Fact-checked by Grok 2 weeks ago

Crank–Nicolson method

The Crank–Nicolson method is a scheme for numerically solving time-dependent partial differential equations (PDEs), particularly parabolic ones such as the \frac{\partial u}{\partial t} = \alpha \frac{\partial^2 u}{\partial x^2}. Developed by John Crank and Phyllis Nicolson in 1947, it represents an implicit time-stepping approach that averages the spatial derivatives at the and time levels, yielding second-order accuracy in both and time (O((\Delta x)^2 + (\Delta t)^2)). In its standard formulation for the one-dimensional on a [0, L] \times [0, T] with Dirichlet conditions, the method discretizes using a uniform grid with spacing \Delta x and time with \Delta t, approximating the solution u(x,t) at grid points (x_i, t_n) as U_i^n. The scheme is given by: \frac{U_i^{n+1} - U_i^n}{\Delta t} = \frac{\alpha}{2} \left[ \delta_x^2 U_i^{n+1} + \delta_x^2 U_i^n \right], where \delta_x^2 U_i = \frac{U_{i+1} - 2U_i + U_{i-1}}{(\Delta x)^2} is the central second difference operator. This results in a tridiagonal at each time step, which is symmetric and positive definite, solvable efficiently via methods like the Thomas algorithm. A key advantage of the Crank–Nicolson method is its unconditional stability for the , meaning solutions remain bounded regardless of the ratio \frac{\alpha \Delta t}{(\Delta x)^2}, unlike explicit schemes that require restrictive time-step limits. This stability, combined with high accuracy, makes it suitable for long-time simulations without excessive computational cost from small steps. However, the implicit nature introduces mild oscillations near sharp gradients, which can be mitigated by extensions like upwinding or preconditioning. Beyond heat conduction, the method has broad applications in fields requiring solutions to diffusion-like PDEs. In , it is extensively used to price options under the by solving the corresponding parabolic PDE for asset prices. In , it simulates time evolution in the , enabling studies of dynamics and tunneling with unitary-preserving properties when adapted appropriately. Extensions to higher dimensions, nonlinear PDEs, and adaptive grids further enhance its versatility in and .

History and Development

Origins

The Crank–Nicolson method was developed in as a numerical technique for solving the heat conduction equation and related partial differential equations. This approach emerged in the context of early methods, which were foundational to for physical problems during the mid-20th century. The method's creation was motivated by wartime demands during for efficient solvers of partial differential equations in physics and engineering, particularly applications involving heat diffusion such as modeling wood combustion for incendiary devices developed by the UK's . These needs arose from efforts to improve sabotage tools and understand thermal processes in materials under extreme conditions. The scheme originated from Phyllis Nicolson's 1946 PhD thesis under , which developed the first model of solid fuel burning using the heat diffusion equation with an Arrhenius source term, in collaboration with Clement Bamford and David Malan for SOE applications. It was first published by John Crank and Phyllis Nicolson in the Mathematical Proceedings of the Cambridge Philosophical Society, titled "A practical method for numerical evaluation of solutions of partial differential equations of the heat-conduction type." The scheme emphasized implicit formulations to achieve unconditional , addressing limitations of explicit methods like the forward-time centered-space (, which suffered from severe restrictions. This innovation allowed for larger time steps in simulations without numerical instability, making it particularly valuable for practical computations in problems.

Key Contributors

John Crank (1916–2006) was a British mathematician renowned for his contributions to numerical methods for solving partial differential equations, particularly those modeling processes. After earning his M.Sc. from the in 1938, Crank joined the Fundamental Research Laboratory in 1945, where he focused on mathematical modeling of heat conduction and material throughout his career. His expertise in these areas stemmed from wartime research on industrial applications, leading to seminal work on approximations for parabolic equations. Phyllis Nicolson (1917–1968), also a British mathematician, collaborated closely with Crank during her postdoctoral work. Born Phyllis Lockett, she obtained her B.Sc. and M.Sc. from the in 1938 and 1939, respectively, followed by a Ph.D. in physics there in 1946 under . After her doctorate, Nicolson moved to , where she pursued research in ; she later held academic positions, including a lectureship in physics at the . Her work emphasized practical computational techniques for physical problems, building on her wartime experience working on and problems for the . The key collaboration between and Nicolson occurred at the in the mid-1940s, culminating in their 1947 joint paper, "A Practical Method for Numerical Evaluation of Solutions of Partial Differential Equations of the Heat-Conduction Type," published in the Proceedings of the Cambridge Philosophical Society. In this work, they introduced a scheme that demonstrated unconditional stability for parabolic partial differential equations, such as the , enabling reliable long-time simulations without restrictive time-step constraints. This innovation addressed limitations in explicit methods by averaging forward and backward differences, resulting in a tridiagonal system solvable at each step. Their method drew influence from earlier numerical techniques, including Lewis Fry Richardson's 1910 finite difference approaches for diffusion problems, which highlighted stability issues in explicit schemes, and Richard V. Southwell's relaxation methods for iterative solutions of elliptic and parabolic equations developed in the 1940s. These foundations, amid the era's advancements in during and after , informed and Nicolson's emphasis on and practicality.

Mathematical Principles

Finite Difference Methods for PDEs

Finite difference methods provide a foundational approach for numerically solving partial differential equations (PDEs) by approximating derivatives with discrete differences on a grid. For time-dependent PDEs, such as the one-dimensional heat equation u_t = a u_{xx}, where a > 0 is the diffusion coefficient, these methods discretize both space and time, leading to a system of ordinary differential equations or algebraic equations. The key distinction lies between explicit and implicit schemes, which differ in how they treat the time derivative and spatial derivatives at the current and next time levels. Explicit schemes compute the solution at the next time step directly from known values, offering simplicity but often requiring restrictive time step conditions for stability, while implicit schemes solve a system of equations involving unknown future values, providing greater stability at the cost of increased computational effort per step. A canonical explicit scheme is the forward-time central-space (FTCS) method, which approximates the time derivative with a forward difference and the spatial with a central difference. On a uniform grid with spatial step \Delta x and time step \Delta t, the scheme at interior point i and time level n is given by u_i^{n+1} = u_i^n + r (u_{i+1}^n - 2u_i^n + u_{i-1}^n), where r = a \Delta t / (\Delta x)^2 is the mesh . This method is conditionally stable, requiring the Courant-Friedrichs-Lewy (CFL) condition r \leq 1/2 to prevent oscillatory growth, as determined by . Violation of this condition leads to , limiting \Delta t relative to \Delta x and potentially increasing overall computational cost for fine spatial resolutions. In contrast, the is an implicit scheme that uses a backward difference for time and central differences for space, yielding u_i^{n+1} - r (u_{i+1}^{n+1} - 2u_i^{n+1} + u_{i-1}^{n+1}) = u_i^n. This formulation is unconditionally stable for any r > 0, allowing larger time steps without instability, though it is only first-order accurate in time (O(\Delta t)) compared to the second-order spatial accuracy (O((\Delta x)^2)). At each time step, the scheme results in a A \mathbf{u}^{n+1} = \mathbf{u}^n, where A is a tridiagonal matrix with 1 + 2r on the diagonal, -r on the sub- and super-diagonals, enabling efficient solution via the Thomas algorithm, a direct solver with O(N) complexity for N grid points. This efficiency makes implicit methods practical for stiff problems or long-time simulations in diffusion processes.

The Crank-Nicolson Scheme

The Crank–Nicolson method is a for solving time-dependent partial equations, particularly parabolic ones such as the . It approximates the time derivative using a forward difference: \frac{u^{n+1} - u^n}{\Delta t}, where u^n denotes the solution at time level n and \Delta t is the time step. The spatial derivatives are approximated by averaging the spatial operator L (e.g., L = a \partial_{xx} for a a > 0) evaluated at both the current and next time levels: \frac{1}{2} [L u^{n+1} + L u^n]. This averaging, derived from the trapezoidal rule for time integration, combines explicit and implicit treatments to achieve a balance between computational efficiency and numerical stability. The resulting scheme forms an implicit that must be solved at each time step: (I - \frac{\Delta t}{2} L) u^{n+1} = (I + \frac{\Delta t}{2} L) u^n, where I is the identity operator. For spatial using central differences, this yields a system that can be efficiently solved using methods like Thomas algorithm. The Crank–Nicolson method corresponds to the \theta-method with \theta = 1/2, where \theta weights the implicit contribution. In terms of accuracy, the scheme is second-order in time, with a local of O(\Delta t^2), due to the trapezoidal . When paired with second-order central differences , it also achieves O(\Delta x^2) spatial accuracy, making it suitable for problems requiring precise temporal . For linear parabolic partial differential equations, the Crank–Nicolson method exhibits unconditional , allowing arbitrary ratios of \Delta t / \Delta x^2 without , unlike explicit schemes that impose restrictive conditions. This property stems from the averaging that damps high-frequency error modes over time steps.

Derivation and Analysis

Discretization of the Heat Equation

The one-dimensional , which models the diffusion of through a medium, is given by \frac{\partial u}{\partial t} = \alpha \frac{\partial^2 u}{\partial x^2}, where u(x, t) represents the at position x and time t, and \alpha > 0 is the constant. This equation is typically supplemented with u(x, 0) = f(x) for $0 \leq x \leq L and conditions at x = 0 and x = L, such as Dirichlet conditions u(0, t) = g_0(t), u(L, t) = g_L(t), or Neumann conditions specifying the \frac{\partial u}{\partial x}. To apply the Crank–Nicolson method, discretize the spatial domain [0, L] into N+1 points x_i = i \Delta x for i = 0, 1, \dots, N, where \Delta x = L/N, and the time domain [0, T] into steps t_n = n \Delta t for n = 0, 1, \dots, M, where \Delta t = T/M. The approximation u_i^n \approx u(x_i, t_n) is used at interior points i = 1, \dots, N-1. The Crank–Nicolson averages the explicit and implicit central approximations for the second spatial . The time is approximated by the forward difference \frac{u_i^{n+1} - u_i^n}{\Delta t}, while the spatial at time level n + 1/2 uses the average \frac{\alpha}{2 (\Delta x)^2} \left[ (u_{i+1}^{n+1} - 2 u_i^{n+1} + u_{i-1}^{n+1}) + (u_{i+1}^n - 2 u_i^n + u_{i-1}^n) \right]. This yields the full scheme \frac{u_i^{n+1} - u_i^n}{\Delta t} = \frac{\alpha}{2 (\Delta x)^2} \left[ (u_{i+1}^{n+1} - 2 u_i^{n+1} + u_{i-1}^{n+1}) + (u_{i+1}^n - 2 u_i^n + u_{i-1}^n) \right], for i = 1, \dots, N-1 and n = 0, 1, \dots, M-1. Rearranging terms defines the r = \frac{\alpha \Delta t}{ (\Delta x)^2} and collects unknowns at the new time level on the left, resulting in the tridiagonal system -\frac{r}{2} u_{i-1}^{n+1} + (1 + r) u_i^{n+1} - \frac{r}{2} u_{i+1}^{n+1} = \frac{r}{2} u_{i-1}^n + (1 - r) u_i^n + \frac{r}{2} u_{i+1}^n. This linear system can be solved efficiently using methods like the Thomas algorithm due to its tridiagonal structure. Boundary conditions are incorporated directly into the system. For Dirichlet conditions, set u_0^{n+1} = g_0(t_{n+1}) and u_N^{n+1} = g_L(t_{n+1}), modifying the equations at i=1 and i=N-1 accordingly. For Neumann conditions, approximate the flux using centered differences, such as \frac{u_1^{n+1} - u_{-1}^{n+1}}{2 \Delta x} = h_0(t_{n+1}) at x=0, introducing ghost points u_{-1}^{n+1} eliminated by substitution into the i=1 equation. Similar adjustments apply at x=L.

Stability and Convergence

The Crank–Nicolson method is unconditionally for the , as demonstrated by . This analysis assumes a of the form u_j^n = g^n e^{i \xi j \Delta x}, leading to the amplification factor g(\xi) = \frac{1 - 2 r \sin^2(\xi \Delta x /2)}{1 + 2 r \sin^2(\xi \Delta x /2)}, where r = \alpha \Delta t / \Delta x^2 is the diffusion number and \alpha is the . For all r > 0 and wave numbers \xi, |g(\xi)| \leq 1, with equality only when \sin^2(\xi \Delta x /2) = 0; thus, no mode grows unbounded, confirming unconditional stability independent of the time step size. The method is consistent with the , achieving a local of O(\Delta t^2 + \Delta x^2). This order arises from Taylor expansions around the in time and centered differences in space, which approximate the time to second order and the spatial to second order. Convergence follows from the Lax equivalence theorem, which states that for linear problems, a consistent scheme converges it is . Given the second-order consistency and unconditional stability, the Crank–Nicolson method exhibits second-order global error convergence under sufficiently smooth solutions. Despite its stability, the method can produce non-physical oscillations in numerical solutions with sharp gradients or discontinuities, as high-frequency modes are preserved without damping.

Basic Applications

One-Dimensional Diffusion

The Crank–Nicolson method finds frequent application in simulating one-dimensional diffusion processes, such as heat conduction in a thin rod, subject to homogeneous Dirichlet boundary conditions and an initial Gaussian temperature distribution. Consider the heat equation u_t = \alpha u_{xx} defined on the spatial domain $0 \leq x \leq L, where \alpha is the thermal diffusivity, with boundary conditions u(0, t) = u(L, t) = 0 for t > 0, and initial condition u(x, 0) = A \exp\left( -\frac{(x - x_0)^2}{2\sigma^2} \right), representing a centered Gaussian profile with amplitude A, mean x_0 = L/2, and standard deviation \sigma. This setup models scenarios like localized heat sources in insulated rods with fixed endpoints at ambient temperature. The discretized form of the scheme, derived from averaging the explicit and implicit finite difference approximations, yields a linear system in matrix notation: \mathbf{A} \mathbf{u}^{n+1} = \mathbf{B} \mathbf{u}^n, where \mathbf{u}^n is the vector of solution values at time level n on the spatial grid with N interior points, \Delta x = L / (N+1), and r = \alpha \Delta t / (\Delta x)^2. The matrices \mathbf{A} and \mathbf{B} are tridiagonal, with \mathbf{A} having diagonal elements $1 + r/2, sub- and super-diagonal elements -r/2, and \mathbf{B} having diagonal elements $1 - r/2, sub- and super-diagonal elements r/2; boundary conditions are enforced by setting endpoint values to zero. This system is solved iteratively at each time step using the Thomas algorithm (tridiagonal matrix algorithm), which exploits the banded structure for efficient forward and backward substitution. Implementation involves discretizing the and time, initializing the , and advancing in time via the tridiagonal solve. for a basic solver is as follows:
Set parameters: L, alpha, T_final, sigma, A, x0 = L/2
N = number of spatial intervals (e.g., 100)
dx = L / (N + 1)
x = [dx, 2*dx, ..., N*dx]
dt = 0.5 * dx^2 / alpha  # Chosen for accuracy
Nt = round(T_final / dt)
r = alpha * dt / dx^2

Initialize u[1:N] = A * exp( -(x - x0)^2 / (2*sigma^2) )

For n = 1 to Nt:
    # Build tridiagonal A: diag(1 + r/2), off-diag(-r/2)
    # Build RHS: B * u = diag(1 - r/2)*u + off-diag(r/2)*u (with boundaries 0)
    rhs[1] = (1 - r/2) * u[1] + (r/2) * u[2]
    for i = 2 to N-1:
        rhs[i] = (r/2) * u[i-1] + (1 - r/2) * u[i] + (r/2) * u[i+1]
    rhs[N] = (r/2) * u[N-1] + (1 - r/2) * u[N]
    
    # Thomas algorithm solve A * u = rhs (update in place)
    # Forward elimination:
    c[1] = -r/2 / (1 + r/2)
    d[1] = rhs[1] / (1 + r/2)
    for i = 2 to N-1:
        denom = 1 + r/2 + (r/2) * c[i-1]
        c[i] = - (r/2) / denom
        d[i] = (rhs[i] + (r/2) * d[i-1]) / denom
    denom_N = 1 + r/2 + (r/2) * c[N-1]
    d[N] = (rhs[N] + (r/2) * d[N-1]) / denom_N
    
    # Back substitution:
    u[N] = d[N]
    for i = N-1 downto 1:
        u[i] = d[i] - c[i] * u[i+1]

Output u at desired times for plotting
This loop runs for Nt steps, with each iteration requiring O(N) operations due to the tridiagonal solve. Numerical results demonstrate the method's accuracy and stability. Plots of u(x, t) at multiple time levels reveal diffusive spreading, while surface plots over x-t illustrate the temporal decay. Comparison to the exact solution for the infinite domain, u(x, t) = \frac{\sigma}{\sqrt{\sigma^2 + 4\alpha t}} \exp\left( -\frac{(x - x_0)^2}{2(\sigma^2 + 4\alpha t)} \right) (which can be related to the error function via integration), shows close agreement in the domain interior when L is sufficiently large relative to the diffusion length \sqrt{4\alpha t}, consistent with the scheme's second-order accuracy. In cases approximating semi-infinite conditions (e.g., large L, one-sided Gaussian near x=0 with u(L,t) \approx 0), the numerical profile aligns well with the complementary error function solution u(x,t) = \erfc\left( \frac{x}{\sqrt{4\alpha t}} \right) for step-like boundary inputs, validating the method's fidelity. The overall computational cost is O(N Nt), linear in the grid size per time step, making it efficient for moderate N up to thousands.

Two-Dimensional Diffusion

The two-dimensional describes the of heat in a , given by \frac{\partial u}{\partial t} = a \left( \frac{\partial^2 u}{\partial x^2} + \frac{\partial^2 u}{\partial y^2} \right), where u(x,y,t) is the , a > 0 is the , and the domain is typically a or square. The Crank-Nicolson method extends to this equation by applying the average of explicit and implicit finite difference approximations in time, using central differences in space on a uniform with spacing \Delta x = \Delta y = h and time step \Delta t = k. This results in a fully implicit at each time step, yielding a sparse linear of equations whose is block tridiagonal, with each block being a of size (M-1) \times (M-1) for an M \times M interior . For efficiency in solving this system of O(N^2) unknowns, where N \approx M^2, direct methods like banded (costing O(N^2)) or iterative methods such as Gauss-Seidel can be employed, leveraging the sparsity. An Alternating Direction Implicit (ADI) variant, originally developed by Peaceman and Rachford, splits the problem into sequential one-dimensional solves along x and y directions, reducing computational cost while maintaining second-order accuracy and unconditional stability. Unlike the explicit method, which requires a CFL-like stability condition \mu = a k / h^2 \leq 1/4 to prevent oscillations, the Crank-Nicolson is unconditionally , allowing larger time steps. A representative example is on a domain with insulated () boundary conditions \partial u / \partial n = 0 on all edges, initialized with a central hot spot such as a Gaussian u(x,y,0) = \exp(-( (x-0.5)^2 + (y-0.5)^2 ) / 0.01 ). The method accurately captures the symmetric spreading and decay of the heat profile over time.

Advanced Applications

Advection-Diffusion Equations

The one-dimensional - equation, which models transport processes combining convective flow and diffusive spreading, is given by \frac{\partial u}{\partial t} + v \frac{\partial u}{\partial x} = a \frac{\partial^2 u}{\partial x^2}, where u(x,t) represents the concentration or quantity of interest, v > 0 is the constant advection velocity, and a > 0 is the . This equation arises in environmental modeling, such as pollutant dispersion in flowing fluids. To apply the Crank-Nicolson method, spatial discretization employs central differences for the term \frac{\partial^2 u}{\partial x^2} to maintain second-order accuracy, while the term \frac{\partial u}{\partial x} uses an upwind-biased , such as \frac{u_i^n - u_{i-1}^n}{\Delta x} for positive v, to prevent non-physical oscillations in convection-dominated regimes. The time-stepping averages the right-hand side ( and contributions) between time levels n and n+1, yielding an implicit scheme: \frac{u_i^{n+1} - u_i^n}{\Delta t} + \frac{v}{2} \left[ \left( \frac{u_i^{n+1} - u_{i-1}^{n+1}}{\Delta x} \right) + \left( \frac{u_i^n - u_{i-1}^n}{\Delta x} \right) \right] = \frac{a}{2 \Delta x^2} \left[ (u_{i-1}^{n+1} - 2u_i^{n+1} + u_{i+1}^{n+1}) + (u_{i-1}^n - 2u_i^n + u_{i+1}^n) \right], where u_i^k \approx u(i \Delta x, k \Delta t). This formulation ensures second-order accuracy in time, first-order spatial accuracy for the advection term, and second-order spatial accuracy for the diffusion term while leveraging the implicit nature for stability. In multi-channel systems, such as river networks with steady flow, the scheme is implemented separately on each channel segment, with coupling at junctions enforced by through flux continuity: the outgoing advective-diffusive flux from upstream channels equals the incoming flux to downstream ones. For instance, in simulating solute across connected river branches, this approach accurately captures flow partitioning while maintaining numerical efficiency. Stability of the scheme remains unconditional with respect to the time step \Delta t due to its implicit averaging, but spatial accuracy and monotonicity depend on the cell . For high \mathrm{Pe} > 2, the upwind bias for mitigates dispersive errors and ensures oscillation-free solutions, though coarser grids may require additional stabilization to preserve positivity. In steady-flow examples, such as in branched waterways, the method demonstrates robust performance, with errors below 1% for moderate \mathrm{Pe} when \Delta x is refined appropriately.

Nonlinear Problems

The Crank–Nicolson method, originally formulated for linear partial differential equations, can be extended to nonlinear problems by incorporating iterative linearization techniques at each time step. A prototypical example is the nonlinear u_t = \nabla \cdot (k(u) \nabla u), where the k(u) depends on the solution itself, modeling phenomena such as temperature-dependent . The scheme discretizes the time derivative using the centered difference at the midpoint t^{n+1/2}, averaging the nonlinear flux terms between time levels n and n+1, resulting in an implicit nonlinear algebraic system that requires iterative solution. To resolve the nonlinearity, Picard iteration or is commonly employed to linearize the system around iterates from the previous step. In Picard iteration, the nonlinear operator is frozen at the current approximation u^k, yielding a fixed-point formulation such as u^{k+1} = \left( I - \frac{\Delta t}{2} L(u^k) \right)^{-1} \left( u^n + \frac{\Delta t}{2} L(u^k) u^n \right), where L(u^k) represents the discrete divergence of the flux k(u^k) \nabla u, and the inverse involves solving a akin to the linear case (e.g., via tridiagonal matrices in one dimension). , in contrast, linearizes via Taylor expansion, incorporating the of the nonlinear residual for faster quadratic convergence when the initial guess is sufficiently close. Each iteration solves a linear Crank–Nicolson-like system, preserving the scheme's second-order accuracy in time provided the linear solves are exact. For quadratic nonlinearities, such as those arising in coefficients or terms, convergence of these iterations typically requires 3–5 steps per time step, with often needing fewer due to its quadratic rate, while may benefit from under-relaxation to accelerate linear . The process continues until a tolerance is met, ensuring the is resolved to machine precision without degrading the overall temporal order. A representative application is the reaction-diffusion equation with logistic growth, exemplified by u_t = \nabla \cdot (k(u) \nabla u) + u(1 - u), which models or front propagation. Here, the Crank–Nicolson scheme discretizes the diffusion implicitly while treating the logistic reaction term via iteration, often using to solve the resulting nonlinear tridiagonal (or pentadiagonal in higher-order spatial schemes) system at each step. Convergence is monitored through residual norms, such as the Euclidean norm of the difference between successive iterates \| u^{k+1} - u^k \| < \epsilon, typically achieving stability and second-order accuracy for traveling wave solutions.

Specialized Uses

Financial Mathematics

In financial mathematics, the Crank–Nicolson method is widely applied to solve the (PDE) for pricing options and other derivatives. The governs the value V(S, t) of a European option on an underlying asset with price S at time t, volatility \sigma, and risk-free rate r: \frac{\partial V}{\partial t} + \frac{1}{2} \sigma^2 S^2 \frac{\partial^2 V}{\partial S^2} + r S \frac{\partial V}{\partial S} - r V = 0, with terminal condition V(S, T) = \max(S - K, 0) for a call option with strike K and maturity T. To facilitate numerical solution via Crank–Nicolson, a logarithmic transformation x = \log(S/K) is commonly employed, along with a time reversal \tau = T - t, converting the PDE into a convection-diffusion equation resembling the : \frac{\partial V}{\partial \tau} = \frac{\sigma^2}{2} \frac{\partial^2 V}{\partial x^2} + \left(r - \frac{\sigma^2}{2}\right) \frac{\partial V}{\partial x} - r V, where the diffusion coefficient is \sigma^2 / 2. This transformation linearizes the coefficients and maps the semi-infinite domain S > 0 to x \in (-\infty, \infty), simplifying boundary handling while preserving the parabolic structure suitable for implicit schemes. The Crank–Nicolson scheme discretizes this transformed PDE on a , averaging explicit and implicit Euler steps for second-order accuracy in both and time, yielding an unconditionally stable tridiagonal system solvable via Thomas algorithm. In practice, non-uniform grids in the original S-variable are preferred for computational efficiency, with finer spacing near the K to capture the option's payoff discontinuity and reduce truncation errors without excessive nodes across the wide domain. For American options, which allow early exercise, the method incorporates penalty terms in the PDE to approximate the free boundary, enforcing the constraint V \geq \max(S - K, 0) iteratively while maintaining stability. As an example, consider pricing a European call option with S_0 = 100, K = 100, \sigma = 0.2, r = 0.05, T = 1. The exact value is 10.4506, and the Crank–Nicolson method converges to this value demonstrating second-order accuracy O(\Delta t^2 + \Delta x^2). Key advantages of Crank–Nicolson in this context include its ability to handle variable coefficients inherent to the (e.g., S-dependent diffusion and convection) without instability, alongside second-order accuracy that outperforms first-order schemes like fully implicit methods for smooth solutions. However, care is needed to mitigate oscillations near discontinuities, often via smoothing or fitted schemes.

Quantum Mechanics

The Crank–Nicolson method finds significant application in for numerically solving the time-dependent , which governs the evolution of a quantum system's \psi(x, t). In one dimension, the equation takes the form i \hbar \frac{\partial \psi}{\partial t} = -\frac{\hbar^2}{2m} \frac{\partial^2 \psi}{\partial x^2} + V(x) \psi, where \hbar is the reduced Planck's constant, m is the particle mass, and V(x) is the . This describes unitary evolution, requiring numerical schemes that preserve key quantum properties such as the of the , \int |\psi|^2 \, dx = 1. For the free-particle case where V(x) = 0, the Hamiltonian H = -\frac{\hbar^2}{2m} \frac{\partial^2}{\partial x^2} is time-independent, and the Crank–Nicolson scheme emerges as the Cayley transform of the exact unitary propagator e^{-i H \Delta t / \hbar}. The discretized update is \left( I + i \frac{\Delta t}{2 \hbar} H \right) \psi^{n+1} = \left( I - i \frac{\Delta t}{2 \hbar} H \right) \psi^n, where \Delta t is the time step and I is the identity operator. This formulation ensures unitarity of the evolution operator, as the Cayley transform maps skew-Hermitian operators to unitary ones, thereby conserving the L^2-norm of \psi exactly in the discrete setting. The method is second-order accurate in time and unconditionally stable, exhibiting minimal numerical dissipation compared to explicit schemes. When a non-zero potential V(x) is present, the kinetic and potential operators do not commute, necessitating a split-operator approach based on the Trotter approximation to factor the evolution operator. The Hamiltonian splits as H = T + V, with kinetic operator T = -\frac{\hbar^2}{2m} \frac{\partial^2}{\partial x^2} and potential V(x). The second-order Strang splitting approximates the time step as e^{-i H \Delta t / \hbar} \approx e^{-i T \Delta t / (2\hbar)} e^{-i V \Delta t / \hbar} e^{-i T \Delta t / (2\hbar)}, where the kinetic evolution is efficiently computed in momentum space via fast Fourier transforms, and the potential evolution is diagonal in position space. Crank–Nicolson can be integrated into this framework for the kinetic part to maintain unitarity and accuracy. This hybrid approach preserves the wave function norm and enables simulations of complex quantum dynamics. In practice, the method excels in benchmark problems like the particle in an infinite square well or the , where initial Gaussian wave packets evolve without artificial norm leakage or excessive phase errors. For the with V(x) = \frac{1}{2} m \omega^2 x^2, the scheme accurately captures periodic revivals and spreading, demonstrating second-order and long-term over many periods. These properties make Crank–Nicolson a cornerstone for quantum simulations, particularly in linear systems. Nonlinear extensions exist for more advanced quantum models but lie beyond the standard linear application here.

Modern Extensions

Fractional and Singularly Perturbed Problems

The Crank–Nicolson method has been adapted for time-fractional equations of the form {}^C D_t^\alpha u(x,t) = a u_{xx}(x,t), where {}^C D_t^\alpha denotes the Caputo fractional derivative of order $0 < \alpha < 1. In these schemes, the L1 approximation discretizes the fractional time derivative, providing an accuracy of O(\Delta t^{2-\alpha}) for sufficiently smooth solutions, while the Crank–Nicolson averaging is applied to the spatial second derivative for second-order accuracy O(\Delta x^2). This combination preserves the unconditional stability of the classical method and effectively captures the subdiffusive behavior modeled by the fractional order. For singularly perturbed problems, where a small diffusion parameter \epsilon > 0 introduces sharp boundary layers, standard Crank–Nicolson discretizations can exhibit oscillations unless modified. Adaptations incorporate upwind schemes or non-standard approximations in the spatial direction to stabilize the solution across layers, often on uniform . These modifications maintain the method's second-order temporal accuracy while ensuring robustness to \epsilon, particularly in convection-dominated regimes. Fitted approaches have also been integrated to refine near layers without excessive computational cost. Post-2020 has advanced these adaptations with uniformly convergent schemes for time-fractional singularly delay partial equations, achieving estimates of O(\Delta t^2 + \Delta x) that are of the parameter. These schemes employ –Nicolson for the fractional temporal component alongside specialized spatial discretizations, demonstrating practical efficiency in numerical experiments for delay-involved subdiffusion models. A notable application arises in financial mathematics, where the Crank–Nicolson method solves the time-fractional to model in European option pricing. These extensions accurately simulate heavy-tailed asset returns and long-memory effects. Recent advances as of 2025 include Crank–Nicolson finite difference schemes for the chiral , enabling simulations of nonlinear optical phenomena with second-order accuracy, and for the three-dimensional Allen-Cahn equation, providing error estimates in phase-field models of material science. Additionally, fast element-free Galerkin methods combined with Crank–Nicolson have been developed for the , achieving high-order spatial accuracy for studies.

Two-Grid and Iterative Methods

The two-grid Crank–Nicolson (CN) method enhances the efficiency of solving nonlinear parabolic partial differential equations (PDEs) by decoupling the into a coarse for handling the nonlinear terms and a fine for applying . On the coarse of size H, the full nonlinear CN scheme is solved, which captures the dominant behavior at reduced resolution. Subsequently, on the finer of size h (with h = O(H^2)), only linear systems derived from the linearized CN scheme are solved using the coarse solution as an initial guess, yielding optimal estimates while minimizing iterations. This approach reduces the computational cost per time step from O(N^3) for direct nonlinear solves on the fine to O(N^{3/2}) in two dimensions, where N \propto 1/h^2 represents the number of . In the context of convection-diffusion equations, a second-order characteristic mixed combines Raviart–Thomas elements for with piecewise constant elements for the scalar variable. This formulation preserves mass conservation and provides error estimates of O(h^2 + \Delta t^2) in the L^2-norm for the scalar solution under suitable regularity assumptions, enabling robust simulations of advection-dominated flows without excessive oscillations. Post-2020 research has advanced two-grid CN techniques for higher-order problems, including reduced-dimension variants that leverage to further compress solution coefficient vectors while maintaining accuracy. For instance, a reduced-dimension two-grid mixed finite element CN algorithm for the fourth-order nonlinear Rosenau decomposes the biharmonic term into coupled second-order systems and applies POD to shrink the effective dimensionality from O(10^6) to O(10) unknowns, achieving error bounds of O(\Delta t^2 + h^{r}) (with r \geq 2) and substantial savings in numerical benchmarks. Similarly, the two-grid CN mixed finite element (TGCNMFE) method for the convective FitzHugh–Nagumo system—a nonlinear reaction-diffusion model with —employs lowest-order mixed s and proves unconditional with errors O(\Delta t^2 + h^2 + H^2), demonstrating superior efficiency over single-grid Newton solvers in 2D simulations. An illustrative application involves accelerating the iterative Newton method for nonlinear CN discretizations using two-grid preconditioning: the coarse-grid nonlinear solve provides a warm start for fine-grid Newton iterations, converging in fewer steps (typically 3–5) compared to standard Newton (10+ iterations), as verified for semilinear parabolic interface problems. For implementation, multigrid cycles—such as V-cycles or W-cycles—integrate seamlessly with two-grid CN by using the coarse solution for prolongation and restriction operators, accelerating overall convergence to machine precision in O(N) work per cycle for linear systems arising post-linearization.

References

  1. [1]
    A practical method for numerical evaluation of solutions of partial ...
    This paper is concerned with methods of evaluating numerical solutions of the non-linear partial differential equation.
  2. [2]
    [PDF] Crank Nicolson Scheme for the Heat Equation
    The goal of this section is to derive a 2-level scheme for the heat equation which has no stability requirement and is second order in both space and time.
  3. [3]
    [PDF] Crank-Nicolson time-marching - People
    Crank and P. Nicolson. A practical method for numerical integra- tion of solutions of partial differential equations of heat-conduction type. In Proceedings ...Missing: original | Show results with:original
  4. [4]
    Numerical Simulation of the Time-Dependent Schrodinger Equation ...
    Oct 14, 2024 · This study presents a numerical simulation of a quantum electron confined in a 10 nm potential well, using the Crank-Nicolson numerical technique to solve the ...
  5. [5]
  6. [6]
    Historical investigation into the first burning model of a solid fuel and ...
    Oct 9, 2025 · Crank and Nicolson were primarily concerned with the solving of wood burning problem and hence the new numerical, the Crank–Nicolson (CN) method ...Missing: motivation | Show results with:motivation
  7. [7]
    [PDF] Historical investigation into the first burning model of a solid fuel and ...
    Sep 15, 2025 · The first burning model of a solid fuel, the heat diffusion equation with an Arrhenius-type source was solved by Phyllis. Nicolson during the ...
  8. [8]
    John Crank (1916 - 2006) - Biography - MacTutor
    Crank and Nicolson's method, which is numerically stable, requires the solution of a very simple system of linear equations (a tridiagonal system) at each time ...
  9. [9]
    Phyllis Nicolson (1917 - 1968) - Biography - MacTutor
    Phyllis Nicolson was an English mathematician best known for her work in numerical analysis. ... Crank and Nicolson's method, which is numerically stable, ...
  10. [10]
    Finite Difference Methods for Ordinary and Partial Differential ...
    This book introduces finite difference methods for both ordinary differential equations (ODEs) and partial differential equations (PDEs)
  11. [11]
    None
    Below is a merged summary of the Crank-Nicolson method based on the provided segments from "Numerical Analysis 2e" by Sauer. Since the original content does not explicitly detail the Crank-Nicolson method in many segments, I have synthesized the information by combining the most consistent and detailed descriptions, supplemented with standard knowledge where necessary (noted where applicable). To retain all information in a dense and organized manner, I will use a table in CSV format for key details, followed by a narrative summary that integrates additional context and useful URLs.
  12. [12]
    [PDF] Numerical Methods for Partial Differential Equations - Seongjai Kim
    Dec 11, 2023 · • An adaptive method is a numerical method which uses a variable step ... – The Crank-Nicolson method can be viewed as an explicit method in.
  13. [13]
    [PDF] Lecture notes on Numerical Analysis of Partial Differential Equations
    is always done in numerical analysis: we consider not a single discretization, but a sequence ... resulting fully discrete scheme is called the Crank–Nicolson ...
  14. [14]
    Numerical Solution of Partial Differential Equations - G. D. Smith
    Free delivery 25-day returnsThis authoritative study covers the standard finite difference methods of parabolic, hyperbolic, and elliptic equations.
  15. [15]
    The Crank-Nicholson scheme - Richard Fitzpatrick
    The well-known Crank-Nicholson implicit method for solving the diffusion equation involves taking the average of the right-hand side between the beginning and ...
  16. [16]
    [PDF] 11. Finite difference methods for the heat equation
    For the Crank-Nicholson method, one can show that the local truncation error is O(k2)+O(h2).
  17. [17]
    [PDF] Chapter 4. Accuracy, Stability, and Convergence - People
    Crank-Nicolson formulas for ut =ux, ut =uxx, ut =iuxx. In (3.5.35) we found the amplification factor for the Crank-Nicolson formula to be g( ) = 1;2 k h2 ...
  18. [18]
    [PDF] Eigenvalue Dependence of Numerical Oscillations in Parabolic ...
    The Crank-Nicolson method can be susceptible to these discontinuities, which propagate oscillations. (stable and unstable) in a normally consistent and ...
  19. [19]
    [PDF] 5.4 The Heat Equation and Convection-Diffusion
    0 t in (5). Example 1 Suppose the initial function is a bell-shaped Gaussian u(x, 0) = e−x2/2 .
  20. [20]
    [PDF] Crank-Nicolson Scheme for the 1D Heat Equation ME 448/548 in a ...
    3. Stability: The Crank-Nicolson method is unconditionally stable for the heat equation. The benefit of stability comes at a cost of increased complexity ...
  21. [21]
    [PDF] NOV 071995 Scru - DSpace@MIT
    above equations a semi-infinite medium ... Erfc(3.2)) which could be taken as zero for ... Difference method and for the Crank-Nicolson method, respectively.
  22. [22]
    CS267: Notes for Lecture 13, Feb 27, 1996 - People @EECS
    Feb 27, 1996 · Let us show how to solve the 1-dimensional heat equation. For simplicity, we take C=1. Since the problem is continuous, we need to discretize it to reduce it ...
  23. [23]
    (PDF) An Efficient Explicit Scheme for Solving the 2D Heat Equation ...
    Aug 10, 2025 · This paper presents a comprehensive numerical study of the two-dimensional time-dependent heat conduction equation using the Forward Time ...<|control11|><|separator|>
  24. [24]
    Numerical solution of unsteady advection dispersion equation ...
    A large Peclet number Pe > 100 shows the advection dominates dispersion [36]. ... A new difference scheme with high accuracy and absolute stability for solving ...
  25. [25]
    (PDF) Modeling and Numerical Simulation of River Pollution Using ...
    Aug 8, 2025 · In the present study we have applied diffusion – reaction equation to describe the dynamics of river pollution and drawn numerical solution ...
  26. [26]
    Advection Term - an overview | ScienceDirect Topics
    We see that the Crank Nicolson scheme performs well in solving the advection ... Here, the upwind differencing becomes effective only when the local Peclet number ...
  27. [27]
    A modified Crank-Nicolson scheme with incremental unknowns for ...
    Aug 6, 2025 · A modified Crank–Nicolson scheme based on one-sided difference approximations is proposed for solving time-dependent convection dominated ...
  28. [28]
    [PDF] Numerical Solution of One-dimensional Advection-diffusion ...
    Regarding. Fig. 1, the Crank-Nicolson scheme establishes results closer to the analytical solution. BTCS scheme provides suitable results too, and BTBSCS ...
  29. [29]
    [PDF] Diffusive Wave Models for Operational Forecasting of Channel ...
    Oct 7, 2022 · The second model uses the CN scheme over three temporal nodes and is referred to as Crank– Nicolson over. Time (CNT). Both models can properly ...
  30. [30]
    Lagrangian modeling of advection‐diffusion transport in open ...
    Dec 9, 2009 · The method employs a Crank-Nicolson type scheme that can be used in explicit, semi-implicit, or fully implicit modes.
  31. [31]
    (PDF) Assessment of some numerical methods for estimating the ...
    Aug 7, 2025 · For Pe > 5 Crank–Nicolson and MacCormack gave slightly overestimated dispersion coefficients whilst the other methods gave significantly ...
  32. [32]
    [PDF] Solving nonlinear ODE and PDE problems - Hans Petter Langtangen
    To solve the nonlinear algebraic equations we need to apply the Picard iteration method or Newton's method to the variational form directly, as shown next ...
  33. [33]
    [PDF] a crank nicholson method with five points to solve fisher's equation
    In Fisher's equation, the mechanism of logistic growth and linear diffusion are com- bined in order to model the spreading and proliferation of population ...
  34. [34]
    [PDF] 5 Solving the heat conduction and Black-Scholes equations
    This chapter is concerned with the nature of these equations, focusing attention on the heat conduction equation and then extending to the Black-Scholes ...
  35. [35]
    [PDF] A Critique of the Crank Nicolson Scheme Strengths and ...
    The Crank Nicolson method has become one of the most popular finite difference schemes for approximating the solution of the Black. Scholes equation and its ...
  36. [36]
    Option Pricing Using The Crank -Nicolson Finite Difference Method
    This tutorial discusses the specifics of the Crank-Nicolson finite difference method as it is applied to option pricing.
  37. [37]
    Efficient explicit numerical solutions of the time-dependent Schr ...
    Feb 4, 2022 · The advantage of the generalized Crank-Nicolson approach is that it preserves unitarity and is unconditionally stable.
  38. [38]
    Numerical solutions of the time-dependent Schrödinger equation in ...
    Feb 23, 2017 · The generalized Crank-Nicolson method is employed to obtain numerical solutions of the two-dimensional time-dependent Schrödinger equation.Article Text · INTRODUCTION · ACCURATE TIME... · DISCUSSION
  39. [39]
    A Crank-Nicolson Approximation for the time Fractional...
    A Crank-Nicolson Approximation for the time Fractional Burgers Equation. M ... Caputo sense. The L1 type discretization formula is going to be applied ...
  40. [40]
    Full article: Crank–Nicolson method for solving time-fractional ...
    The proposed numerical scheme consists of the Crank–Nicolson method for discretizing the time-fractional derivative, which is given in Caputo sense and the non- ...
  41. [41]
    Practical finite difference method for solving multi-dimensional black ...
    We use explict and Crank-Nicolson scheme to explore the one-asset fractional Black-Scholes model. •. We use OSM method to establish semi-implicit scheme to ...
  42. [42]
    Two-grid finite element methods combined with Crank-Nicolson ...
    Aug 31, 2018 · In this paper, we present a second-order accurate Crank-Nicolson scheme for the two-grid finite element methods of the nonlinear Sobolev ...
  43. [43]
    A Second Order Characteristic Mixed Finite Element Method for ...
    The stability is proved and the L2-norm error estimates are derived for both the scalar unknown variable and its flux. The scheme is of second order accuracy in ...
  44. [44]
    Reduced-Dimension Study of Two-Grid Mixed Finite Element Crank ...
    May 27, 2025 · This paper mainly studies the reduction dimensionality of two-grid mixed finite element (TGMFE) Crank-Nicolson (CN) (TGMEFCN) algorithm of the fourth-order ...
  45. [45]
    A New Two-grid Crank-Nicolson Mixed Finite Element Method for ...
    In this article, we mainly develop a new two-grid Crank-Nicolson (CN) mixed finite element (FE) (TGCNMFE) method for the convective FitzHugh-Nagumo equation ...
  46. [46]
    A two-grid immersed finite element method with the Crank-Nicolson ...
    In this paper, we propose and analyze a new two-grid partially penalized immersed finite element method for solving the semilinear parabolic interface ...
  47. [47]
    [PDF] Crank-Nicolson Method of a Two-Grid Finite Volume Element ...
    Firstly, we choose different time and space steps and use Newton iteration method to deal with the nonlinear system. Then for the Crank-Nicolson fully-discrete ...