Fact-checked by Grok 2 weeks ago

Five-point stencil

The five-point stencil is a approximation scheme used in to discretize the two-dimensional Laplacian operator on a uniform rectangular , involving a central grid point and its four immediate orthogonal neighbors (to the east, west, north, and south). This method provides a second-order accurate representation of the second partial derivatives, with a local of O(h^2), where h is the grid spacing, making it suitable for solving elliptic partial differential equations (PDEs) like the Poisson equation -\nabla^2 u = f. The discrete form of the Laplacian using the five-point stencil at an interior grid point (i,j) is given by \nabla^2 u_{i,j} \approx \frac{u_{i+1,j} + u_{i-1,j} + u_{i,j+1} + u_{i,j-1} - 4u_{i,j}}{h^2}, which rearranges into a -u_{i-1,j} - u_{i+1,j} - u_{i,j-1} - u_{i,j+1} + 4u_{i,j} = h^2 f_{i,j} for the Poisson equation on a with appropriate conditions, such as Dirichlet conditions where u=0 on the . This stencil produces a sparse, block-tridiagonal when applied across the grid, which is efficiently solved using iterative techniques like the Gauss-Seidel method or (SOR), often enhanced with ordering strategies such as red-black coloring for parallel computation. The is widely applied in and to model steady-state problems, including electrostatic potential distributions, heat conduction in solids, fields in incompressible , and load-bearing in seals or gas systems. Its simplicity and accuracy have made it a of methods since the early development of numerical PDE solvers, though extensions like nine-point stencils exist for higher-order accuracy in anisotropic or irregular domains.

Overview

Definition and Motivation

Finite difference methods provide a foundational approach in for approximating continuous derivatives by discretizing the domain into a grid of discrete points, enabling the solution of ordinary differential equations (ODEs) and partial differential equations (PDEs) through algebraic equations. These methods replace derivatives with differences between function values at grid points, facilitating computational simulations of physical phenomena where analytical solutions are intractable. The five-point stencil specifically refers to a that incorporates five equally spaced points—typically the central point and two points on each side in one dimension, or the central point and its four orthogonal neighbors in two dimensions—to estimate derivatives with second-order accuracy. This configuration allows for a more refined than basic schemes by capturing higher-order terms in the , while maintaining a compact support that limits the number of neighboring points involved. The motivation for employing the five-point stencil lies in its optimal balance between computational efficiency and accuracy, outperforming simpler three-point stencils in resolving complex behaviors without excessively increasing the stencil width or resource demands. It is extensively applied in solving PDEs such as the for diffusion processes and for steady-state potential fields, where second-order precision is essential for reliable simulations in fields like and . Historically, the five-point stencil originated in early 20th-century , first prominently featured in the 1928 work of Courant, Friedrichs, and Lewy for approximating , and gained widespread adoption in methods for and following the advent of digital computers in the post-1950s era.

Comparison to Lower-Order Stencils

The three-point stencil approximates the using the central grid point and its immediate neighbors, achieving second-order accuracy with an error term of O(h^2), where h is the grid spacing; however, it is limited in providing higher precision for more demanding applications due to its lower-order . In contrast, the five-point stencil extends this by incorporating two neighboring points on each side, enabling fourth-order accuracy for the or approximations. The seven-point stencil further broadens the support to three points on each side, attaining sixth-order accuracy with an error of O(h^6) for similar approximations, but at the expense of greater complexity in . This higher order reduces dispersion errors in wave-like problems, making it suitable for simulations requiring long-term , such as . Key trade-offs among these stencils involve balancing accuracy against stencil width and resulting system properties. The five-point stencil provides O(h^4) accuracy for second derivatives or enhanced in two-dimensional settings, using a moderate five-point width that avoids the excessive bandwidth of wider alternatives. The table below summarizes these aspects:
StencilGrid PointsAccuracy Order for Second DerivativeTypical Use CasesMatrix Bandwidth (1D PDE)
Three-point3O(h^2)Basic ODEs and simple 1D PDEs3 (tridiagonal)
Five-point5O(h^4)Higher-accuracy 1D problems5 (pentadiagonal)
Seven-point7O(h^6)High-precision simulations (e.g., wave propagation)7 (heptadiagonal)
Computationally, narrower stencils like the three-point require fewer floating-point operations (approximately 3-5 per approximation point) and less memory for storing sparse matrices, facilitating efficient direct solvers such as the algorithm. Wider stencils, such as the seven-point, demand more (up to 13-15 per point) and increased memory due to additional non-zero entries, while the elevated in resulting matrices from discretizations raises the cost of iterative or direct solvers, often scaling as O(n \cdot b^2) where b is the and n the grid size. The five-point stencil strikes a balance, offering substantial accuracy gains over the three-point with only modestly higher overhead, making it preferable for many practical schemes.

One-Dimensional Formulation

Central Difference for First Derivative

The central difference approximation for the first derivative using the five-point stencil in one dimension provides a fourth-order accurate estimate at an interior grid point x_i on a uniform grid with spacing h, given by f'(x_i) \approx \frac{-f(x_{i+2}) + 8f(x_{i+1}) - 8f(x_{i-1}) + f(x_{i-2})}{12h}. This formula is derived by expanding f(x_{i \pm k}) in around x_i for k = 1, 2, substituting into the with coefficients [-1/12, 8/12, 0, -8/12, 1/12] (multiplied by $1/h), and solving the system to match the coefficient of the linear term while canceling the constant, h, h^2, and h^3 terms, yielding O(h^4) . The leading term is -\frac{h^4}{30} f^{(5)}(\xi) for some \xi in the [x_{i-2}, x_{i+2}], obtained from the in the Taylor expansions after the fourth-order terms. To illustrate, consider f(x) = \sin x at x_i = 0, where the exact is f'(0) = 1. The approximations for decreasing h exhibit fourth-order , as the reduces by a factor of approximately 16 when h is halved, consistent with the analysis.
hApproximationAbsolute ErrorError Ratio (previous/current)
0.20.99994692$5.31 \times 10^{-5}
0.10.99999667$3.33 \times 10^{-6}15.94
0.050.99999979$2.08 \times 10^{-7}16.00

Higher-Order Derivatives

The five-point stencil can be extended to approximate higher-order derivatives in one dimension beyond the first derivative, leveraging symmetric points around the central grid point x_i to achieve central difference schemes. These approximations utilize function values at x_{i-2}, x_{i-1}, x_i, x_{i+1}, and x_{i+2}, with grid spacing h. The resulting formulas exhibit coefficient symmetry: even-order derivatives have symmetric coefficients (invariant under reversal), while odd-order derivatives have antisymmetric coefficients (changing sign under reversal), which aligns with the parity of the underlying Taylor series terms. For the second derivative, the five-point central difference formula is f''(x_i) \approx \frac{-f(x_{i-2}) + 16f(x_{i-1}) - 30f(x_i) + 16f(x_{i+1}) - f(x_{i+2})}{12h^2}, achieving fourth-order accuracy, O(h^4), due to cancellation of lower-order error terms up to h^2. This improves upon the three-point stencil's second-order accuracy and is particularly useful in applications requiring precise estimates, such as solving elliptic PDEs. The third derivative approximation using the five-point stencil is f'''(x_i) \approx \frac{f(x_{i+2}) - 2f(x_{i+1}) + 2f(x_{i-1}) - f(x_{i-2})}{2h^3}, with second-order accuracy, O(h^2). This lower order relative to the second derivative arises because odd-order central differences inherently retain a leading h^2 error term from the expansion, without the even-order cancellation benefits. For the fourth derivative, the formula is f^{(4)}(x_i) \approx \frac{f(x_{i-2}) - 4f(x_{i-1}) + 6f(x_i) - 4f(x_{i+1}) + f(x_{i+2})}{h^4}, also second-order accurate, O(h^2), with the involving the sixth derivative. This is exact for polynomials of degree up to 4, as the coefficients match the fourth forward difference operator precisely for such functions. The symmetric coefficients facilitate extensions to mixed partial derivatives in multidimensional settings, where products of one-dimensional approximate cross terms like \partial^2 f / \partial x \partial y while preserving properties.

Derivation Using Taylor Expansion

The derivation of five-point stencil coefficients in one dimension employs expansions to approximate derivatives by matching series terms and eliminating lower-order errors. Consider a smooth function f(x) evaluated on a with spacing h, at points x_i + kh for integers k. The expansion around the central point x_i for each neighboring point is f(x_i + kh) = f(x_i) + (kh) f'(x_i) + \frac{(kh)^2}{2!} f''(x_i) + \frac{(kh)^3}{3!} f'''(x_i) + \frac{(kh)^4}{4!} f^{(4)}(x_i) + O(h^5), where higher-order terms are truncated for analysis up to fourth-order accuracy. To derive a stencil approximating the m-th f^{(m)}(x_i), form a of the five function values at k = -2, -1, 0, 1, 2: \sum_{k=-2}^{2} c_k f(x_i + kh) = a h^m f^{(m)}(x_i) + O(h^{5}), where the coefficients c_k and scaling a are chosen to match the coefficient of f^{(m)}(x_i) while canceling contributions from derivatives of orders n = 0 to m-1 and m+1 to $4, ensuring fourth-order accuracy. Substituting the Taylor expansions yields a system of five equations by equating coefficients of like powers of h from h^0 to h^4. This system can be expressed in matrix form as \mathbf{V} \mathbf{c} = \mathbf{b}, where \mathbf{c} = [c_{-2}, c_{-1}, c_0, c_1, c_2]^T is the of coefficients, \mathbf{b} has a single nonzero entry a in the row corresponding to order m (and zeros elsewhere for the canceled terms), and \mathbf{V} is a $5 \times 5 with entries V_{j,k} = \frac{k^j}{j!} for row j = 0 to $4 (orders of derivatives) and column k = -2 to $2. Solving this determines the c_k, which, when normalized by h^m, provide the approximation to f^{(m)}(x_i). For instance, this approach confirms the standard centered five-point stencil for the first derivative without relying on methods.

Connection to Polynomial Interpolation

The five-point stencil in one dimension provides a framework for approximating derivatives by constructing a of degree at most 4 that passes through the values at five equally spaced points, typically denoted as x_{i-2}, x_{i-1}, x_i, x_{i+1}, and x_{i+2}, with uniform spacing h = x_{i+1} - x_i. The interpolating P(x) is given by P(x) = \sum_{k=-2}^{2} f(x_{i+k}) \, l_k(x), where the are l_k(x) = \prod_{\substack{m=-2 \\ m \neq k}}^{2} \frac{x - x_{i+m}}{x_{i+k} - x_{i+m}}. This construction ensures P(x_{i+j}) = f(x_{i+j}) for j = -2, -1, 0, 1, 2, providing an exact representation for any of at most 4. To approximate the first derivative at x_i, evaluate the derivative of the interpolant: f'(x_i) \approx P'(x_i). Substituting the form of P(x) yields P'(x_i) = \sum_{k=-2}^{2} f(x_{i+k}) \, l_k'(x_i), where the derivatives of the basis functions at the evaluation point are l_k'(x_i) = \sum_{\substack{m=-2 \\ m \neq k}}^{2} \frac{1}{x_i - x_{i+m}} \prod_{\substack{j=-2 \\ j \neq k, j \neq m}}^{2} \frac{x_i - x_{i+j}}{x_{i+k} - x_{i+j}}. For a uniform grid, these evaluate to specific coefficients, resulting in the central difference approximation f'(x_i) \approx \frac{1}{12h} \left[ -f(x_{i+2}) + 8f(x_{i+1}) - 8f(x_{i-1}) + f(x_{i-2}) \right], which matches the coefficients derived via expansion and achieves fourth-order accuracy, exact for polynomials up to degree 4. This interpolative approach unifies methods with theory, offering an alternative perspective to direct series methods while confirming equivalent results. A key advantage of this formulation is its theoretical insight into error bounds and exactness: since P(x) exactly matches f(x) for polynomials of \leq 4, the P'(x_i) (or higher s) is exact in those cases, with the error for general smooth functions governed by the term in the error, leading to O(h^4) for the first . This connection also extends naturally to higher derivatives; for the second , f''(x_i) \approx P''(x_i), where P''(x) is obtained by twice differentiating the interpolant. For uniform spacing, this yields the centered formula f''(x_i) \approx \frac{ -f(x_{i-2}) + 16f(x_{i-1}) - 30f(x_i) + 16f(x_{i+1}) - f(x_{i+2}) }{12h^2}, again exact for polynomials up to 4 and fourth-order accurate.

Two-Dimensional Formulation

Standard Cross-Shaped Stencil

The standard cross-shaped five-point stencil in two dimensions approximates the Laplacian operator \nabla^2 u = \frac{\partial^2 u}{\partial x^2} + \frac{\partial^2 u}{\partial y^2} at an interior grid point (i, j) using the values of the function u at five points: the central point (i, j) and its four orthogonal neighbors (i+1, j), (i-1, j), (i, j+1), and (i, j-1). This arrangement forms a cross pattern on the grid, which is particularly suited for Cartesian coordinates due to its alignment with the coordinate axes. The approximation is given by the finite difference formula: \nabla^2 u_{i,j} \approx \frac{u_{i+1,j} + u_{i-1,j} + u_{i,j+1} + u_{i,j-1} - 4u_{i,j}}{h^2}, where h is the uniform grid spacing, assuming \Delta x = \Delta y = h. This stencil achieves second-order accuracy, with a local of O(h^2) in each spatial direction, and it exhibits isotropic behavior on square grids where the spacing is equal in both directions. The method can be generalized to rectangular grids with unequal spacings \Delta x and \Delta y, yielding: \nabla^2 u_{i,j} \approx \frac{u_{i+1,j} - 2u_{i,j} + u_{i-1,j}}{\Delta x^2} + \frac{u_{i,j+1} - 2u_{i,j} + u_{i,j-1}}{\Delta y^2}.[18] The assumes a uniform rectangular grid over the computational , enabling straightforward implementation via tensor products of one-dimensional difference operators. Near boundaries, handling requires modifications such as incorporating Dirichlet or conditions directly into the or using auxiliary techniques like ghost points, which extend the grid fictitiously beyond the to maintain the standard cross pattern while satisfying constraints.

Derivation and Isotropic Variants

The derivation of the two-dimensional five-point stencil for approximating the Laplacian operator \nabla^2 u = \frac{\partial^2 u}{\partial x^2} + \frac{\partial^2 u}{\partial y^2} begins with multivariable Taylor expansions around the grid point (x_i, y_j), assuming uniform spacing h in both directions for simplicity. Consider the expansions for the neighboring points: u(x_i + h, y_j) = u + h u_x + \frac{h^2}{2} u_{xx} + \frac{h^3}{6} u_{xxx} + \frac{h^4}{24} u_{xxxx} + O(h^5), u(x_i - h, y_j) = u - h u_x + \frac{h^2}{2} u_{xx} - \frac{h^3}{6} u_{xxx} + \frac{h^4}{24} u_{xxxx} + O(h^5), where all partial derivatives are evaluated at (x_i, y_j). Adding these yields u(x_i + h, y_j) + u(x_i - h, y_j) - 2u = h^2 u_{xx} + 2 \cdot \frac{h^4}{24} u_{xxxx} + O(h^6) = h^2 u_{xx} + \frac{h^4}{12} u_{xxxx} + O(h^6). Dividing by h^2 gives \frac{u(x_i + h, y_j) + u(x_i - h, y_j) - 2u}{h^2} = u_{xx} + \frac{h^2}{12} u_{xxxx} + O(h^4). An analogous expansion in the y-direction produces \frac{u(x_i, y_j + h) + u(x_i, y_j - h) - 2u}{h^2} = u_{yy} + \frac{h^2}{12} u_{yyyy} + O(h^4). Summing these approximations combines the one-dimensional central second differences, yielding the five-point stencil operator \Lambda u_{i,j} = \frac{u_{i+1,j} + u_{i-1,j} + u_{i,j+1} + u_{i,j-1} - 4u_{i,j}}{h^2} = \nabla^2 u + \frac{h^2}{12} \left( \frac{\partial^4 u}{\partial x^4} + \frac{\partial^4 u}{\partial y^4} \right) + O(h^4). This discretization achieves second-order accuracy, O(h^2), for the Poisson equation \nabla^2 u = f. For the Laplace equation (f = 0), the leading term can be rewritten using the PDE. Since \nabla^2 u = 0, applying \nabla^2 again gives \nabla^4 u = \frac{\partial^4 u}{\partial x^4} + 2 \frac{\partial^4 u}{\partial x^2 \partial y^2} + \frac{\partial^4 u}{\partial y^4} = 0, so \frac{\partial^4 u}{\partial x^4} + \frac{\partial^4 u}{\partial y^4} = -2 \frac{\partial^4 u}{\partial x^2 \partial y^2}. Substituting yields a of -\frac{h^2}{6} \frac{\partial^4 u}{\partial x^2 \partial y^2} + O(h^4), highlighting the anisotropic nature of the error, which depends on mixed derivatives and favors alignment with the grid axes. On non-square grids with spacings h_x \neq h_y, the stencil generalizes to \Lambda u_{i,j} = \frac{u_{i+1,j} + u_{i-1,j} - 2u_{i,j}}{h_x^2} + \frac{u_{i,j+1} + u_{i,j-1} - 2u_{i,j}}{h_y^2}, with \frac{h_x^2}{12} \frac{\partial^4 u}{\partial x^4} + \frac{h_y^2}{12} \frac{\partial^4 u}{\partial y^4} + O(\max(h_x^4, h_y^4)). This introduces additional , as the differing coefficients amplify directional biases in the error, potentially slowing or distorting solutions in stretched domains. For the Laplace case, a similar using \nabla^4 u = 0 yields a more complex mixed-derivative form that does not simply scale as in the equal-spacing case, further emphasizing the . To mitigate rotational invariance issues in the standard axis-aligned five-point stencil, modifications such as weighted crosses or 45-degree rotated variants improve angular accuracy while retaining a five-point support. A 45-degree rotated stencil uses the center and four diagonal neighbors (i \pm 1, j \pm 1), scaled by the diagonal distance h\sqrt{2}; the approximation becomes \Lambda u_{i,j} = \frac{u_{i+1,j+1} + u_{i+1,j-1} + u_{i-1,j+1} + u_{i-1,j-1} - 4u_{i,j}}{2h^2}, derived similarly via Taylor expansions along diagonal directions, yielding second-order accuracy but with error terms rotated by 45 degrees, better suiting problems with diagonal features. Weighted variants adjust coefficients (e.g., non-equal weights on arms) to balance isotropy, often tuned via moment-matching for specific angular responses. In contrast, the nine-point stencil incorporates diagonal points with weights (e.g., \frac{1}{6h^2} [4(u_{i+1,j} + u_{i-1,j} + u_{i,j+1} + u_{i,j-1}) + (u_{i+1,j+1} + u_{i+1,j-1} + u_{i-1,j+1} + u_{i-1,j-1}) - 20 u_{i,j} ]), achieving a leading truncation error of O(h^2) proportional to \nabla^4 u, which is fully rotationally invariant; for solutions to the Laplace equation where \nabla^4 u = 0, the error is O(h^4). As an illustrative application, consider the Poisson equation \nabla^2 u = -1 on the unit disk with Dirichlet u=0 on r=1. Discretizing with the five-point stencil on a Cartesian (approximating the circular via ) yields solutions converging at second order; numerical tests on with h = 1/64 to $1/256 show maximum errors reducing by a factor of approximately 4 per halving of h, confirming the O(h^2) rate, though irregularities introduce minor logarithmic factors.

Applications and Extensions

Use in Partial Differential Equations

The five-point stencil plays a central role in schemes for solving partial differential equations (PDEs) by discretizing spatial derivatives, particularly the Laplacian operator, on structured grids. This approach transforms continuous PDEs into algebraic systems that can be solved numerically, enabling simulations of physical phenomena involving , , and wave propagation. In two dimensions, the stencil approximates the second derivative terms using the cross-shaped arrangement of a central point and its four immediate neighbors, as detailed in the standard formulation. For the , \partial u / \partial t = \alpha \nabla^2 u, the five-point discretizes the term \nabla^2 u in both one- and two-dimensional settings. In explicit time-stepping methods, such as the forward Euler scheme, the solution at the next time step is computed as u^{n+1}_j = u^n_j + \alpha \Delta t \, D^2_{FD} u^n_j, where D^2_{FD} denotes the Laplacian approximated via the ; this requires satisfying a condition. In 1D, \Delta t \leq 2 \Delta x^2 / (\alpha \pi^2 L^2) (domain-dependent); in 2D, \Delta t \leq h^2 / (4 \alpha). Implicit methods, including Crank-Nicolson, incorporate the into a (I - \alpha \Delta t / 2 \, D^2_{FD}) u^{n+1} = (I + \alpha \Delta t / 2 \, D^2_{FD}) u^n, offering unconditional for larger time steps and suitability for long-term simulations of heat conduction. In solving Poisson and Laplace equations, such as -\nabla^2 u = f, the two-dimensional five-point stencil leads to the discrete form -\Delta_h u_{i,j} = (4u_{i,j} - u_{i+1,j} - u_{i-1,j} - u_{i,j+1} - u_{i,j-1}) / h^2 = f_{i,j}, assembling a sparse A for the A u = f. Boundary conditions, like Dirichlet or , modify the matrix entries and right-hand side accordingly, preserving sparsity with five non-zero entries per row. Iterative solvers, such as Gauss-Seidel, exploit this structure by updating each grid point using the latest neighbor values, converging efficiently for well-posed problems in electrostatic potential modeling. For advection-diffusion equations, like \partial u / \partial t + \mathbf{v} \cdot \nabla u = \kappa \nabla^2 u, the five-point stencil approximates the term while combining with upwind schemes for the to ensure stability. In two dimensions, compact integrated variants generate a five-point stencil that yields tridiagonal systems when paired with alternating direction implicit (ADI) time-stepping, reducing computational cost and achieving higher accuracy in flows with dominant , as verified in benchmarks like lid-driven cavity simulations. Real-world applications leverage the stencil's efficiency in image processing, where the Laplacian via the five-point serves as an filter by highlighting intensity discontinuities; for instance, convolving an image with the stencil kernel [0, 1, 0; 1, -4, 1; 0, 1, 0] identifies boundaries in data. In geophysical modeling, the discretizes equations for seismic , using five-point operators in staggered-grid schemes to simulate wavefields in heterogeneous media, supporting inversion tasks in exploration with resolutions of 10 points per . Recent extensions include high-order five-point stencils based on integrated (IRBF) s for solving biharmonic equations in , achieving higher accuracy on irregular grids.

Error Analysis and Stability Considerations

The five-point stencil provides a higher-order approximation for derivatives compared to lower-order schemes, but its truncation error depends on the specific and . In one , the centered five-point stencil for the first , given by u'(x_j) = \frac{-u_{j+2} + 8u_{j+1} - 8u_{j-1} + u_{j-2}}{12h} + O(h^4), achieves a truncation error of order O(h^4), where h is the grid spacing. This fourth-order accuracy arises from canceling lower-order terms in the expansion up to h^3. In two dimensions, the standard cross-shaped five-point stencil for the Laplacian , \Delta u_{i,j} \approx \frac{u_{i+1,j} + u_{i-1,j} + u_{i,j+1} + u_{i,j-1} - 4u_{i,j}}{h^2} + O(h^2), yields a truncation error of order O(h^2), as the leading error terms involve fourth derivatives scaled by h^2. For wave equations, the five-point stencil introduces both dispersion errors, which manifest as deviations in phase velocity for different Fourier modes, and dissipation errors, leading to artificial amplitude damping; these errors are minimized in optimized schemes that target waves with at least five points per wavelength. Round-off errors in the five-point stencil stem from limitations during the evaluation of differences and summations across the stencil points, with the total error bounded by \epsilon \approx 10^{-16} amplified by factors such as O(1/h^p) for p-th derivatives. This amplification occurs because small perturbations in function values are scaled inversely by the grid spacing in the denominator, potentially dominating the for sufficiently small h. The discrete matrices arising from five-point discretizations, such as the Laplacian on a uniform grid, exhibit a spectral \kappa_2(A) = O(1/h^2), as the smallest eigenvalue remains O(1) while the largest grows as O(1/h^2); this ill-conditioning can exacerbate propagation in iterative solvers. Stability analysis for schemes employing the five-point stencil, particularly in time-dependent methods, relies on the approach, which decomposes solutions into modes e^{i \kappa n} and examines the amplification factor G(\kappa) for each \kappa; stability requires |G(\kappa)| \leq 1 + O(\Delta t) for all modes to prevent exponential growth. For explicit schemes, such as the forward-time central-space method for the two-dimensional using the five-point Laplacian , the Courant-Friedrichs-Lewy (CFL) condition imposes \Delta t \leq h^2 / 4 (assuming unit ) to ensure the numerical domain of dependence encompasses the physical one and bounds the . This condition arises from the maximum eigenvalue of the spatial operator being approximately $8/h^2, requiring the time step to damp high-frequency modes without instability. Despite its advantages, the five-point stencil has limitations on irregular grids, where stencil asymmetry leads to reduced accuracy (e.g., O(h) errors on high-aspect-ratio meshes) and requires adaptive compact stencils to maintain second-order by adjusting point selections. In flows with high Reynolds numbers, the stencil's fixed structure struggles with thin layers and convective dominance, often necessitating adaptive refinement or upwinding to avoid oscillations, as standard central differences amplify . Compared to methods, which achieve for smooth solutions via global basis functions, the five-point stencil's algebraic O(h^2) accuracy demands finer grids for equivalent precision, making it less efficient for periodic or analytic problems but more robust on complex domains.

References

  1. [1]
    [PDF] Finite Difference Methods For Poisson Equation
    Oct 5, 2025 · It is called five point stencil since there are only five points involved. When hx = hy, it is simplified to. (4). − (∆hu)i,j = 4ui,j − ui+ ...
  2. [2]
    Finite difference methods - Computational Physics
    The expression above is known as a five-point stencil as it uses five points to calculate the Laplacian. Keypoints. Finite difference methods are a family of ...
  3. [3]
    [PDF] Lecture 9: Numerical Partial Differential Equations(Part 2)
    • Stencil of finite difference approximation. 2. Page 3. • Finite ... • Five-point stencil of the finite difference approximation. 5. Page 6. Natural ...
  4. [4]
    Point Finite Difference - an overview | ScienceDirect Topics
    The 5-point finite difference scheme is defined as a method for approximating the Laplacian operator using a grid of nodes, where the value at a given node ...
  5. [5]
  6. [6]
    [PDF] Accurate Finite Difference Algorithms
    grid point. These al_thms have three, five, seven, and nine point central sten- c/Is, they are second, fourth, sixth, and eighth order accurate in both.Missing: cost | Show results with:cost
  7. [7]
    None
    Below is a merged summary of the five-point stencil from LeVeque's "Finite Difference Methods," consolidating all information from the provided segments into a comprehensive response. To retain maximum detail and clarity, I will use a structured format with a table for key details, followed by additional narrative information where necessary. The response avoids redundancy while ensuring all unique points are included.
  8. [8]
    [PDF] FINITE DIFFERENCE METHODS (II): 1D EXAMPLES IN MATLAB
    We may use fdcoefs to derive general finite difference formulas. Let's compute, for example, the weights of the 5-point, centered formula for the first ...
  9. [9]
    [PDF] Numerical differentiation: finite differences
    If we need to estimate the rate of change of y with respect to x in such a situation, we can use finite difference formulas to compute approximations of f0(x).
  10. [10]
    [PDF] Homework 5 Solutions
    = h4 20 f(5)(ξ). Note that we used the Intermideate Value Theorem above, similar to: f(5)(ξ) = 1 2 (f(5)(ξ1) + f(5)(ξ2)). = − h4 30 f(5)(ξ). Thus, the error ...
  11. [11]
    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)
  12. [12]
    [PDF] Brief Summary of Finite Difference Methods
    The algorithm is particularly convenient to use when only a small number of stencils are considered, and when one wants to obtain the weights in exact rational ...
  13. [13]
    [PDF] Finite Difference Schemes
    5 Finite difference formulas for fourth derivatives. Central finite differece scheme second order: ∂4u. ∂x4 xi. = ui+2 − 4ui+1 + 6ui − 4ui−1 + ui−2. ∆x4.
  14. [14]
    [PDF] Lecture 1 Taylor series and finite differences
    To find a formula for the second derivative, f , we must use a stencil of at least 3 points. Here, we'll consider a uniform grid with spacing x. The. Taylor ...
  15. [15]
    [PDF] an introduction to compact finite differences - The University of Bath
    For second order methods it is generally the case that each derivative is simply replaced by its finite difference stencil, but for compact fourth order methods ...
  16. [16]
    [PDF] Polynomial approximation of functions and derivatives
    (Lagrange interpolation formula). (11). Let us verify that Πnf(x) indeed ... Figure 10: Five-point stencil for the fourth-order finite difference ...
  17. [17]
    [PDF] Numerical Differentiation is Easy - arnold@uark.edu
    All of the commonly used formulas for approximating f0(c) (and f00, etc.) can be derived from the Lagrange interpolation result: f(x) = Pn(x) + Rn(x), where R(x) ...
  18. [18]
    [PDF] MA615 Numerical Methods for PDEs Spring 2024 Lecture Notes
    ... finite difference and finite element methods are completely different. To be precise, the traditional finite difference approach simply approximates the ...
  19. [19]
    [PDF] Stencil Computations for PDE-based Applications with Examples ...
    Jan 29, 2016 · In general, stencils represent computational patterns that repeat across the computational domain. They apply to finite differences, finite ...
  20. [20]
    None
    Below is a merged summary of all the provided segments on the Two-Dimensional Poisson Equation and related topics (Five-Point Stencil, Derivation Using Taylor, Error Analysis, and Convergence Rates). To retain all information in a dense and organized manner, I will use a combination of narrative text and a table in CSV format for detailed section-specific data. The narrative provides an overview, while the table captures specifics such as section references, equations, and key findings across the document.
  21. [21]
    [PDF] Finite Difference Methods for Elliptic Equations
    Five point stencil for approximating the Laplacian. Inserting the approximation of the Laplacian with the five point stencil (2.4) for x = (x, y) ∈ ω◦h in ...
  22. [22]
    [PDF] Finite Difference Method for the Solution of Laplace Equation
    5-point stencil for Laplace equation. A finite difference equation suitable ... system can be obtained by considering Laplace's equation at the nodal point i, j ...Missing: five- expansion
  23. [23]
    [PDF] Indoor Path Planning for Mobile Robot using Half-Sweep Gauss ...
    5-point stencil shown in Figure 2 (a), whereas half- sweep iteration is actually based on the five-point. 45-degree rotated finite difference approximation as.
  24. [24]
    Isotropic Finite Differences Based on Lattice-Boltzmann Stencils - PMC
    We propose isotropic finite differences for high-accuracy approximation of high-rank derivatives. These finite differences are based on direct application ...Missing: sources | Show results with:sources<|control11|><|separator|>
  25. [25]
    [PDF] collocation spectral methods in the solution of - UBC Open Collections
    We solved the Poisson on a unit disk with spectral method and finite difference method with SOR. In both the spectral method and the SOR, the exact value of ...
  26. [26]
    [PDF] Numerical methods for the heat equation - Daniele Venturi
    For example, we could have used a stencil with 5 points in space and the BDF3 method in time. Absolute stability analysis of finite-difference methods.
  27. [27]
  28. [28]
    Mixed-grid and staggered-grid finite-difference methods for ...
    This leads to five-point and nine-point stencils for the acoustic and elastic wave equations respectively. Their scheme required 10 gridpoints per shortest ...<|control11|><|separator|>
  29. [29]
    High-order, low dispersive and low dissipative explicit schemes for ...
    Jun 10, 2007 · They are constructed by minimizing the dispersion and the dissipation errors in the wave number space for waves down to four points per ...
  30. [30]
    [PDF] Notes: von Neumann Stability Analysis - MIT Mathematics
    Apr 26, 2022 · When these Fourier modes are evaluated on the numerical grid, as in equation (1.2), they give rise to a dependence on the grid indexes of the ...
  31. [31]
    [PDF] 5 Finite differences: and what about 2D?
    The 2D heat equation can be discretized using explicit or implicit methods. Explicit methods are simple, while implicit methods can solve stability issues. ...
  32. [32]
    [PDF] Comparison of node-centered and cell-centered unstructured finite ...
    The adaptive compact stencil (CC-CS) is important for discretizations on high-aspect-ratio grids of types (II) and (III) to represent correctly the direction of ...
  33. [33]
    [PDF] Adaptive Grid Techniques for Elliptic Fluid-Flow Problems, - DTIC
    for the five-point stencil shown in Fig. 3.5. We have assumed that the ... using CD for cell Reynolds numbers as high as 150 (see Chapter 8). Kim and ...
  34. [34]
    [PDF] Spectral and High-Order Methods with Applications - Purdue Math
    Aug 1, 2012 · This book expands lecture notes by the authors for a course on Introduction of Spec- tral Methods taught in the past few years at Penn State ...