Fact-checked by Grok 2 weeks ago

Heun's method

Heun's method is an explicit, second-order technique for solving initial value problems of ordinary differential equations (ODEs) of the form y' = f(t, y), y(t_0) = y_0, named after the German mathematician Karl Heun who introduced it in 1900 as part of early developments in . It improves upon the first-order by employing a predictor-corrector approach, where an initial prediction is made using the Euler step, followed by a correction that averages the function evaluations (slopes) at the current point and the predicted endpoint to approximate the integral more accurately using the . The method's algorithm proceeds in two stages per step of size h: first, compute the predictor \tilde{y}_{n+1} = y_n + h f(t_n, y_n), then evaluate the corrector slope f(t_{n+1}, \tilde{y}_{n+1}), and update the solution as y_{n+1} = y_n + \frac{h}{2} \left[ f(t_n, y_n) + f(t_{n+1}, \tilde{y}_{n+1}) \right]. This formulation yields a local truncation error of O(h^3) and global error of O(h^2), making it suitable for moderately accurate approximations of non-stiff ODEs, though it requires two function evaluations per step. As a foundational member of the Runge–Kutta family, Heun's method bridges simple one-stage methods like Euler's and higher-order variants such as the classical fourth-order Runge–Kutta method, and it can be extended to systems of ODEs or differential equations with appropriate modifications. While effective for educational purposes and basic simulations, its fixed step size limits efficiency for stiff problems, where implicit methods are preferred; however, adaptive step-size implementations enhance its practical utility in computational software.

Background

Numerical solution of ODEs

The (IVP) for an (ODE) is typically formulated as y'(t) = f(t, y(t)), subject to the initial condition y(t_0) = y_0, where f is a given and the goal is to determine the y(t) over some interval containing t_0. This setup ensures the existence and uniqueness of a solution under suitable conditions, such as continuity and of f with respect to y. Exact analytical solutions to IVPs are available only for a limited class of ODEs, particularly linear ones; for nonlinear ODEs, closed-form solutions are rare due to the complexity of integrating such equations, often requiring or approximations even when solvable. Nonlinearities can lead to phenomena like finite-time or non-uniqueness, further complicating analytical resolution. Consequently, numerical methods become essential for approximating solutions in practical applications, such as modeling physical systems in and sciences. Numerical approaches to solving IVPs involve discretizing the time domain into a grid of points t_n = t_0 + n h, where h > 0 is the step size, and approximating the values y(t_n) \approx y_n at these nodes. These methods advance the approximation step by step using schemes to estimate the y'(t_n), effectively replacing the continuous with a over the grid. A key concept in these methods is the local truncation error, which measures the discrepancy between the exact solution and the numerical approximation over a single step, assuming exact input data from the previous step; this error arises from approximating the smooth solution curve with a linear or higher-order . For instance, in simple schemes like Euler's method, the local is on the order of O(h^2). Over multiple steps, these local errors accumulate to form the error, which represents the total deviation of the numerical solution from the true solution at t_n; of the method ensures this accumulation remains bounded, typically resulting in a global error of order one less than the local truncation error for consistent schemes.

Euler's method

Euler's method, also known as the forward Euler method, is a explicit numerical technique for approximating solutions to equations (ODEs) of the form y' = f(t, y) with y(t_0) = y_0. It serves as the simplest one-step method in , relying on the at the current point to advance the solution by a fixed step size h. Developed in the and formalized in modern numerical contexts, it provides a foundational approach but exhibits limitations that have spurred more accurate variants. The algorithm proceeds iteratively as follows:
  1. Initialize t_0 = t_{\text{start}}, y_0 = y(t_0), and choose step size h > 0.
  2. For each step n = 0, 1, 2, \dots, until t_n \geq t_{\text{end}}:
    • Compute y_{n+1} = y_n + h f(t_n, y_n).
    • Update t_{n+1} = t_n + h.
  3. The approximate solution is the sequence \{y_n\} at points \{t_n\}.
This can be expressed in formula form as y_{n+1} = y_n + h f(t_n, y_n). Geometrically, Euler's method interprets the solution curve as a sequence of straight-line segments, each tangent to the direction field defined by f(t, y) at (t_n, y_n). This steps forward along the tangent line, effectively polygonalizing the by connecting these tangent segments, known as the Euler polygon. The , which measures the discrepancy per step assuming previous steps are exact, arises from the expansion of the exact solution around t_n: y(t_n + h) = y(t_n) + h y'(t_n) + \frac{h^2}{2} y''(\xi) for some \xi \in (t_n, t_n + h), where y'(t_n) = f(t_n, y(t_n)). Substituting the Euler step yields an error of \frac{h^2}{2} y''(\xi) = O(h^2). For illustration, consider the y' = y, y(0) = [1](/page/1), whose exact solution is y(t) = e^t. With step size h = 0.1, the first step computes y_1 = 1 + 0.1 \cdot f(0, 1) = 1 + 0.1 \cdot 1 = [1.1](/page/1.1), while the exact value at t = 0.1 is e^{0.1} \approx 1.10517, showing a small local discrepancy. Despite its simplicity, Euler's method is only first-order accurate, accumulating a global error of O(h) over an interval, which can lead to significant deviations for problems that are stiff—characterized by widely varying timescales—or highly oscillatory, where small steps are required for but increase computational cost. These shortcomings motivate improvements such as Heun's method, which averages slopes for better accuracy.

Formulation

Algorithm steps

Heun's method is an explicit two-stage technique for approximating solutions to initial value problems of the form y' = f(t, y), y(t_0) = y_0, where y is the solution function and f is a given . The method proceeds iteratively over discrete time steps, producing a sequence of approximations \{ y_n \} at grid points t_n = t_0 + n h, where h > 0 is the fixed step size and n = 0, 1, \dots, N with t_N approximating the desired end time. To implement Heun's method, the required inputs are the step size h > 0, the y_0, the initial time t_0, the end time T, and the function f(t, y). The algorithm begins at n = 0 with y_0 and advances step by step until t_n \geq T, outputting the sequence \{ y_n \}. The core update at each iteration n consists of a predictor step to estimate an intermediate value \hat{y}_{n+1} = y_n + h f(t_n, y_n), followed by a corrector step that computes the final as y_{n+1} = y_n + \frac{h}{2} \left[ f(t_n, y_n) + f(t_{n+1}, \hat{y}_{n+1}) \right], where t_{n+1} = t_n + h. This predictor step aligns with the forward Euler update, providing a provisional estimate before refinement. The following outlines the iterative procedure for a scalar :
function HeunMethod(f, t0, y0, h, T)
    n = 0
    t = t0
    y = y0
    while t < T do
        k1 = f(t, y)
        y_hat = y + h * k1
        k2 = f(t + h, y_hat)
        y = y + (h / 2) * (k1 + k2)
        t = t + h
        n = n + 1
        store y as y_n  // optional, for output sequence
    end while
    return sequence {y_n for n=0 to N}
end function
This implementation assumes uniform steps and stores approximations for later use if needed. For systems of ODEs, where y \in \mathbb{R}^m and f: \mathbb{R} \times \mathbb{R}^m \to \mathbb{R}^m, the method extends component-wise by treating vectors throughout: the predictor becomes \hat{y}_{n+1} = y_n + h f(t_n, y_n) with vector addition and scalar multiplication, and the corrector averages the vector-valued slopes similarly. The pseudocode adapts directly by replacing scalar operations with vector equivalents, enabling solution of coupled systems like predator-prey models.

Predictor-corrector approach

Heun's method functions as a predictor-corrector scheme, designed to enhance the accuracy of numerical solutions to initial value problems for ordinary differential equations of the form y' = f(t, y), y(t_0) = y_0. Introduced by in 1900, this approach refines the first-order by incorporating an explicit prediction followed by a correction step. The predictor phase employs the to generate an initial approximation \hat{y}_{n+1} at the end of the interval [t_n, t_n + h]:
\hat{y}_{n+1} = y_n + h f(t_n, y_n).
This estimate assumes a constant slope equal to the derivative at the starting point, providing a linear extrapolation along the tangent.
In the corrector phase, the prediction is adjusted using the trapezoidal rule, which averages the slopes at the current and predicted points to better approximate the integral of f over the step:
y_{n+1} = y_n + \frac{h}{2} \left[ f(t_n, y_n) + f(t_n + h, \hat{y}_{n+1}) \right].
Geometrically, the predictor traces the tangent from (t_n, y_n), while the corrector constructs a secant line averaging the tangents at the start and the predicted endpoint, yielding a path that more closely follows the solution curve's average direction. This averaging captures curvature effects neglected by the 's single tangent, resulting in second-order local accuracy by mimicking the trapezoidal integration's quadratic error term.
Although the standard formulation of Heun's method applies the corrector once for computational efficiency, iterative variants repeat the corrector step—re-evaluating the slope with the updated estimate—until convergence, potentially reducing truncation error further at the cost of additional function evaluations. These iterations align the method more closely with implicit schemes but are not part of the classical explicit version.

Derivation

Taylor series expansion

The Taylor series expansion of the exact solution to the initial value problem y' = f(t, y), y(t_n) = y_n, around t_n is given by y(t_n + h) = y_n + h y'(t_n) + \frac{h^2}{2} y''(t_n) + O(h^3), where higher-order terms are neglected for the local analysis up to second order. Since y'(t_n) = f(t_n, y_n), the second derivative y''(t_n) is obtained via the chain rule as y''(t_n) = \frac{\partial f}{\partial t}(t_n, y_n) + \frac{\partial f}{\partial y}(t_n, y_n) f(t_n, y_n), where the partial derivatives f_t and f_y account for the dependence of f on both t and y. Heun's method approximates the solution at the next step as y_{n+1} = y_n + \frac{h}{2} \left[ f(t_n, y_n) + f(t_n + h, y_n + h f(t_n, y_n)) \right], which incorporates an average of the function value at the current point and at a predicted point one step ahead. To verify the order of accuracy, the predictor term f(t_n + h, \hat{y}_{n+1}), where \hat{y}_{n+1} = y_n + h f(t_n, y_n), is expanded using the multivariate Taylor series: f(t_n + h, \hat{y}_{n+1}) = f(t_n, y_n) + h \frac{\partial f}{\partial t}(t_n, y_n) + h f(t_n, y_n) \frac{\partial f}{\partial y}(t_n, y_n) + O(h^2). This expansion captures the first-order changes in both arguments of f. Substituting this expansion into the Heun approximation yields y_{n+1} = y_n + h f(t_n, y_n) + \frac{h^2}{2} \left[ \frac{\partial f}{\partial t}(t_n, y_n) + \frac{\partial f}{\partial y}(t_n, y_n) f(t_n, y_n) \right] + O(h^3), which matches the exact Taylor series up to the second-order term O(h^2); thus, the local truncation error is O(h^3).

Butcher tableau representation

Heun's method belongs to the family of explicit Runge-Kutta methods and is compactly represented using the Butcher tableau, which organizes the coefficients \mathbf{c}, A, and \mathbf{b} in a standardized format. In general, an s-stage explicit Runge-Kutta method computes intermediate stages as k_i = f(t_n + c_i h, y_n + h \sum_{j=1}^{i-1} a_{ij} k_j), \quad i = 1, \dots, s, followed by the solution update y_{n+1} = y_n + h \sum_{i=1}^s b_i k_i. Heun's method corresponds to s = 2, with the Butcher tableau \begin{array}{c|cc} 0 & 0 & \\ 1 & 1 & 0 \\ \hline & \frac{1}{2} & \frac{1}{2} \end{array} where \mathbf{c} = \begin{pmatrix} 0 \\ 1 \end{pmatrix}, the matrix A = \begin{pmatrix} 0 & 0 \\ 1 & 0 \end{pmatrix}, and \mathbf{b}^\top = \left( \frac{1}{2}, \frac{1}{2} \right). This tableau yields the stage values k_1 = f(t_n, y_n) and k_2 = f(t_n + h, y_n + h k_1), with the update y_{n+1} = y_n + h \left( \frac{1}{2} k_1 + \frac{1}{2} k_2 \right). These steps precisely reproduce the explicit predictor-corrector formulation of , where the predictor uses the Euler step and the corrector averages the slopes. The coefficients in the tableau satisfy the simplifying assumptions and order conditions for a second-order method, specifically \sum_{i=1}^2 b_i = 1 and \sum_{i=1}^2 b_i c_i = \frac{1}{2}.

Properties

Order of accuracy

Heun's method is a second-order explicit , meaning it achieves a global error of order O(h^2) for a step size h. This order arises from its design, which matches the first two terms of the Taylor series expansion of the exact solution, leading to a local truncation error of O(h^3). The local truncation error is derived by expanding the exact solution and the numerical approximation around a grid point, revealing that the principal error term involves third derivatives of the solution or function f, assuming sufficient smoothness. Under standard assumptions for initial value problems (IVPs) of the form y' = f(t,y) with y(t_0) = y_0, where f satisfies a Lipschitz condition in y (ensuring uniqueness and bounded growth) and is sufficiently differentiable, the global error over a fixed interval [t_0, T] is bounded by C h^2 for some constant C > 0 independent of h. This follows from the theorem for one-step methods, which links the global error order to the local order minus one, provided the method is consistent and zero-stable. The method is consistent because the local approaches zero as h \to 0, satisfying the necessary condition for . In practical implementations, the choice of step size h involves a between (which decreases with smaller h) and (which accumulates and dominates for excessively small h due to limitations in evaluating f). For Heun's method, as an explicit two-stage process, round-off propagation remains moderate, but optimal accuracy requires balancing these errors to minimize the total computed error.

Stability analysis

The stability of Heun's method is analyzed using the linear test equation y' = \lambda y, where \lambda \in \mathbb{C} with \operatorname{Re}(\lambda) < 0, which models the behavior of dissipative systems. Applying the method to this equation yields the recurrence y_{n+1} = R(z) y_n, where z = h \lambda and the stability function is the quadratic polynomial R(z) = 1 + z + \frac{1}{2} z^2. The region of absolute stability is the set of z \in \mathbb{C} such that |R(z)| \leq 1, ensuring that the numerical solution remains bounded as n \to \infty for fixed h > 0. This region lies in the left half of the , symmetric about the , and forms a bounded area that includes a wedge-shaped portion around the negative , extending up to approximately z = -2 along the real line. Compared to the forward Euler method's stability region—a disk of 1 centered at z = -1—Heun's region is larger overall, particularly in directions away from the real axis, but it remains finite and does not encompass the entire left half-plane. As an explicit two-stage Runge-Kutta method, Heun's method is not A-stable, meaning its stability region excludes parts of the negative real axis for sufficiently large |z|, leading to instability when the step size h is too large relative to $1/|\lambda|. In the context of the Dahlquist test equation, the method achieves second-order accuracy but is not L-stable, as |R(z)| grows unbounded for large |z| due to the nature of the function, unlike implicit L-stable methods where R(z) \to 0 as |z| \to \infty. For practical applications in dissipative systems with real negative eigenvalues, stability requires h < 2 / |\lambda|, the same restriction as forward Euler, which limits the method's utility for stiff problems where large |\lambda| demand very small steps to avoid oscillations or divergence. This conditional underscores the need for adaptive step sizing or switching to implicit methods in such scenarios.

Applications

Example computation

To illustrate the application of Heun's method, consider the y' = -y + 1 - x, with y(0) = 3, over the interval [0, 1]. The exact solution is y(x) = 2 - x + e^{-x}. This test problem is linear and solvable analytically, making it suitable for verifying numerical approximations. We apply Heun's method with step size h = 0.1, computing the first two steps explicitly to demonstrate the predictor-corrector . For the first step, at t_0 = 0, y_0 = 3:
The predictor is \tilde{y}_1 = y_0 + h f(t_0, y_0), where f(t, y) = -y + 1 - t.
Thus, f(t_0, y_0) = -3 + 1 - 0 = -2, so \tilde{y}_1 = 3 + 0.1 \cdot (-2) = 2.8.
The corrector is y_1 = y_0 + \frac{h}{2} \left[ f(t_0, y_0) + f(t_0 + h, \tilde{y}_1) \right].
Here, f(t_0 + h, \tilde{y}_1) = f(0.1, 2.8) = -2.8 + 1 - 0.1 = -1.9,
so y_1 = 3 + 0.05 \cdot (-2 + (-1.9)) = 3 + 0.05 \cdot (-3.9) = 3 - 0.195 = 2.805.
For the second step, at t_1 = 0.1, y_1 = 2.805:
The predictor is \tilde{y}_2 = 2.805 + 0.1 \cdot f(0.1, 2.805).
f(0.1, 2.805) = -2.805 + 1 - 0.1 = -1.905, so \tilde{y}_2 = 2.805 + 0.1 \cdot (-1.905) = 2.6145.
The corrector is y_2 = 2.805 + 0.05 \cdot [f(0.1, 2.805) + f(0.2, 2.6145)].
f(0.2, 2.6145) = -2.6145 + 1 - 0.2 = -1.8145,
so y_2 = 2.805 + 0.05 \cdot (-1.905 - 1.8145) = 2.805 + 0.05 \cdot (-3.7195) = 2.805 - 0.185975 = 2.61903.
Continuing this process for additional steps yields the approximations shown in the table below, including the exact values and absolute errors |y_n - y(t_n)|. The computations were performed to five places for precision.
t_nHeun approximation y_n y(t_n) error
0.03.000003.000000.00000
0.12.805002.804840.00016
0.22.619032.618730.00030
0.32.441222.440820.00040
0.42.270802.270320.00048
0.52.107082.106530.00055
Errors remain on the order of $10^{-4} or smaller across these steps, consistent with the second-order accuracy of Heun's method. For comparison, applying the (first-order) with the same h = 0.1 to this problem yields larger errors: at t = 0.1, the approximation is 2.80000 with error 0.00484; at t = 0.2, it is 2.61000 with error 0.00873. Thus, Heun's errors are approximately smaller, highlighting its improved accuracy over for the same step size. In practice, Heun's method can be implemented in programming languages like or using a simple : initialize y and t with the given values, then iteratively compute the predictor \tilde{y} = y + h \cdot f(t, y), followed by the corrector y = y + \frac{h}{2} [f(t, y) + f(t + h, \tilde{y})], and update t = t + h. Vectorized forms are possible for efficiency, but the explicit predictor-corrector structure ensures clarity.

Comparisons and extensions

Heun's method offers improved accuracy over the forward , which is a technique with global error proportional to the step size h. As a second-order method, Heun's approach reduces the global error to O(h^2), allowing for roughly twice the step size of Euler's method to achieve comparable accuracy or halving the error for the same step size in typical non-stiff problems. Compared to the classical fourth-order Runge-Kutta (RK4) method, Heun's method is simpler, requiring only two function evaluations per step versus four for RK4, making it computationally less intensive for applications where second-order accuracy suffices. However, RK4 provides higher precision with global error O(h^4), rendering it preferable for non-stiff ordinary differential equations demanding greater accuracy without excessive computational cost. Heun's method is a specific implementation of the modified , also known as the explicit trapezoidal or improved , where it employs a predictor-corrector strategy to average the slope at the initial point and a predicted . The method traces its origins to Carl Runge's 1895 work on techniques, which introduced early multistage approaches, with Karl Heun formalizing the second-order variant in 1900 as part of broader Runge-Kutta developments. Extensions of Heun's method include adaptive step-size control, often achieved by pairing it with embedded lower-order estimators like the to dynamically adjust h based on local estimates, enhancing efficiency for varying solution behaviors. Such adaptive variants appear in numerical software libraries, analogous to MATLAB's ode23 solver, which implements a low-order explicit Runge-Kutta pair with step-size adaptation for non-stiff problems. As an explicit method, Heun's approach is not well-suited for stiff ordinary differential equations, where stability requires impractically small step sizes due to the method's stability function growing unbounded for large negative eigenvalues.

References

  1. [1]
    [PDF] NUMERICAL SOLUTION OF ORDINARY DIFFERENTIAL EQUATIONS
    known as Heun's method. The Heun method is still of second-order accuracy ... Rewrite the original equation (12.2) using the trapezoidal numerical integration ...
  2. [2]
    [PDF] Euler's and Heun's method
    Oct 3, 2021 · The topic of this note is the numerical solution of systems of ordinary differential equations (ODEs). This has been discussed in previous ...
  3. [3]
  4. [4]
    None
    ### Summary of Heun's Method from Section 6: Ordinary Differential Equations
  5. [5]
    [PDF] An Introduction to Numerical Methods for ODEs
    We will introduce first-order differential equations, providing examples with and without solutions, as well as a statement on existence and uniqueness. We then ...
  6. [6]
    [PDF] Chapter 5 Methods for ordinary differential equations
    Initial-value problems (IVP) are those for which the solution is entirely known at some time, say t = 0, and the question is to solve the ODE.<|control11|><|separator|>
  7. [7]
    Solving Ordinary Differential Equations I - SpringerLink
    This book is a valuable tool for students of mathematics and specialists concerned with numerical analysis, mathematical physics, mechanics, system engineering.
  8. [8]
    Euler Forward Method -- from Wolfram MathWorld
    A method for solving ordinary differential equations using the formula y_(n+1)=y_n+hf(x_n,y_n), which advances a solution from x_n to x_(n+1)=x_n+h.
  9. [9]
    [PDF] Euler's Method - Whitman People
    From this, we say that the local truncation error is approximately proportional to h2. It can be shown that the global truncation error for Euler's method is ...
  10. [10]
    MATHEMATICA TUTORIAL, Part 1.3: Heun Methods
    The Improved Euler's method, also known as the Heun formula or the average slope method, gives a more accurate approximation than the Euler rule.
  11. [11]
    [PDF] A history of Runge-Kutta methods f ~(z) dz = (x. - x.-l) - People
    This paper constitutes a centenary survey of Runge--Kutta methods. It reviews some of the early contributio~ due to Runge, Heun, Kutta ... In 1900, K. Heun ...Missing: Karl | Show results with:Karl
  12. [12]
    Numerical Methods - Heun Method | San Joaquin Delta College
    Heun's clever approach to this problem is to consider the tangent lines to the solution curve at both ends of the interval we are investigating. As we have ...Missing: integration ODE
  13. [13]
    [PDF] multistage and predictor–corrector methods
    The midpoint method, also called modified Euler, and the improved Euler (Heun) method are two-stage methods for solving ODEs. Heun is a predictor-corrector ...Missing: approach | Show results with:approach
  14. [14]
    MATHEMATICA TUTORIAL, Part 1.3: Heun Method
    Our next step in this direction includes Heun's method, which was named after a German mathematician Karl Heun (1859--1929), who made significant contributions ...
  15. [15]
    None
    ### Summary of Taylor Series Derivation for Heun's Method
  16. [16]
    8.03: Runge-Kutta 2nd-Order Method for Solving Ordinary ...
    Oct 5, 2023 · To understand the Runge-Kutta 2nd order method, we need to first derive Euler's method from the Taylor series. ... Heun's method formula is given ...Lesson 1: Theory of Runge... · Runge-Kutta 2nd order method · Heun's Method
  17. [17]
    [PDF] Numerical Solutions of Differential Equations (2)
    expanded into a Taylor series in h: ○ I.e. Expansion of Heun's method and of exact solution coincide up to and including 2nd order y(t n. +h)≈ y(t n. )+h y ...
  18. [18]
    [PDF] High order Runge-Kutta methods
    Oct 20, 2021 · Applying order conditions to Heun's method. Apply the conditions to Heun's method, for which s = 2 and the Butcher tableau is c1 a11 a12 c2.
  19. [19]
    [PDF] Numerical solution of ordinary differential equations
    Apr 14, 2020 · 5.1 Local and global truncation error of OSM ... Thanks to Theorem 5.1, we can conclude that Heun's method is convergent of order 2.
  20. [20]
    [PDF] CHAPTER VI NUMERICAL INTEGRATION OF ORDINARY ...
    with a local truncation error of O(h3). 2. The Trapezoidal Rule method: Heun's method. The second Runge Kutta method of order 2- Heun's method- is also called.
  21. [21]
    [PDF] Internal error propagation in explicit Runge--Kutta methods
    Eventually, the truncation errors become so small that the accumulation of roundoff errors is dominant and the overall error cannot be decreased further.
  22. [22]
    [PDF] AM 213B Prof. Daniele Venturi Absolute stability of numerical ...
    ... stability of the Heun method (shaded area). ... It is seen in Figure 6 that these methods are in fact not A-stable. Absolute stability analysis of Runge-Kutta ...
  23. [23]
    [PDF] Numerical Methods for ODE - Dipartimento di Matematica
    – Local Truncation Error (LTE error for a single step). – propagated truncation error (approximations produced during the previous steps) (zero-stability).<|control11|><|separator|>
  24. [24]
    [PDF] First order numerical methods
    6 Example (Euler and Heun Methods) Solve y′ = −y + 1 − x, y(0) = 3 by both Euler's method and Heun's method for x = 0 to x = 1 in steps of h = 0.1. Produce a ...
  25. [25]
    [PDF] Euler, Heun and RK4 Methods Mathematics 5410
    Compute 20 data points (tn,yn) using the Runge-Kutta 4 method, as follows: Step 1. Define t0 = 0, y0 = 1, from the initial condition y(0) = 1. Step 2 ...
  26. [26]
    [PDF] Chapter 08.03 Runge-Kutta 2nd Order Method for Ordinary ...
    Oct 13, 2010 · ... Euler. Heun θ(K. ) Figure 4 Comparison of Euler and Runge Kutta methods with exact results over time. How do these three methods compare with ...
  27. [27]
    [PDF] Runge-Kutta 4 (and Other Numerical Methods for ODEs)
    We will follow Kutta's seminal paper [Kutta, 1901] which contains the pertinent contributions by Euler, Heun, and Runge as special cases. 4 On pg. 437 Kutta ...
  28. [28]
    Runge–Kutta Methods for Ordinary Differential Equations
    Since their first discovery by Runge (Math Ann 46:167-178, 1895), Heun (Z Math Phys 45:23-38, 1900) and Kutta (Z Math Phys 46:435-453, 1901), Runge-Kutta ...
  29. [29]
    ODE Solvers · DifferentialEquations.jl
    Heun - The second order Heun's method. Uses embedded Euler method for adaptivity.
  30. [30]
    [PDF] Adaptive Methods and Stiff Systems
    Use ode23 to solve the following ODE from t = 0 to 4;. Where y(0) = 0.5. Obtain solutions for the default (10-3) and for a more stringent (10- ...Missing: extensions | Show results with:extensions
  31. [31]
    ode23 - Solve nonstiff differential equations — low order method
    ode23 solves nonstiff differential equations using a low-order method, integrating from t0 to tf with initial conditions y0. It is an explicit Runge-Kutta (2,3 ...Missing: Heun's | Show results with:Heun's
  32. [32]
    [PDF] Stiff differential equations.
    Nov 3, 2020 · For all explicit methods, like Euler's and Heun's, the stability function will be a polynomial, and |R(z)|→∞ as <z → −∞. Explicit methods can ...