-contractive maps can exhibit multiple with intertwined basins separated by repelling fixed points or orbits. Lyapunov stability provides a foundational criterion for : a fixed point x^* is Lyapunov stable if for every \epsilon > 0, there exists \delta > 0 such that if \|x_0 - x^*\| < \delta, then \|f^n(x_0) - x^*\| < \epsilon for all n \geq 0; it is asymptotically stable (hence an ) if additionally orbits converge to x^*, which occurs in one dimension when |f'(x^*)| < 1. A classic example illustrating the progression from fixed-point attractors to more complex sets arises in the f(x) = r x (1 - x) for x \in [0,1] and parameter r \in [0,4], studied via iterative dynamics. For $1 < r < 3, the attracting fixed point x^* = (r-1)/r draws orbits from the B(x^*) = (0,1), with governed by |f'(x^*)| = |2 - r| < 1. As r increases beyond 3, period-doubling bifurcations create attracting periodic orbits (e.g., a period-2 for $3 < r < 1 + \sqrt{6}), each with its own within [0,1], until the accumulation of bifurcations at r \approx 3.56995 leads to a transition to . At r=4, iterations produce a filling [0,1] as the , where dense unstable periodic orbits and sensitive dependence on initial conditions prevail, yet the iterative process confines orbits to this invariant set.

References

  1. [1]
    [PDF] Fixed-Point Iteration - MATH 375 Numerical Analysis
    Definition. A fixed-point for a function f is a number p such that f(p) = p. Page 4. Equivalence. The process of root-finding and the process of finding fixed ...
  2. [2]
    4.2. Fixed-point iteration — Fundamentals of Numerical Computation
    Fixed point iteration for a differentiable g ( x ) converges to a fixed point p if the initial error is sufficiently small.<|control11|><|separator|>
  3. [3]
    [PDF] Fixed point iteration - Georgia State University
    Proof. 1. If g(a) = a or g(b) = b, then done. Otherwise, g(a) > a and g(b) < b. Define f (x) = x − g(x), then f (a) = a − g(a) < 0,.
  4. [4]
    Fixed point iteration
    Definition: A fixed point ${\bf x}^*$ of a function ${\bf g}({\bf x})$ is a point in its domain that is mapped to itself: $\displaystyle {\bf x}^*={\bf g} ...Missing: mathematical | Show results with:mathematical
  5. [5]
    [PDF] Chapter 3: The Contraction Mapping Theorem - UC Davis Math
    is called a fixed point of T. The contraction mapping theorem states that a strict contraction on a complete metric space has a unique fixed point. The ...
  6. [6]
    [PDF] 2.2. Fixed-Point Iteration
    Dec 17, 2022 · Definition. A number p is a fixed point for a given function g ... as to use Fixed-Point Iteration to solve the equation. Notice that ...
  7. [7]
    [PDF] a short survey of the development of fixed point theory
    The method of successive approximation introduced by Liouville in 1837 and systematically developed by Picard in 1890 culminated in formulation by Banach known ...
  8. [8]
  9. [9]
    [PDF] 2.2 Fixed-Point Iteration
    If 𝑔 𝑎 = 𝑎, or 𝑔 𝑏 = 𝑏, then 𝑔 has a fixed point at the endpoint. • Otherwise, 𝑔 𝑎 > 𝑎 and 𝑔 𝑏 < 𝑏. • Define a new function ℎ 𝑥 = 𝑔 𝑥 − 𝑥.Missing: mathematical | Show results with:mathematical
  10. [10]
    [PDF] Babylonian Method of Computing the Square Root
    Each iteration of the Babylonian algorithm consists of one division, one multiplication, and one addition. Since division is the most time-consuming of the ...
  11. [11]
    [PDF] Iterative Methods for Solution of Linear Equations
    Define r = r(x) = b − Ax, as the residual at x with respect to the system (A, b). The residual at the exact solution x? is zero. The fixed-point iteration.
  12. [12]
    Using cobwebbing as a graphical solution technique for discrete ...
    Cobwebbing is a graphical method to explore function iteration, moving vertically to the function's graph, then horizontally to the y-axis, and then to the x- ...Missing: fixed source
  13. [13]
    [PDF] Numerical Analysis, 9th ed.
    ... Numerical Analysis,. Ninth Edition. Richard L. Burden and J. Douglas Faires. Editor-in-Chief: Michelle Julet. Publisher: Richard Stratton. Senior Sponsoring ...
  14. [14]
    [PDF] Lecture notes 1 Fixed point iterations
    Feb 13, 2020 · 1.2 Stability of fixed points. Some fixed points are attracting and some are repelling, the following definitions classify these fixed points.
  15. [15]
    [PDF] 3 Fixed points — summary
    If |f0(x∗)| < 1 then we call x∗ an attracting fixed point. • If |f0(x∗)| > 1 then we call x∗ a repelling fixed point. • If |f0(x∗)| = 1 then we call x∗ a ...
  16. [16]
    [PDF] THE CONTRACTION MAPPING THEOREM 1. Introduction Let f
    We will discuss here the most basic fixed-point theorem in analysis. It is due to Banach and appeared in his Ph.D. thesis (1920, published in 1922). Theorem 1.1 ...<|control11|><|separator|>
  17. [17]
    Sur les opérations dans les ensembles abstraits et leur application ...
    Le but de cette note est d'établir quelques théorèmes valables pour différents champs fonctionnels.
  18. [18]
    [PDF] The Banach Fixed Point Theorem: selected topics from its hundred ...
    The Banach theorem is simple in its formulation, the fixed point is always unique and it is obtained by an explicit calculation. Its disadvantage is that the ...
  19. [19]
    [PDF] Fixed Point Methods in Nonlinear Analysis - UChicago Math
    Aug 29, 2014 · We begin with the Banach fixed point theorem, which we use to prove the inverse and implicit mapping theorems and the Picard-Lindelöf theorem ...
  20. [20]
    [PDF] Lecture 8 : Fixed Point Iteration Method, Newton's Method
    When does the sequence (xn) obtained from the iterative process (3) converge ? The following result is a consequence of the mean value theorem. Theorem 8.1: Let ...<|control11|><|separator|>
  21. [21]
    [PDF] FIXED POINT ITERATION We begin with a computational example ...
    We are going to use a numerical scheme called `fixed point iteration'. It amounts to making an initial guess of x0 and substituting this into the right side of ...
  22. [22]
    [PDF] Nonlinear Systems - Stanford Computer Graphics Laboratory
    Thus, in this case Ek is quadratic in Ek−1, so we say fixed point iteration can have quadratic con- vergence; notice this proof of quadratic convergence only ...
  23. [23]
    [PDF] 1. Fixed Point Iteration for Non-linear Equations - NTNU
    Fixed point iteration solves F(x)=0 by defining x(k+1) = Φ(x(k)), where x is a fixed point of Φ if x solves Φ(x) = x.
  24. [24]
    [PDF] Lecture 3: Solving Equations Using Fixed Point Iterations - cs.wisc.edu
    Sep 14, 2010 · We introduce now two notions concerning orders of convergence. 1.1.1 Linear Convergence. Linear convergence requires that the error is reduced ...
  25. [25]
    Nonlinear Systems of Equations: Fixed-Point Iteration Method
    The fixed-point iteration method proceeds by rearranging the nonlinear system such that the equations have the form. Note that any other norm function can work ...
  26. [26]
    Newton-Raphson Method (Univariate)
    The Newton-Raphson method can be considered as the fixed point iteration $x_{n+1}=g(x_n)$ based on $\displaystyle g(x)=xf(x)/f'(x)$.
  27. [27]
    Convergence of Newton-Raphson method: - NPTEL Archive
    Newton Raphson Method is said to have quadratic convergence. Note: Alternatively, one can also prove the quadratic convergence of Newton-Raphson method based on ...
  28. [28]
    [PDF] MATH 4513 Numerical Analysis Chapter 2. Solutions of Equations in ...
    Sep 3, 2020 · Secant Method converges slightly slower than Newton Method, but much faster than other Fixed-point iterations. Newton's method or the Secant ...<|control11|><|separator|>
  29. [29]
    [PDF] 7.3 The Jacobi and Gauss-Seidel Iterative Methods
    With the Jacobi method, the values of obtained in the th iteration remain unchanged until the entire th iteration has been calculated. With the Gauss-Seidel ...
  30. [30]
    [PDF] 1 Review of Fixed Point Iterations - cs.wisc.edu
    Sep 21, 2010 · In the secant method, we use the difference between consecutive values of xn for h, so that the approximation of the derivative should improve.
  31. [31]
    [PDF] Topic 3 Iterative methods for Ax = b
    Example. Solve the following system of equations using (a) 3 iterations of the Jacobi method and (b) 3 iterations of the. Gauss-Seidel method. Start with x = y ...
  32. [32]
    [PDF] Acceleration methods for fixed point iterations
    Nov 16, 2024 · The non-practical character of this technique prompted Peter Wynn in a 1956 article to explore an alternative implementation and this resulted ...
  33. [33]
    Anderson Acceleration for Fixed-Point Iterations
    This paper concerns an acceleration method for fixed-point iterations that originated in work of DG Anderson [J. Assoc. Comput. Mach., 12 (1965), pp. 547–560]
  34. [34]
    The chaos game on a general iterated function system
    Sep 13, 2010 · The chaos game on a general iterated function system. Published online by Cambridge University Press: 13 September 2010. MICHAEL F. BARNSLEY and.Missing: original | Show results with:original
  35. [35]
    Iterated function systems and the global construction of fractals
    BARNSLEY M and VINCE A (2010) The chaos game on a general iterated function system, Ergodic Theory and Dynamical Systems, 10.1017/S0143385710000428, 31:4 ...<|control11|><|separator|>
  36. [36]
    Fractals Everywhere - ScienceDirect.com
    Fractals Everywhere. Book • Second Edition • 1993. Author: MICHAEL F. BARNSLEY ... This 10-chapter text is based on a course called "Fractal Geometry", which ...
  37. [37]
    Attractor - Scholarpedia
    Nov 3, 2006 · An attracting set for a dynamical system is a closed subset A of its phase space such that for many choices of initial point the system will evolve towards A.
  38. [38]
    [PDF] Discrete Dynamical Systems: The Linear, the ... - Ohio University
    For example, if x∗ is a locally asymptotically stable fixed point, then {x∗} is an attractor. More generally, if x is locally asymptotically stable fixed ...
  39. [39]
    [PDF] 2 Discrete Dynamical Systems: Maps - Complexity Sciences Center
    The sets of points whose orbits converge to an attractor of a system is called the basin of attraction of this point. From Figure 2.5 we see that the unstable ...