Fact-checked by Grok 2 weeks ago

Stiffness matrix

The stiffness matrix is a square in that linearly relates the of applied nodal forces \mathbf{F} to the of corresponding nodal displacements \mathbf{u} in a discretized , expressed by the equation \mathbf{K} \mathbf{u} = \mathbf{F}, where \mathbf{K} denotes the stiffness matrix itself. This encapsulates the inherent resistance of the to deformation, with its elements representing stiffness coefficients that quantify how forces at one node influence displacements at others. It forms the cornerstone of computational methods for analyzing determinate and indeterminate structures, such as trusses, beams, and , by enabling the systematic solution of equilibrium equations. In practice, the stiffness matrix is constructed using the direct stiffness method, a matrix-based approach where individual element stiffness matrices—derived from material properties, geometry, and boundary conditions—are transformed from local to global coordinates and assembled into a global stiffness matrix that enforces overall equilibrium. This assembly process accounts for degrees of freedom at nodes (e.g., two per node for 2D trusses or three for 2D frames), allowing the solution for unknown displacements at free degrees of freedom before back-substituting to find reactions and internal forces. The method's efficiency stems from its adaptability to computer implementation, making it indispensable for finite element analysis (FEA) in engineering disciplines like civil, mechanical, and aerospace. Key mathematical properties of the stiffness matrix include (\mathbf{K} = \mathbf{K}^T), arising from Maxwell's reciprocity theorem, which states that the displacement at one point due to a unit force at another equals the displacement at the second point due to a unit force at the first. For stable structures, \mathbf{K} is positive definite, with all eigenvalues positive, ensuring a unique solution to the displacement equations; zero or negative eigenvalues indicate instability or rigid body modes. These properties facilitate numerical solving techniques, such as , while the matrix's sparsity (due to localized element interactions) optimizes computational performance in large-scale simulations. The development of the stiffness matrix traces its roots to 19th-century structural theory, with early formulations appearing in Alfred Clebsch's 1862 work on three-dimensional pinned trusses, building on principles from Louis Navier (1826) and Barré de Saint-Venant (1838). Iterative advancements, including Heinrich Manderla's prize-winning method for secondary stresses (1880) and Otto Mohr's modifications (1892), laid groundwork for matrix approaches in the early , alongside Hardy Cross's (1930). The modern emerged in the 1950s amid demands for analyzing complex frames, with pivotal contributions from Gabriel Kron's tensor-based formulations (1944) and John H. Argyris's matrix extensions (1954–1955). Its formalization as part of the occurred in the seminal 1956 paper by M. J. , R. W. Clough, H. C. , and L. J. Topp, which introduced displacement-based stiffness assembly for continuum structures. This evolution transformed stiffness matrix applications from discrete frames to versatile FEA tools for simulating real-world phenomena like vibrations, , and nonlinear behaviors.

Fundamentals

Definition and basic properties

In linear elasticity and structural mechanics, the stiffness matrix, denoted as \mathbf{K}, is a square matrix that relates the applied force vector \mathbf{F} to the resulting displacement vector \mathbf{u} through the linear equation \mathbf{F} = \mathbf{K} \mathbf{u}. This relation captures the system's resistance to deformation under external loads, assuming small displacements and linear material behavior where holds. The stiffness matrix exhibits several properties. It is symmetric, satisfying \mathbf{K}^T = \mathbf{K}, which follows from reciprocity theorem stating that the displacement at one point due to a force at another equals the displacement at the second point due to the same force at the first. For stable, physically realistic systems without or instability, \mathbf{K} is positive definite, meaning all its eigenvalues are positive and the quadratic form \mathbf{u}^T \mathbf{K} \mathbf{u} > 0 for any nonzero \mathbf{u}, ensuring positive storage. However, if the system is unconstrained (e.g., free-floating), \mathbf{K} is singular with zero eigenvalues corresponding to rigid body modes, allowing arbitrary translations or rotations without deformation. The stiffness matrix arises from the principle of minimum in conservative systems. The total potential energy \Pi is the sum of the strain energy U (internal elastic energy) and the potential of external forces \Omega (negative work done by loads). displacements minimize \Pi, and setting the first variation \delta \Pi = 0 yields the force-displacement equations; the second variation corresponds to the of U, which is precisely \mathbf{K}. A simple illustration is a single spring with constant k, fixed at one end and loaded at the other, representing a one-degree-of-freedom where the stiffness is the scalar K = k, so F = k u. This extends to multi-degree-of-freedom systems, such as a connected between two nodes with displacements u_1 and u_2, yielding the stiffness matrix \mathbf{k} = k \begin{bmatrix} 1 & -1 \\ -1 & 1 \end{bmatrix}, relating nodal forces to relative displacements via \mathbf{f} = \mathbf{k} \mathbf{u}.

Physical interpretation

The stiffness matrix generalizes the scalar stiffness concept from , which describes the linear relationship between force and for a single-degree-of-freedom as F = k x, to multi-dimensional where deformations occur in multiple directions simultaneously. In this matrix form, the relationship becomes \mathbf{F} = \mathbf{K} \mathbf{u}, where \mathbf{K} captures the coupled resistance of the structure to vector \mathbf{u}, reflecting material properties, geometry, and connectivity that determine how loads induce deformations across . The entries of the stiffness matrix provide a direct physical into these interactions: the element K_{ij} denotes the force developed at the i-th degree of freedom due to a unit imposed at the j-th degree of freedom, with all other displacements constrained to zero. Positive diagonal terms K_{ii} indicate the inherent against direct deformation in that direction, while off-diagonal terms K_{ij} (for i \neq j) reveal effects, where the sign determines whether the induced force opposes or aids the displacement, arising from the structure's and conditions. In the context of structural , the stiffness matrix represents the collective of the system, such that the product \mathbf{K} \mathbf{u} generates the nodal forces necessary to counteract applied external loads \mathbf{F} in \mathbf{K} \mathbf{u} = \mathbf{F}. This ensures that the deformed balances internal restoring forces against external actions, maintaining static . For instance, in a simple 2D frame composed of beams, the diagonal entries embody direct stiffness contributions like axial rigidity along member lengths and flexural resistance to transverse loads, whereas off-diagonal entries account for geometric , such as how a horizontal at a induces vertical forces due to inclined or connected members. The of the stiffness matrix further underscores its physical reliability, guaranteeing a unique and stable response to any load without spontaneous deformation.

Formulation in mechanics

Discrete systems

In discrete systems, such as lumped-parameter models, the stiffness matrix assembles from individual element contributions to relate nodal forces to displacements across the . These models approximate continuous structures by concentrating and at discrete nodes connected by elements like springs or bars, enabling efficient computation of static and dynamic responses. The in such systems correspond to the possible nodal and associated forces. In one-dimensional models, like axial chains, each has a single of along the , with forces acting in the same direction. For two-dimensional structures, each typically has two —horizontal and vertical —with corresponding force components. In three dimensions, nodes in space exhibit three translational per . Lumped models assemble the global stiffness matrix by superimposing element-level matrices, ensuring at shared nodes. For a mass-spring , consider two springs in series connecting three nodes, with spring constants k_1 and k_2. The local stiffness matrix for each spring is \begin{bmatrix} k & -k \\ -k & k \end{bmatrix}, relating end forces to relative displacements. Assembly yields the 3×3 global stiffness matrix: K = \begin{bmatrix} k_1 & 0 & -k_1 \\ 0 & k_2 & -k_2 \\ -k_1 & -k_2 & k_1 + k_2 \end{bmatrix}, where the off-diagonal terms reflect force at nodes. In structures, element matrices account for and . For a two-dimensional bar element with cross-sectional area A, E, and length L at \theta to the global x-axis, the element stiffness matrix is K_e = \frac{AE}{L} \mathbf{T}^T \begin{bmatrix} 1 & -1 \\ -1 & 1 \end{bmatrix} \mathbf{T}, where \mathbf{T} is the incorporating \cos\theta and \sin\theta to rotate between local and global coordinates. Explicitly, K_e = \frac{AE}{L} \begin{bmatrix} c^2 & cs & -c^2 & -cs \\ cs & s^2 & -cs & -s^2 \\ -c^2 & -cs & c^2 & cs \\ -cs & -s^2 & cs & s^2 \end{bmatrix}, with c = \cos\theta and s = \sin\theta. For a two-bar , the global matrix forms by adding contributions from each element's K_e, expanded to the full and condensed for connectivity. Boundary conditions incorporate supports by modifying the stiffness matrix or the force . For fixed supports (zero ), partition the matrix by removing rows and columns corresponding to constrained , reducing the system to free nodes. For prescribed non-zero displacements, adjust the force by subtracting stiffness terms multiplied by the known displacements, maintaining . This ensures the solved displacements satisfy structural constraints without altering the matrix symmetry or positive definiteness.

Continuous systems via finite elements

In the finite element method (FEM), stiffness matrices for continuous systems are derived by discretizing the domain into finite elements and approximating the governing partial differential equations through variational principles. This approach transforms the continuous problem into a system of algebraic equations where the element stiffness matrix relates nodal displacements to forces, enabling the analysis of structures like beams, plates, and solids under elastic deformation. The derivation begins with the weak form of the equilibrium equations, obtained by multiplying the strong form by a test and integrating over the , often using to reduce derivative orders and incorporate natural boundary conditions. For , the weak form expresses the principle of : the internal virtual work equals the external virtual work, leading to an over the element volume. The then approximates both the solution and test functions within this weak form using the same basis, ensuring symmetry in the resulting matrices for self-adjoint problems and promoting . Within each element, displacements are approximated as \mathbf{u} = \mathbf{N} \mathbf{d}, where \mathbf{N} is the matrix of shape functions and \mathbf{d} contains the nodal displacement values. Shape functions, typically polynomials like linear or quadratic, interpolate the field variables and ensure inter-element continuity (e.g., C^0 continuity for displacements). Strains are then derived as \boldsymbol{\epsilon} = \mathbf{B} \mathbf{d}, with the strain-displacement matrix \mathbf{B} formed from derivatives of the shape functions. The material constitutive relation links stresses to strains via \boldsymbol{\sigma} = \mathbf{D} \boldsymbol{\epsilon}, where \mathbf{D} is the material matrix (e.g., for isotropic elasticity, involving Young's modulus and Poisson's ratio). Substituting into the weak form yields the element stiffness matrix as \mathbf{K}_e = \int_{V_e} \mathbf{B}^T \mathbf{D} \mathbf{B} \, dV, which represents the resistance to deformation within the element. For complex geometries, isoparametric elements map the physical element coordinates to a simple reference (parent) domain using the same shape functions for both geometry and displacements: \mathbf{x} = \sum N_i \mathbf{x}_i and \mathbf{u} = \sum N_i \mathbf{d}_i. This mapping involves the Jacobian matrix \mathbf{J} for coordinate transformation, adjusting the integration to the reference space as \mathbf{K}_e = \int_{\hat{V}_e} \mathbf{B}^T \mathbf{D} \mathbf{B} \det(\mathbf{J}) \, d\hat{V}, where d\hat{V} = d\xi \, d\eta \, d\zeta in natural coordinates (e.g., \xi \in [-1,1]). This formulation allows flexible handling of curved boundaries while maintaining computational efficiency through numerical quadrature. A simple example is the 1D bar element under axial loading, with length L, cross-sectional area A, and E. Using linear shape functions N_1 = 1 - \xi and N_2 = \xi (where \xi = x/L), the is u(x) = (1 - \xi) d_1 + \xi d_2, and the strain is constant: \epsilon = (d_2 - d_1)/L. The B is \mathbf{B} = [-1/L, 1/L], and with \mathbf{D} = E, the stiffness integrates to \mathbf{K}_e = \frac{EA}{L} \begin{bmatrix} 1 & -1 \\ -1 & 1 \end{bmatrix}. This 2x2 captures the element's axial stiffness, with off-diagonals indicating load transfer between nodes. The physical interpretation aligns with the embodying resistance to relative nodal , analogous to spring constants in systems.

Specific cases

Poisson equation

The Poisson equation, a fundamental , models fields such as in conduction or electrostatic potential, and is given by -\nabla \cdot (k \nabla u) = f in a domain \Omega, where k > 0 is a coefficient, u is the unknown , and f is a source term. To apply the (FEM), the equation is first reformulated in its weak form by multiplying by a test function \phi \in H^1_0(\Omega) (the of functions with square-integrable first derivatives vanishing on the boundary for homogeneous Dirichlet conditions) and integrating by parts: \int_\Omega k \nabla \phi \cdot \nabla u \, d\Omega = \int_\Omega f \phi \, d\Omega + \int_{\partial \Omega_N} g \phi \, d\Gamma, where \partial \Omega_N denotes the Neumann boundary portion and g is the specified flux. This bilinear form on the left-hand side ensures well-posedness under suitable conditions on k. In the Galerkin FEM discretization, the solution is approximated as u_h = \sum_{j=1}^N u_j \phi_j, where \{\phi_j\} are basis functions spanning a finite-dimensional subspace V_h \subset H^1_0(\Omega), such as continuous piecewise polynomials on a mesh. Substituting into the weak form yields the discrete system K \mathbf{u} = \mathbf{F}, where \mathbf{u} collects the nodal coefficients u_j, the load vector has entries F_j = \int_\Omega f \phi_j \, d\Omega + \int_{\partial \Omega_N} g \phi_j \, d\Gamma (involving a mass matrix M_{ij} = \int_\Omega \phi_i \phi_j \, d\Omega when f is projected), and the stiffness matrix K is symmetric positive definite with entries K_{ij} = \int_\Omega k \nabla \phi_i \cdot \nabla \phi_j \, d\Omega. This matrix encodes the diffusion operator and is the core of the FEM approximation for the Poisson problem, with its sparsity arising from local support of the basis functions. For two-dimensional problems on triangular meshes, linear Lagrange basis functions are commonly used, where on each element T_e with vertices \mathbf{r}_1, \mathbf{r}_2, \mathbf{r}_3 and area A_e, the local basis \phi_i|_{T_e} = \frac{1}{2A_e} (a_i + b_i x + c_i y) has constant gradients \nabla \phi_i = \frac{1}{2A_e} (b_i, c_i). Assuming constant k for simplicity, the element stiffness matrix K_e is then explicit: K_e = \frac{k}{4 A_e} \begin{pmatrix} |\mathbf{r}_2 - \mathbf{r}_3|^2 & (\mathbf{r}_2 - \mathbf{r}_3) \cdot (\mathbf{r}_3 - \mathbf{r}_1) & (\mathbf{r}_2 - \mathbf{r}_3) \cdot (\mathbf{r}_1 - \mathbf{r}_2) \\ (\mathbf{r}_2 - \mathbf{r}_3) \cdot (\mathbf{r}_3 - \mathbf{r}_1) & |\mathbf{r}_3 - \mathbf{r}_1|^2 & (\mathbf{r}_3 - \mathbf{r}_1) \cdot (\mathbf{r}_1 - \mathbf{r}_2) \\ (\mathbf{r}_2 - \mathbf{r}_3) \cdot (\mathbf{r}_1 - \mathbf{r}_2) & (\mathbf{r}_3 - \mathbf{r}_1) \cdot (\mathbf{r}_1 - \mathbf{r}_2) & |\mathbf{r}_1 - \mathbf{r}_2|^2 \end{pmatrix}, which is assembled into the global K by summing contributions from shared nodes. This form highlights the dependence on element and ensures second-order accuracy in the H^1-norm for smooth solutions. Boundary conditions influence the stiffness matrix differently based on type. Essential (Dirichlet) conditions, prescribing u = \bar{u} on \partial \Omega_D, are enforced by restricting V_h to functions satisfying them (e.g., setting corresponding rows and columns of K to zero except the diagonal to identity, and adjusting \mathbf{F}), preserving symmetry. Natural (Neumann or Robin) conditions, specifying flux k \nabla u \cdot \mathbf{n} = g or a convective form, appear only in \mathbf{F} via boundary integrals and do not modify K directly, allowing the matrix to remain unchanged while incorporating inhomogeneous data.

Elasticity and structural problems

In , the stiffness matrix plays a central role in the finite element formulation of problems, which govern the deformation of deformable bodies under external loads. The fundamental equilibrium equation is \nabla \cdot \boldsymbol{\sigma} = \mathbf{f}, where \boldsymbol{\sigma} is the tensor, \mathbf{f} is the , and the stress-strain relation follows \boldsymbol{\sigma} = \mathbf{D} \boldsymbol{\varepsilon}, with the infinitesimal strain tensor defined as \boldsymbol{\varepsilon} = \frac{1}{2} (\nabla \mathbf{u} + (\nabla \mathbf{u})^T). Here, \mathbf{u} is the displacement , and \mathbf{D} is the constitutive (elasticity) that encapsulates material properties such as E and \nu. This vector-valued framework distinguishes elasticity from scalar problems, as it accounts for coupled multi-directional deformations and components. Within the finite element method, the element stiffness matrix \mathbf{K}_e for linear elasticity is derived from the principle of virtual work and expressed as the volume integral \mathbf{K}_e = \int_V \mathbf{B}^T \mathbf{D} \mathbf{B} \, dV, where \mathbf{B} is the strain-displacement matrix relating strains to nodal displacements. The matrix \mathbf{B} incorporates the gradients of the shape functions and explicitly includes terms for normal strains (\varepsilon_{xx}, \varepsilon_{yy}, \varepsilon_{zz}) and shear strains (\varepsilon_{xy}, \varepsilon_{xz}, \varepsilon_{yz}), ensuring symmetry in 2D and 3D formulations; for instance, in 2D, \mathbf{B} is a $3 \times 8 matrix for a 4-node element, with rows corresponding to \varepsilon_{xx} = \sum \frac{\partial N_i}{\partial x} u_i, \varepsilon_{yy} = \sum \frac{\partial N_i}{\partial y} v_i, and \varepsilon_{xy} = \frac{1}{2} \sum \left( \frac{\partial N_i}{\partial y} u_i + \frac{\partial N_i}{\partial x} v_i \right), where N_i are the shape functions and u_i, v_i are nodal displacements. The integral is typically evaluated numerically via Gauss quadrature to handle arbitrary geometries. For two-dimensional approximations, and plane strain assumptions simplify the 3D problem by reducing the dimensionality while preserving key . In , applicable to thin plates where out-of-plane stresses vanish (\sigma_{zz} = \sigma_{xz} = \sigma_{yz} = 0), the reduced constitutive matrix is \mathbf{D} = \frac{E}{1 - \nu^2} \begin{bmatrix} 1 & \nu & 0 \\ \nu & 1 & 0 \\ 0 & 0 & \frac{1 - \nu}{2} \end{bmatrix}, relating in-plane stresses \{\sigma_{xx}, \sigma_{yy}, \sigma_{xy}\} to strains. Conversely, plane strain, suitable for long bodies constrained in the z-direction (\varepsilon_{zz} = \varepsilon_{xz} = \varepsilon_{yz} = 0), uses \mathbf{D} = \frac{E}{(1 + \nu)(1 - 2\nu)} \begin{bmatrix} 1 - \nu & \nu & 0 \\ \nu & 1 - \nu & 0 \\ 0 & 0 & \frac{1 - 2\nu}{2} \end{bmatrix}. These forms adjust the effective to reflect the assumptions, enabling efficient 2D simulations without full 3D meshing. A representative example is the 4-node quadrilateral for 2D , commonly used in or analyses due to its and ability to model irregular shapes via isoparametric mapping. The shape functions in natural coordinates (\xi, \eta) are N_i = \frac{1}{4} (1 + \xi_i \xi)(1 + \eta_i \eta), leading to a \mathbf{B} matrix that varies within the . For a rectangular of width $2a and height $2b (with coordinates aligned), the position-dependent \mathbf{B} matrix is \mathbf{B} = \begin{bmatrix} -\frac{b - y}{4ab} & 0 & \frac{b - y}{4ab} & 0 & \frac{b + y}{4ab} & 0 & -\frac{b + y}{4ab} & 0 \\ 0 & -\frac{a - x}{4ab} & 0 & -\frac{a + x}{4ab} & 0 & \frac{a + x}{4ab} & 0 & \frac{a - x}{4ab} \\ -\frac{a - x}{4ab} & -\frac{b - y}{4ab} & -\frac{a + x}{4ab} & \frac{b - y}{4ab} & \frac{a + x}{4ab} & \frac{b + y}{4ab} & \frac{a - x}{4ab} & -\frac{b + y}{4ab} \end{bmatrix}, where x, y are local coordinates within [-a, a] \times [-b, b]; the full \mathbf{K}_e = t \int_{-a}^{a} \int_{-b}^{b} \mathbf{B}^T \mathbf{D} \mathbf{B} \, dx \, dy can be computed analytically for uniform \mathbf{D} to yield an explicit $8 \times 8 , or numerically via Gauss quadrature to account for the varying \mathbf{B}. This captures shear locking if not integrated properly, typically requiring reduced (1-point) or selective integration.

Assembly and computation

Element-level construction

The construction of the local stiffness matrix for a finite element begins with evaluating the integral derived from the weak form of the governing equations, which relates the element's contributions to the overall system through shape functions and material properties. This process is performed in the element's local coordinate system to isolate computations before global assembly. For linear elastic materials, the element stiffness matrix \mathbf{k}^e is computed as \mathbf{k}^e = \int_{V^e} \mathbf{B}^T \mathbf{D} \mathbf{B} \, dV, where \mathbf{B} is the strain-displacement matrix, \mathbf{D} is the constitutive matrix, and the integration is over the element volume V^e. This integral is typically evaluated numerically using Gauss-Legendre quadrature, which approximates the integrand at specific Gauss points \xi_i with weights w_i, ensuring exact integration for polynomials up to a certain degree matching the element's order. For a 2D quadrilateral element, a 2x2 Gauss quadrature rule (four points) is often sufficient for bilinear interpolation, transforming the integral to \mathbf{k}^e \approx \sum_{i=1}^{n_g} w_i \mathbf{B}(\xi_i)^T \mathbf{D} \mathbf{B}(\xi_i) \det(\mathbf{J}(\xi_i)), where n_g is the number of Gauss points and \mathbf{J} is the Jacobian of the coordinate transformation. In isoparametric elements, which use the same shape functions for geometry and displacements, the physical coordinates are mapped from a reference (parent) element via \mathbf{x} = \sum N_i(\xi) \mathbf{x}_i, requiring the Jacobian matrix \mathbf{J} = \partial \mathbf{x}/\partial \xi to account for distortions in curved or irregular elements; its determinant scales the differential volume in the integral. This transformation ensures accurate representation of arbitrary shapes while maintaining computational efficiency. While the above focuses on linear problems, extensions to material nonlinearity involve the tangent stiffness matrix, which updates \mathbf{D} to the current or tangent modulus at each of a nonlinear solver, preserving the form but requiring re-evaluation per load step. As a specific example, consider a bilinear in 2D for the Poisson equation -\nabla^2 u = f, where the stiffness matrix simplifies to \mathbf{k}^e = \int_{\Omega^e} \nabla \mathbf{N}^T \nabla \mathbf{N} \, d\Omega with constitutive matrix. For a rectangular element of width a and b, the shape function derivatives in local coordinates yield explicit entries after 2x2 Gauss ; for instance, the (1,1) entry is \frac{1}{3} \left( \frac{a}{b} + \frac{b}{a} \right), but numerical at points like (\pm 1/\sqrt{3}, \pm 1/\sqrt{3}) with weights 1 handles general cases efficiently.

Global matrix assembly

The global stiffness matrix in the is constructed by superimposing the contributions from each element's stiffness matrix, based on the , to form a single matrix that governs the entire discretized domain. This assembly process, central to the , ensures kinematic compatibility across element boundaries and force at shared nodes. Connectivity between elements is defined by node numbering in the mesh, which maps local degrees of freedom in each element to global indices. During assembly, a scatter-gather operation distributes the local stiffness matrix entries to their corresponding global positions and sums overlapping contributions; specifically, for an element's local entry K_e(m,n), the global update is K_\text{global}(i,j) += K_e(m,n), where i and j are the global node numbers linked to local indices m and n via the element's array. The element stiffness matrices themselves are typically obtained from over the element domain. The assembled global stiffness matrix exhibits a sparse , with non-zero entries limited to connected through the , reflecting the localized interactions in the physical system. To enhance computational in and solution phases, node renumbering techniques, such as the Cuthill-McKee algorithm, are applied to minimize the matrix —the maximum difference in node indices for non-zero off-diagonal terms—thereby reducing fill-in during . The global load vector is assembled analogously, by scattering and summing element-level force contributions (from distributed loads or body forces) to the appropriate global based on . Consider a simple one-dimensional example: a discretized into three equal-length elements with four nodes (1, 2, 3, 4), assuming uniform axial k = \frac{AE}{L} for each element, where A is the cross-sectional area, E is the , and L is the element length. Each element's local stiffness matrix is K_e = k \begin{bmatrix} 1 & -1 \\ -1 & 1 \end{bmatrix}, with local nodes mapped as follows: element 1 (nodes 1–2), element 2 (nodes 2–3), element 3 (nodes 3–4).
  • For element 1, scatter to global rows/columns 1 and 2: adds k to K_\text{global}(1,1) and K_\text{global}(2,2), -k to K_\text{global}(1,2) and K_\text{global}(2,1).
  • For element 2, scatter to global rows/columns 2 and 3: adds k to K_\text{global}(2,2) and K_\text{global}(3,3), -k to K_\text{global}(2,3) and K_\text{global}(3,2); note the summation at row/column 2.
  • For element 3, scatter to global rows/columns 3 and 4: adds k to K_\text{global}(3,3) and K_\text{global}(4,4), -k to K_\text{global}(3,4) and K_\text{global}(4,3); summation occurs at row/column 3.
The resulting 4×4 global stiffness matrix is K_\text{global} = k \begin{bmatrix} 1 & -1 & 0 & 0 \\ -1 & 2 & -1 & 0 \\ 0 & -1 & 2 & -1 \\ 0 & 0 & -1 & 1 \end{bmatrix}, illustrating the banded sparse pattern with 2. If nodal forces are present, such as a distributed load on 2 yielding a local force \{f_e\} = \frac{qL}{2} \{1, 1\}^T (where q is load per unit length), it scatters equally to global 2 and 3, with similar additions for other elements.

Solution and applications

Solving the linear system

Once the global stiffness matrix K and the force vector F have been assembled, the finite element method requires solving the Ku = F for the displacement vector u, where K is typically symmetric positive definite and sparse. This system can be large, with millions of in practical simulations, necessitating efficient numerical techniques that exploit these properties. Direct methods, such as and , are suitable for small to medium-sized systems where exact solutions are feasible. transforms the system into upper triangular form through row operations, allowing back-substitution to find u, and is effective for dense matrices but computationally intensive for large sparse ones due to fill-in. factors K into a lower L and an upper U such that K = LU, enabling forward and backward substitution for multiple right-hand sides; is a variant for symmetric positive definite matrices like K, reducing operations by about half. These methods provide high accuracy but scale poorly, with complexity O(n^3) for an n \times n matrix, making them impractical for systems exceeding a few thousand without specialized sparse implementations. For large-scale problems, iterative methods are preferred, as they approximate solutions through successive refinements and leverage sparsity. The method is widely used for symmetric positive definite K, generating a sequence of residuals orthogonal to previous search directions to converge to u in at most n steps theoretically, though practical is faster for well-conditioned matrices. Preconditioning improves CG performance by solving a modified system \tilde{K} \tilde{u} = \tilde{F} where \tilde{K} = M^{-1} K and M approximates K, reducing the and iteration count; common preconditioners include or multigrid approaches. These methods are memory-efficient and parallelizable, often converging in O(\sqrt{\kappa}) iterations where \kappa is the condition number of K. The condition number \kappa(K) = \|K\| \cdot \|K^{-1}\| (typically the 2-norm of largest to smallest eigenvalues) measures the of the solution to perturbations and affects solver and . In finite element discretizations, ill-conditioning arises from high aspect ratios in mesh elements, which distort the of K and amplify round-off errors, potentially leading to \kappa exceeding $10^{10} and requiring or refinement. Poor aspect ratios, such as elongated elements, introduce near-zero eigenvalues, exacerbating this issue and slowing iterative solvers. Consider a simple two-degree-of-freedom (2-DOF) system of two springs in series, each with stiffness k = 1 (units omitted), where the left end is fixed and the right end is free, with an applied force F_2 = 1 at the second node; the global stiffness matrix is K = \begin{bmatrix} 2 & -1 \\ -1 & 1 \end{bmatrix}, and F = \begin{bmatrix} 0 \\ 1 \end{bmatrix}. Using Gaussian elimination, pivot on K_{11} = 2, with multiplier m = -1/2 = -0.5; the updated K_{22}' = 0.5 and F_2' = 1; back-substitution gives u_2 = 2, u_1 = 1. This explicit solution illustrates the displacements, with the second node shifting by twice the amount of the first node due to the series setup.

Engineering and computational uses

The stiffness matrix, a cornerstone of the direct stiffness method, was developed in the 1950s and 1960s by M.J. Turner and colleagues at Boeing for analyzing complex aerospace structures, enabling efficient computation of displacements and stresses in aircraft components that traditional methods struggled to handle. In structural analysis, stiffness matrices form the basis of finite element analysis (FEA) in software such as ANSYS, where they are assembled to simulate load-bearing behavior in civil engineering applications like bridges and mechanical structures like aircraft frames. For instance, ANSYS uses the global stiffness matrix to predict deflections and stresses under operational loads, ensuring designs meet safety standards for long-span bridges subjected to wind and traffic forces. In aerospace, this approach optimizes aircraft fuselages by evaluating stiffness contributions from composite grid-stiffened panels, balancing weight and rigidity. For and , the stiffness matrix extends to dynamic analyses by coupling with mass matrices in , solving eigenvalue problems to identify natural frequencies and mode shapes in engineering systems. This is critical for assessing risks in structures like turbine blades or vehicle suspensions, where the generalized form \mathbf{K} \phi = \omega^2 \mathbf{M} \phi reveals modes that could lead to fatigue failure if excited. In structural optimization, of the stiffness matrix with respect to design variables—such as material thickness or geometry—guides iterative improvements to enhance performance metrics like weight reduction while maintaining . Techniques compute derivatives of stiffness matrix elements to evaluate how perturbations affect overall system responses, as implemented in optimization procedures for aerospace components.

References

  1. [1]
    [PDF] Mathematical Properties of Stiffness Matrices - Duke People
    Element stiffness matrices [k] relate the forces and displacements of all of the coordinates of the element, regardless of whether or not any of the coordinates ...
  2. [2]
    [PDF] Stiffness Methods for Systematic Analysis of Structures (Ref
    Apr 30, 2014 · The Stiffness method provides a very systematic way of analyzing determinate and indeterminate structures. Displacement (Stiffness) Method. ...
  3. [3]
    [PDF] History of the stiffness method
    In the 1930s, under the it became fashionable to use matrix forms to illustrate the whole process systematically. [21–23]. Similar equations occur in the ...
  4. [4]
    [PDF] Introduction to Linear Elasticity - UCI Mathematics
    Mar 25, 2024 · ABSTRACT. This notes introduces the theory of linear elasticity, which studies the de- formation of elastic solid bodies under external ...
  5. [5]
    [PDF] 17. An introduction to the finite element method
    c) The stiffness matrix is of a “tridiagonal” structure: the major diagonal has one super- diagonal and one sub-diagonal, with the remaining elements of the ...
  6. [6]
    [PDF] Chapter 2 – Introduction to the Stiffness (Displacement) Method
    Derive the Element Stiffness Matrix and Equations -. Define the stiffness ... Potential Energy Approach to Derive Spring. Element Equations. One of the ...
  7. [7]
    [PDF] Finite Element Analysis - MIT
    Feb 28, 2001 · The quantity in brackets, multiplied by AE/L, is known as the “element stiffness matrix” kij. Each of its terms has a physical significance, ...
  8. [8]
    [PDF] Stiffness Method - Professor Terje Haukaas
    Jan 6, 2023 · ... Kij = force along DOF number i due to a unit displacement or rotation along DOF number j, uj = unknown displacement or rotation along DOF ...
  9. [9]
    [PDF] The Matrix Stiffness Method for 2D Frames - Duke People
    You should understand where these equations come from, why this matrix is symmetric, why the diagonal terms are all positive, and what the off-diagonal terms ...
  10. [10]
    [PDF] Development of Truss Equations
    Having set forth the foundation on which the direct stiffness method is based, we will now derive the stiffness matrix for a linear-elastic bar (or truss) ...
  11. [11]
    The Finite Element Method: Its Basis and Fundamentals
    This volume presents a view of the finite element method as a general discretization procedure of continuous systems.
  12. [12]
    [PDF] Lecture Notes on Finite Element Methods for Partial Differential ...
    Jan 19, 2025 · Synopsis: Finite element methods represent a powerful and general class of techniques for the approximate solution of partial differential ...
  13. [13]
    [PDF] 4 Plane stress and plane strain - Unife
    The derived formulae enable the full stiffness matrix of the structure to be assembled, and a solution for displacements to be obtained. The stress matrix given ...
  14. [14]
    Engineering at Alberta Courses » Quadrilateral Elements
    Shape functions distribution on the quadratic quadrilateral. Stiffness Matrix. The constitutive relationship of plane linear elastic materials is defined using ...
  15. [15]
    [PDF] Finite Element Method III: - OSTI.GOV
    This can be evaluated using a product Gauss-Legendre quadrature formula where are the number of quadrature points, are the quadrature points in the interval ...
  16. [16]
    [PDF] YALEU/DCS/TR-743 - Department of Computer Science
    is performed by Gauss quadrature, wherein the finite element is first mapped onto a standard element of regular shape. The computation of an elemental stiffness ...
  17. [17]
    [PDF] Issues of the FE Method in one space dimension
    The integration formulas is the Gaussian quadrature of order one, two, three, or four. • The matrix assembly is element by element. 8.4.1 Gaussian quadrature ...
  18. [18]
    [PDF] CVEN 4511/5511: Introduction to Finite Element Analysis
    We use normal Gauss integration for shear stiffness matrix ¯k e. , and reduced integration for volumetric stiffness k e . This is called “selective reduced ...
  19. [19]
    [PDF] Isoparametric Elements Chapter 10
    First, we will illustrate the isoparametric formulation to develop the simple bar element stiffness matrix. Use of the bar element makes it relatively easy to ...
  20. [20]
    [PDF] CHAP 7 ISOPARAMETRIC ELEMENTS
    – Jacobian matrix is constant throughout the element. – Jacobian matrix ... – [k(e)] is the element stiffness matrix. – Integration domain is a general ...
  21. [21]
    [PDF] Formulation and Calculation of Isoparametric Finite Element Matrices
    ' To be able to evaluate the stiffness matrix of an element, we need to calculate the strain-displacement transformation matrix. The ...
  22. [22]
    [PDF] Chapter 2 Nonlinear Finite Element Formulation for ... - VTechWorks
    nonlinear stiffness matrix. The second source of nonlinearity ... where U is the potential energy of deformation and W is the potential energy of the.
  23. [23]
    Tangent stiffness properties of finite elements - ScienceDirect.com
    The new tangent stiffness matrices developed in this study represent the final conditions of the material, such as the most recent deformed geometry, the ...
  24. [24]
    [PDF] Formulation of Finite Element Method for 1D and 2D Poisson Equation
    The element will have two trial functions and so we make. 2X2 local stiffness matrices. These local stiffness matrices are then combined to produce the global ...
  25. [25]
    Stiffness and Deflection Analysis of Complex Structures - AIAA ARC
    Stiffness and Deflection Analysis of Complex Structures. M. J. TURNER,; R. W. CLOUGH,; H. C. MARTIN and; L. J. TOPP. M. J. TURNER. Search for more papers by ...
  26. [26]
    Direct stiffness method and the global stiffness matrix - DoITPoMS
    The spring stiffness equation relates the nodal displacements to the applied forces via the spring (element) stiffness.
  27. [27]
    [PDF] A Two-Step Approach to Finite Element Ordering. - DTIC
    The finite element mesh can be transformed directly into the finite graph G representing the structure of the global stiffness matrix. We call this graph a ...
  28. [28]
  29. [29]
    How to Choose a Solver for FEM Problems: Direct or Iterative?
    Mar 13, 2024 · More commonly adopted approaches to directly calculate the solution include Gaussian elimination, LU, Cholesky, and QR decompositions. However, ...Missing: decomposition | Show results with:decomposition
  30. [30]
    [PDF] Introduction to the Finite Element Method (FEM) Lecture 1 The Direct ...
    For a more complex spring system, a 'global' stiffness matrix is required – i.e. one that describes the behaviour of the complete system, and not just the ...Missing: interpretation | Show results with:interpretation<|control11|><|separator|>
  31. [31]
    Getting Started with Robust Finite Element Method and Solvers
    The Gauss elimination method and the LU or Cholesky factorization method, are examples of direct solvers. Iterative solvers initialize uwith approximate values ...Missing: decomposition | Show results with:decomposition
  32. [32]
    [PDF] An Introduction to the Conjugate Gradient Method Without the ...
    Aug 4, 1994 · The Conjugate Gradient Method is the most prominent iterative method for solving sparse systems of linear equations.
  33. [33]
    Preconditioned conjugate gradient method for finding minimal ...
    We present an iterative proposal, based on the preconditioned conjugate gradient method, to solve the linear system associated to the problem of approximating ...
  34. [34]
    [PDF] The method of conjugate gradients in finite element applications
    The basically iterative procedure was devised in order to take into account the sparsity of the matrix A to full extent and to operate with the originally given ...
  35. [35]
    [PDF] Conditioning of Finite Element Equations with Arbitrary Anisotropic ...
    Sep 18, 2012 · First, we develop tight bounds on the extreme eigenvalues and the condition number of the stiffness matrix for a general diffusion problem with ...
  36. [36]
    [PDF] What Is a Good Linear Finite Element? Interpolation, Conditioning ...
    Dec 31, 2002 · Poorly conditioned matrices affect linear equation solvers by slowing them down or introducing large roundoff errors into their results.Missing: ill- | Show results with:ill-
  37. [37]
    [PDF] The Origins of the Finite Element Method
    there is no question in the writer's mind: M. J. (Jon) Turner at Boeing over the period 1950–1962. He generalized and perfected the Direct Stiffness Method, and ...
  38. [38]
    [PDF] Structural Analysis Guide - Ansys Help
    ... Structural Analysis Overview. Structural analysis is probably the most common application of the finite element method. The term structural (or structure) ...
  39. [39]
    A consistent dynamic stiffness matrix for flutter analysis of bridge decks
    Oct 1, 2023 · Thus, the proposed dynamic stiffness matrix approach is effective for global analysis of the bridge, such as free vibration and flutter analysis ...
  40. [40]
    finite element analysis, optimum design and cost-effective ...
    The main purposes of this application for structures are high stiffness to weight ratio, increase of bending stiffness preventing from local flutter, vibration ...
  41. [41]
    Lecture 24: Modal Analysis: Orthogonality, Mass Stiffness, Damping ...
    Description: Prof. Vandiver goes over the modal expansion theorem, computing mass and stiffness matrices, obtaining uncoupled equations of motion, ...
  42. [42]
    [PDF] An Algorithm for Synthesizing Mass and Stiffness Matrices From ...
    The algorithm synthesizes mass and stiffness matrices from experimental vibration data, using an improved technique to add structural modifications.
  43. [43]
    Structural Sensitivity Analysis and Optimization 1 - SpringerLink
    Structural design sensitivity analysis concerns the relationship between design variables available to the design engineer and structural responses ...
  44. [44]
    [PDF] Documentation for a Structural Optimization Procedure Developed ...
    The SECT INIT procedure is used to compute stiffness submatrices for both the original and perturbed values of each design variable. The derivative of the ...