Fact-checked by Grok 2 weeks ago

Euler–Maruyama method

The Euler–Maruyama method is a numerical approximation scheme for solving stochastic differential equations (SDEs) of the form dX_t = f(t, X_t) dt + g(t, X_t) dW_t, where f and g are the drift and diffusion coefficients, respectively, and W_t is a standard ; it discretizes the time interval into steps of size \Delta t and approximates the solution at each step as X_{n} = X_{n-1} + f(t_{n-1}, X_{n-1}) \Delta t + g(t_{n-1}, X_{n-1}) \Delta W_n, with \Delta W_n being independent Gaussian increments simulating the . Introduced by Japanese mathematician Gisiro Maruyama in his 1955 paper on continuous Markov processes, the method adapts Leonhard Euler's classical polygonal approximation for deterministic ordinary differential equations to the stochastic setting by incorporating Itô stochastic integrals, providing a practical tool for simulating paths of diffusion processes where analytical solutions are unavailable. Under standard and growth conditions on f and g, the Euler–Maruyama scheme exhibits strong convergence of order 0.5, meaning the expected root-mean-square error over paths converges as O((\Delta t)^{1/2}), and of order 1, where the error in expectations of smooth test functions converges as O(\Delta t). Widely applied in fields such as for pricing derivative securities under models like , population dynamics with environmental noise, and physical systems involving random fluctuations, the method serves as a foundational for more advanced integrators like the Milstein scheme, despite its limitations in accuracy for multiplicative noise or stiff equations. Its simplicity and implementability in standard programming environments make it accessible for simulations, though care must be taken with step size selection to balance computational cost and precision.

Historical Development

Origins and Naming

The Euler–Maruyama method represents a stochastic analogue to the originally developed for solving ordinary differential equations (ODEs), extending it to handle the randomness inherent in stochastic differential equations (SDEs). This adaptation emerged in the mid-20th century as computational tools became available for simulating complex systems influenced by noise, building on the deterministic Euler scheme introduced by Leonhard Euler in the for of ODEs. The method derives its name from Leonhard Euler (1707–1783), whose forward Euler technique for ODEs serves as the foundational discretization strategy, and Gisiro Maruyama (1916–1986), a who pioneered its stochastic formulation. Maruyama's contributions focused on adapting the Euler approach to incorporate the increments of processes, essential for modeling continuous-time dynamics. During the 1940s and 1950s, the development of numerical methods for SDEs was motivated by the need to approximate solutions in physical systems exhibiting random fluctuations, such as in processes, and in engineering applications involving noisy signals and control systems. These efforts coincided with the formalization of , providing the rigorous framework for SDEs. A pivotal advancement came in Maruyama's 1955 paper, "Continuous Markov processes and stochastic equations," which introduced the core scheme for processes, laying the groundwork for practical numerical simulations of such equations.

Key Theoretical Contributions

The Euler–Maruyama method extends the classical for deterministic ordinary differential equations to stochastic differential equations by incorporating an approximation for the stochastic integral. In 1955, G. Maruyama established the foundational strong convergence of order 0.5 for the method, proving in quadratic mean to the true solution under appropriate conditions on the coefficients. During the 1970s, researchers including G.N. Milstein demonstrated of order 1 for the Euler–Maruyama method, providing essential error analysis that confirmed its reliability for moments and expectations. Milstein's contributions in this period, while extending to higher-accuracy schemes like the with strong order 1.0, also solidified the foundational error bounds for the basic Euler–Maruyama approximation through stochastic Taylor expansions. These convergence results hold when the drift and diffusion coefficients satisfy global and linear growth bounds, ensuring the and of strong solutions to the underlying . Kloeden and Platen's 1992 monograph offers a thorough treatment of these properties, deriving explicit error bounds and conditions for both strong and weak convergence in multi-dimensional settings.

Mathematical Formulation

The Euler–Maruyama method provides a discrete-time approximation to the solution of an Itô stochastic differential equation () driven by a . Consider the general scalar Itô of the form dX_t = a(t, X_t) \, dt + b(t, X_t) \, dW_t, \quad t \in [0, T], with deterministic initial condition X_0 = x_0 \in \mathbb{R}, where W_t is a standard Wiener process and the drift coefficient a: [0, T] \times \mathbb{R} \to \mathbb{R} and diffusion coefficient b: [0, T] \times \mathbb{R} \to \mathbb{R} are measurable functions satisfying local integrability conditions that ensure the Itô integrals exist. The method proceeds by discretizing the time interval [0, T] into a partition $0 = t_0 < t_1 < \cdots < t_N = T, with time steps \Delta t_n = t_{n+1} - t_n for n = 0, 1, \dots, N-1. The approximation Y_n to X_{t_n} is then defined recursively by the scheme \begin{equation} Y_{n+1} = Y_n + a(t_n, Y_n) \Delta t_n + b(t_n, Y_n) \Delta W_n, \quad Y_0 = x_0, \end{equation} where the increments \Delta W_n = W_{t_{n+1}} - W_{t_n} are independent and normally distributed as \Delta W_n \sim \mathcal{N}(0, \Delta t_n). The sequence of approximations \{Y_n\}_{n=0}^N constitutes a discrete-time , since each step depends solely on the previous approximation Y_n and the independent noise increment \Delta W_n.

Derivation from Itô Calculus

The Itô stochastic differential equation (SDE) for an Itô process X_t is expressed in integral form as X_t = X_0 + \int_0^t a(s, X_s) \, ds + \int_0^t b(s, X_s) \, dW_s, where a(t, x) and b(t, x) denote the drift and diffusion coefficients, respectively, and W_t is a standard . To obtain a numerical approximation, discretize the time interval [0, T] into a sequence of points $0 = t_0 < t_1 < \cdots < t_m = T, with increments \Delta t_n = t_{n+1} - t_n. Over each subinterval [t_n, t_{n+1}], the integral form yields X_{t_{n+1}} = X_{t_n} + \int_{t_n}^{t_{n+1}} a(s, X_s) \, ds + \int_{t_n}^{t_{n+1}} b(s, X_s) \, dW_s. The Euler–Maruyama scheme approximates these integrals by freezing the coefficients at their values at the left endpoint t_n, assuming they remain approximately constant over the small interval \Delta t_n. This leads to the recursive scheme X_{n+1} = X_n + a(t_n, X_n) \Delta t_n + b(t_n, X_n) \Delta W_n, where \Delta W_n = W_{t_{n+1}} - W_{t_n} is the Wiener increment, normally distributed with mean zero and variance \Delta t_n. This freezing of coefficients is justified under the assumption that a and b are continuous (or Lipschitz continuous) in their arguments, ensuring that their variation over \Delta t_n becomes negligible as \Delta t_n \to 0. The Wiener increment \Delta W_n satisfies \mathbb{E}[\Delta W_n] = 0 and \mathbb{E}[(\Delta W_n)^2] = \Delta t_n, which matches the quadratic variation of the in the mean-square sense, providing a consistent approximation for the stochastic integral. The Euler–Maruyama scheme arises naturally from the Itô-Taylor expansion of the SDE solution, which generalizes the classical Taylor series to stochastic processes using . The expansion includes terms involving multiple Itô integrals; the Euler–Maruyama method retains only the zeroth-order terms for both the drift (ordinary integral) and diffusion (), truncating all higher-order multiple stochastic integrals. This truncation introduces a local truncation error of order O(\Delta t_n) in the mean-square sense for sufficiently small \Delta t_n.

Numerical Properties

Strong Convergence

Strong convergence quantifies the pathwise approximation quality of the Euler–Maruyama method to the true solution of a stochastic differential equation (SDE), focusing on the almost sure behavior of sample paths. It is defined such that the root-mean-square of the supremum error tends to zero as the discretization step size \Delta t approaches zero: \left( \mathbb{E} \left[ \sup_{0 \leq t \leq T} |X_t - Y_t|^2 \right] \right)^{1/2} \to 0, where X_t denotes the exact solution and Y_t the continuous-time extension of the Euler–Maruyama approximation over the interval [0, T]. This metric captures the expected maximum deviation across paths, emphasizing reliable trajectory tracking rather than statistical moments. Under standard regularity conditions—namely, global Lipschitz continuity of the drift a(x) and diffusion b(x) coefficients, along with linear growth bounds |a(x)| + |b(x)| \leq K(1 + |x|) for some K > 0—the method exhibits a strong order of convergence of 0.5. In this setting, the error bound takes the form \left( \mathbb{E} \left[ \sup_{0 \leq t \leq T} |X_t - Y_t|^2 \right] \right)^{1/2} \leq C \sqrt{\Delta t}, with C depending on T, the constant, and the second of the \mathbb{E}[|X_0|^2] < \infty. These assumptions ensure existence, uniqueness, and moment stability of the SDE solution, enabling the discretization to mimic the path properties effectively. The proof establishes this order by analyzing the error process e_t = X_t - Y_t, decomposing it into local truncation errors from the Itô-Taylor expansion, and bounding the accumulated discrepancies. Key tools include the Itô isometry, which controls the variance of stochastic increments via \mathbb{E}\left[ \left( \int_0^T |b(X_u) - b(Y_u)|^2 du \right) \right], and Gronwall's inequality, applied to the mean-square error supremum Z(t) = \sup_{0 \leq s \leq t} \mathbb{E}[|e_s|^2] to yield Z(T) \leq C \Delta t. This framework confirms the \sqrt{\Delta t} scaling inherent to the Brownian motion increments. For SDEs featuring additive noise (constant b) or multiplicative noise where the drift satisfies a one-sided Lipschitz condition alongside a globally Lipschitz diffusion, the strong error remains O(\sqrt{\Delta t}). Here, the one-sided condition on the drift, such as \langle a(x) - a(y), x - y \rangle \leq L |x - y|^2 for L \geq 0, combined with polynomial growth, preserves the order while relaxing global monotonicity requirements on the drift. These extensions broaden applicability to dissipative systems without sacrificing the fundamental convergence rate.

Weak Convergence

Weak convergence assesses the Euler–Maruyama method's performance in approximating statistical properties of the solution to a stochastic differential equation (SDE), such as expectations and moments, rather than individual sample paths. Formally, the approximation Y_t converges weakly to the true solution X_t if \sup_{0 \leq t \leq T} \left| \mathbb{E}[f(X_t)] - \mathbb{E}[f(Y_t)] \right| \to 0 as the time step \Delta t \to 0, for any smooth test function f belonging to the space C_b^2 of twice continuously differentiable functions with bounded derivatives. Under the standard assumptions that the drift a and diffusion b coefficients of the SDE are globally Lipschitz continuous and satisfy linear growth conditions, and that the test function f \in C_b^2(\mathbb{R}^d), the Euler–Maruyama method exhibits a weak order of convergence of 1.0. This implies that the error satisfies \sup_{0 \leq t \leq T} \left| \mathbb{E}[f(X_t)] - \mathbb{E}[f(Y_t)] \right| \leq C \Delta t for some constant C > 0 independent of \Delta t. This order surpasses the strong convergence order of 0.5, as errors from individual increments tend to cancel out when averaging over paths to compute expectations. For linear SDEs, the Euler–Maruyama method achieves the full weak order of 1 exactly under the aforementioned conditions, with potential for higher accuracy when the noise is additive and f is a . In the nonlinear case, maintaining order 1 requires that f has bounded second derivatives to control the remainder terms in the Itô-Taylor expansion. Unlike strong convergence, which demands pathwise mean-square accuracy and is thus stricter, suffices for many applications focused on distributional properties. These results were rigorously established and refined in the comprehensive analysis by Kloeden and Platen.

Stability and Error Analysis

The Euler–Maruyama method preserves mean-square stability for linear equations of the form dX = \lambda X \, dt + \sigma X \, dW with \lambda < 0, provided the time step \Delta t satisfies the condition |1 + \lambda \Delta t|^2 + \sigma^2 \Delta t < 1. This inequality ensures that the second moment of the numerical solution decays to zero as time progresses, mirroring the behavior of the exact solution. For practical implementation, the condition highlights the restrictive role of the diffusion coefficient \sigma in limiting allowable step sizes when noise is prominent. The global strong error of the Euler–Maruyama method under global Lipschitz conditions on the drift and diffusion coefficients is bounded by O(\sqrt{\Delta t}), with an explicit estimate of the form \mathbb{E}[|X_T - \hat{X}_T|] \leq C \sqrt{\Delta t}, where C depends on the Lipschitz constant L, the final time T, and bounds on the coefficients. This bound arises from controlling the local truncation error accumulation over the interval [0, T], ensuring the method's reliability for sufficiently small \Delta t. Weak error analysis complements this by providing O(\Delta t) convergence for expectations, but stability considerations emphasize the strong error bound for pathwise approximations. In stiff stochastic differential equations, characterized by large negative eigenvalues in the drift (high Lipschitz constant L), the Euler–Maruyama method exhibits heightened sensitivity to the step size \Delta t, as the stability region shrinks proportionally to $1/L. This can lead to numerical instability or excessive variance growth if \Delta t is not sufficiently small, necessitating adaptive step-size strategies to dynamically adjust \Delta t based on local error estimates or stability indicators. Such adaptations maintain reliability while minimizing computational cost in systems with disparate time scales. For geometric Brownian motion dS = \mu S \, dt + \sigma S \, dW with negative drift \mu < 0, the method's preservation of variance decay requires \Delta t small enough to satisfy the mean-square stability condition, preventing numerical explosion in the second moment.

Examples and Implementations

Geometric Brownian Motion

The (GBM) serves as a fundamental example for applying the , particularly in modeling asset prices in finance, where the process exhibits multiplicative noise. The (SDE) governing GBM is given by dS_t = \mu S_t \, dt + \sigma S_t \, dW_t, with initial condition S_0 > 0, where \mu is the drift parameter, \sigma > 0 is the volatility, and W_t is a standard Wiener process. The exact solution to this SDE is S_t = S_0 \exp\left\{ \left(\mu - \frac{\sigma^2}{2}\right) t + \sigma W_t \right\}, which follows a log-normal distribution, enabling direct comparison with numerical approximations. The Euler–Maruyama method discretizes this over a time grid t_n = n \Delta t for n = 0, 1, \dots, N with T = N \Delta t, yielding the iterative scheme S_{n+1} = S_n \left(1 + \mu \Delta t + \sigma \Delta W_n \right), where \Delta W_n \sim \mathcal{N}(0, \Delta t) are independent Gaussian increments. This approximation preserves the multiplicative structure of the noise term, producing sample paths that approximate the true GBM trajectories. However, this approximation does not guarantee that sample paths remain positive, unlike the exact GBM solution, and negative values may occur, particularly with larger \Delta t. In simulations, multiple paths generated via Euler–Maruyama exhibit the characteristic and log-normal terminal distribution of GBM, with paths diverging from the due to the random increments. The expected strong error, measured as the from the exact solution over paths, scales as O(\sqrt{\Delta t}), reflecting the method's order-0.5 for such SDEs.

Practical Implementation Guidelines

The Euler–Maruyama method is implemented by discretizing the time [0, T] into N equal steps of size Δt = T/N, initializing the approximation Y_0 = X_0, and iteratively updating the scheme for n = 0 to N-1 as Y_{n+1} = Y_n + f(Y_n) Δt + g(Y_n) ΔW_n, where ΔW_n = √Δt ⋅ Z_n and Z_n is a standard normal N(0,1). To generate the random increments, standard generators are used, such as those producing Gaussian variates, ensuring the is accurate for small Δt. For multi-dimensional stochastic differential equations dX_t = μ(t, X_t) dt + σ(t, X_t) dW_t, where X_t ∈ ℝ^d and W_t is a p-dimensional , the method extends to vector form: Ŷ_{n+1} = Ŷ_n + μ(t_n, Ŷ_n) Δt + σ(t_n, Ŷ_n) ΔW_n, with ΔW_n ∈ ℝ^p generated as √Δt ⋅ Z_n and Z_n ∼ N(0, I_p). If the Wiener processes are correlated with Σ, correlated increments are obtained by applying to Σ, yielding L such that ΔW_n = L ⋅ (√Δt ⋅ Z_n) where Z_n consists of independent N(0,1) components; this ensures the correct correlation structure without altering the diffusion matrix σ. Pseudocode for a single path simulation in using is as follows:
python
import numpy as np

def euler_maruyama(f, g, X0, T, N, seed=None):
    if seed is not None:
        np.random.seed(seed)
    dt = T / N
    t = np.linspace(0, T, N+1)
    Y = np.zeros(N+1)
    Y[0] = X0
    for n in range(N):
        Z = np.random.normal(0, 1)
        dW = np.sqrt(dt) * Z
        Y[n+1] = Y[n] + f(Y[n], t[n]) * dt + g(Y[n], t[n]) * dW
    return t, Y
This implementation leverages for efficient random number generation and array operations; analogous code replaces np.random.normal with randn and uses vectorized loops where possible. In , libraries such as sde provide built-in solvers that implement the Euler–Maruyama scheme via functions like ItoEuler. For efficiency, vectorized operations in libraries like are essential to avoid explicit loops over time steps, achieving a typical of for simulating one path and O(N M) for M independent paths. To ensure reproducibility across runs, set a fixed before generating increments, as in the example. Additionally, monitor for numerical overflow, particularly when the diffusion coefficient g (or σ) is large, by checking intermediate values and using adaptive step sizes if explosions occur.

Applications and Extensions

Financial Modeling

The Euler–Maruyama method plays a central role in financial modeling by enabling the numerical simulation of asset price dynamics governed by stochastic differential equations (SDEs), particularly for derivative pricing under the Black-Scholes framework. In this context, Monte Carlo simulations approximate paths of geometric Brownian motion using the method's discretization scheme, where option prices are estimated by generating numerous paths and averaging the discounted payoffs at maturity. This approach is justified by the method's weak convergence properties, which ensure reliable computation of expected values for European-style options. In , the Euler–Maruyama method supports estimation by simulating scenarios for portfolio SDEs, allowing the quantification of potential losses at specified confidence levels through of tail distributions. For instance, the method discretizes multi-asset models to generate loss paths, from which the VaR is derived as the of simulated portfolio values. This application is particularly valuable for assessing systemic risks in correlated . The method's advantages in include its straightforward implementation for high-dimensional problems, such as simulating portfolios with dozens of assets, where closed-form solutions are infeasible. Extensions, such as truncated schemes for jump-diffusion processes, further enable modeling of sudden market shocks like those in Lévy-driven SDEs. During in volatile periods, the method has been employed with time steps of approximately \Delta t \approx 1/252 to match daily trading data frequencies. However, limitations arise from the method's convergence orders: its weak order of 1 suffices for expectation-based metrics like option prices or , but the strong order of 0.5 can introduce significant pathwise errors in pricing path-dependent options, such as barriers or Asians, where accurate reproduction is essential.

Biological Systems

The Euler–Maruyama method has been widely applied to simulate differential equations (SDEs) modeling biological processes, particularly those involving intrinsic noise from counts or environmental fluctuations. In gene expression modeling, the method facilitates numerical solutions for SDEs approximating bursty mRNA production, where bursts arise from intermittent promoter activity. A common formulation for mRNA concentration X_t is the SDE dX_t = (k - \gamma X_t) \, dt + \sqrt{k + \gamma X_t} \, dW_t, with production rate k, degradation rate \gamma, and Wiener process W_t, capturing the Poisson-like variability in transcription and degradation events. This diffusion approximation to the chemical master equation allows efficient simulation of noise-driven fluctuations in protein levels downstream of mRNA dynamics. For , the Euler–Maruyama method approximates solutions to SDEs incorporating , such as the logistic model dN_t = r N_t \left(1 - \frac{N_t}{K}\right) dt + \sigma N_t \, dW_t, where N_t is , r is intrinsic , K is , \sigma quantifies noise intensity, and W_t models random perturbations like resource variability. This approach reveals risks and persistence thresholds under forcing, outperforming deterministic models in low-population regimes. These applications leverage the method's ability to capture intrinsic noise in small biological systems, such as molecular counts in cells, where discrete stochastic effects dominate, and enable parameter inference from time-series data via likelihood-based methods. In , Higham (2008) demonstrated the Euler–Maruyama method's efficiency in integrating diffusion approximations for toggle switch models, combining it with chemical hybrids to approximate τ-leaping for faster simulations of bistable circuits. More recently, during the COVID-19 outbreaks, the method has supported stochastic extensions of SEIR (susceptible-exposed-infectious-recovered) models, simulating epidemic trajectories with demographic to assess outbreak variability and intervention impacts.

Machine Learning

In , particularly in generative modeling, the Euler–Maruyama method is extensively used to simulate differential equations underlying score-based models. These models, popularized since the early , frame generation as sampling from a reverse-time , where the forward process adds to and the reverse process denoises it to generate new samples. The method discretizes this reverse , often variance-preserving types, enabling efficient training and inference for tasks like image synthesis and text-to-image generation. For instance, in predictor-corrector samplers, Euler–Maruyama steps approximate the drift and terms, balancing computational speed with sample quality. As of 2025, extensions incorporate adaptive step sizes to mitigate errors in high-dimensional settings.

Comparisons to Other Methods

The Euler–Maruyama method, with its strong convergence order of 0.5 and weak order of 1.0, serves as a foundational explicit for simulating equations (SDEs), but it is often contrasted with the , which achieves a strong order of 1.0 by incorporating an additional Itô-Taylor expansion term that accounts for the integral of the coefficient's . This extra term in the Milstein requires computing the of the b(t, X_t), increasing complexity compared to the simpler Euler–Maruyama , while both methods maintain the same weak order of 1.0 for expectations of smooth functions. In comparison to stochastic for SDEs, the Euler–Maruyama scheme represents a explicit approach, whereas higher-order variants, such as weak order 2 Runge–Kutta schemes, offer improved accuracy by evaluating multiple stages per step to better approximate and integrals, though at a significantly higher computational cost due to additional evaluations of the coefficients. These Runge–Kutta extensions are particularly beneficial for problems demanding higher , but their derivative-free formulations still demand more resources than the baseline Euler–Maruyama method. A notable advantage of the Euler–Maruyama method arises in cases of additive noise, where the diffusion coefficient b(t) is constant and independent of the solution X_t; here, the scheme attains a strong order of 1.0, equivalent to the , without needing the additional correction term. Consequently, the Euler–Maruyama method is preferred for rapid simulations where is sufficient, such as estimations of expectations, while the is recommended for path-dependent applications requiring precise strong convergence, like barrier option pricing in . For stiff SDEs, where explicit schemes like Euler–Maruyama may suffer from restrictions necessitating tiny time steps, Rosenbrock-type methods provide semi-implicit alternatives that enhance through linearized Jacobian approximations, positioning Euler–Maruyama as the efficient baseline in most numerical libraries for non-stiff problems.
MethodStrong OrderWeak OrderKey Trade-off
Euler–Maruyama0.5 (1.0 for additive noise)1.0Simple, explicit; limited strong accuracy
Milstein1.01.0Requires ; higher strong order
Stochastic Runge–Kutta (higher-order)Varies (e.g., 1.0+)Varies (e.g., 2.0)More stages; better accuracy at higher cost
Rosenbrock-type (for stiff)0.5–1.01.0Improved ; semi-implicit

References

  1. [1]
    An Algorithmic Introduction to Numerical Simulation of Stochastic ...
    We describe in section 4 how the Euler–Maruyama method can be used to simulate an SDE. We introduce the concepts of strong and weak convergence in section 5 and ...
  2. [2]
    Continuous Markov processes and stochastic equations
    Maruyama, G. Continuous Markov processes and stochastic equations. Rend. Circ. Mat. Palermo 4, 48–90 (1955). https://doi.org/10.1007/BF02846028. Download ...
  3. [3]
    [PDF] Using the Euler-Maruyama Method for Finding a Solution to ...
    Jun 6, 2016 · Euler-Maruyama method is named after. Leonhard Euler and Gisiro Maruyama. II. PRELIMINARIES. In this section, we review some of the basic ...
  4. [4]
  5. [5]
    [PDF] An Algorithmic Introduction to Numerical Simulation of Stochastic ...
    We describe in section 4 how the Euler–Maruyama method can be used to simulate an SDE. We introduce the concepts of strong and weak convergence in sec- tion 5 ...
  6. [6]
    Numerical Treatment of Stochastic Differential Equations - jstor
    (2.4) x9j+j = x-i + a (tj, x-j)h + o-(tj, x-j)Awj. It was shown by G. Maruyama (1955) that the Euler method converges uniformly in. q.m. to xt.
  7. [7]
    Approximate Integration of Stochastic Differential Equations - SIAM.org
    On the existence of solutions of stochastic differential equations with integrable drift coefficient.
  8. [8]
    Numerical Solution of Stochastic Differential Equations - SpringerLink
    The aim of this book is to provide an accessible introduction to stochastic differ ential equations and their applications together with a systematic ...
  9. [9]
    [PDF] Strong convergence of the Euler–Maruyama method - Urbain Vaes
    Theorem 1 (Strong convergence of the Euler–Maruyama method). ... particular, it is possible to show that the order of convergence is 1/2 also when the strong.Missing: 1955 | Show results with:1955
  10. [10]
    Strong Convergence of Euler-Type Methods for Nonlinear ...
    In this work we prove strong convergence results under less restrictive conditions. First, we give a convergence result for Euler--Maruyama requiring only that ...
  11. [11]
  12. [12]
    None
    ### Parameters for Geometric Brownian Motion in Euler-Maruyama Simulation
  13. [13]
    [PDF] Applied Stochastic Differential Equations
    May 3, 2019 · The main motivation for the book is the application of stochastic differential equations (SDEs) in domains such as target tracking and medical ...
  14. [14]
    [PDF] Simulating Stochastic Differential Equations - Columbia University
    In these lecture notes we discuss the simulation of stochastic differential equations (SDEs), focusing mainly on the Euler scheme and some simple ...Missing: Maruyama | Show results with:Maruyama
  15. [15]
    Highly nonlinear model in finance and convergence of Monte Carlo ...
    The convergence result justifies clearly that the Monte Carlo simulations based on the Euler–Maruyama scheme can be used to compute the expected payoff of ...
  16. [16]
    Gradient-based optimisation of the conditional-value-at-risk using ...
    Dec 15, 2023 · The ODE is discretised using the Euler–Maruyama scheme, which reads ... Monte Carlo estimation of value-at-risk, conditional value-at-risk and ...
  17. [17]
    [PDF] Convergence of the Euler-Maruyama method for multidimensional ...
    Jan 22, 2019 · In this paper we prove strong convergence of order 1/4 − for arbitrarily small > 0 of the Euler-Maruyama method for multidimensional SDEs with ...
  18. [18]
    Jump Models with Delay—Option Pricing and Logarithmic Euler ...
    We propose a logarithmic Euler–Maruyama scheme to approximate the equation and prove that all the approximations remain positive and the rate of convergence of ...
  19. [19]
    [PDF] Stochastic Processes in Actuarial Surplus Modelling
    Aug 12, 2025 · The processes are discretised for simulation using the Euler-Maruyama method, as analytical solutions are impractical for complex surplus models ...
  20. [20]
    [PDF] Monte Carlo Methods - Mathematical Institute
    Barrier option. Some path-dependent options give only O(. √ h) weak convergence if the numerical payoff is not constructed carefully. A down-and-out call ...<|control11|><|separator|>
  21. [21]
    (PDF) Performance of Euler-Maruyama, 2-Stage SRK and 4-Stage ...
    Aug 6, 2025 · ... Euler-Maruyama, 2-stage SRK and 4-stage SRK in approximating the strong solutions of stochastic logistic model which describe the cell ...
  22. [22]
    Extinction-time for stochastic population models - ScienceDirect
    Our numerical simulations for the stochastic model (2) have been done using the Euler–Maruyama ... On the distribution of the time to extintion in the stochastic ...
  23. [23]
    Switching and Diffusion Models for Gene Regulation Networks
    Abstract. We analyze a hierarchy of three regimes for modeling gene regulation. The most complete model is a continuous time, discrete state space, Markov jump ...<|control11|><|separator|>
  24. [24]
    Stochastic modeling, analysis, and simulation of the COVID-19 ...
    This model incorporates the spread of COVID-19 impacted by social behaviors in the population and allows for projecting the number of infected, recovered, and ...
  25. [25]
  26. [26]
    [PDF] Computational solution of stochastic differential equations
    Runge-Kutta methods for numerical solution of stochastic differential equations. J Comput Appl Math 2002, 138:219–241. 19. Platen E, Wagner W. On a Taylor ...Missing: history | Show results with:history
  27. [27]
    [PDF] Lecture 5: Stochastic Runge–Kutta Methods
    Nov 25, 2014 · The simplest Runge–Kutta method is the (forward) Euler scheme. ... Stochastic Runge–Kutta methods are derivative-free methods for solving SDEs.
  28. [28]
    [PDF] Lecture 16 Numerical SDEs: Basics 1 Schemes
    From the stated result in Theorem 1.3, we know that the Euler-. Maruyama scheme is of weak order 1. Before we go to the rigorous proof, let us give a more.
  29. [29]
    Balanced Implicit Methods for Stiff Stochastic Systems - SIAM.org
    We present and analyze a new class of numerical methods for the solution of stiff stochastic differential equations (SDEs). These methods, called S-ROCK (for ...