Fact-checked by Grok 2 weeks ago

Euler–Bernoulli beam theory

The Euler–Bernoulli beam theory, also known as classical beam theory, is a foundational simplification of theory used to predict the deflection, , and internal forces in slender beams subjected to transverse loads. It models the beam as a one-dimensional along its longitudinal , focusing on deformation while neglecting effects like shear distortion and rotational inertia. Developed in the mid-18th century by Swiss mathematicians Leonhard Euler and , the theory emerged from early efforts to mathematically describe elastic curves and vibrations in , building on prior incomplete attempts by figures like . Euler contributed analyses of beam shapes under boundary conditions, while Bernoulli derived key differential equations for dynamic behavior. This framework became a cornerstone of engineering mechanics, enabling practical calculations for structures long before widespread computational tools. Central to the theory are kinematic assumptions that the beam material is linearly , homogeneous, and isotropic, with the length significantly greater than cross-sectional dimensions for slenderness. Specifically, plane cross-sections to the undeformed remain plane and after deformation, implying no strain in the cross-section and uniform longitudinal varying linearly with distance from the . These lead to the field: axial u(x, y, z) = -z \frac{dw}{dx} and transverse w(x, y, z) = w(x), where w(x) is the deflection of the . The governing equation for static bending under distributed load q(x) and constant flexural rigidity EI (product of E and I) is EI \frac{d^4 w}{dx^4} = q(x). For free vibration, it extends to EI \frac{\partial^4 w}{\partial x^4} + \rho A \frac{\partial^2 w}{\partial t^2} = 0, where \rho is and A is cross-sectional area, yielding natural frequencies and mode shapes. Bending stress follows from \sigma_x = -E z \frac{d^2 w}{dx^2}. Widely applied in civil, mechanical, and for designing beams in bridges, frames, and machine components, the theory provides accurate results for slender structures where dominates. However, it overestimates in shorter or thicker beams due to ignored effects, prompting refinements like the Timoshenko beam theory in the early to account for deformation and rotary .

Fundamentals

Assumptions and limitations

The Euler–Bernoulli beam theory traces its origins to the foundational contributions of Leonhard Euler in the 1740s, who published key derivations of the for the elastic curve of a deflected in his 1744 work Methodus Inveniendi Lineas Curvas Maximi Minive Proprietate Gaudentes, and , who proposed solving the general elastica problem to Euler in 1742, with the theory coalescing around 1750 based on material elasticity linking bending moments to deflections. These early works established the theoretical framework for analyzing bending under transverse loads, emphasizing kinematic simplifications that enable closed-form solutions. The theory relies on several core assumptions to simplify the complex three-dimensional elasticity problem into a one-dimensional model. The must be slender, with its significantly greater than its cross-sectional dimensions (typically length-to-depth ratio exceeding 10), ensuring that dominates over other deformation modes. Deflections are assumed small relative to the , permitting linear strain-displacement relationships and neglecting higher-order geometric effects. A fundamental kinematic states that plane cross-sections perpendicular to the beam's before deformation remain plane and perpendicular after deformation, implying negligible transverse deformation and of cross-sections due solely to . The stress state is uniaxial, with normal stresses varying linearly across the cross-section and no significant transverse or stresses considered in the primary formulation. Additionally, the material is assumed homogeneous and isotropic, exhibiting linear elastic behavior with constant properties such as throughout the . These assumptions limit the theory's applicability to scenarios where they hold true. It becomes inaccurate for short or thick beams, where deformation significantly influences deflections, often requiring alternatives like Timoshenko beam theory that incorporate effects. For large deflections comparable to the beam depth, geometric nonlinearities arise, invalidating the linear approximations and necessitating nonlinear extensions of the theory. The homogeneous and isotropic material assumption restricts use to uniform, single-material beams, rendering the theory unsuitable without modifications for composites, laminates, or anisotropic materials where properties vary directionally. In dynamic contexts, the theory supports linear vibration analysis but fails for high-amplitude or impact-induced responses involving nonlinear effects or significant rotatory inertia.

Kinematics of deformation

In Euler–Bernoulli beam theory, the of deformation describes the geometric relationships between the beam's , , and under , assuming that plane cross-sections perpendicular to the beam remain plane and perpendicular after deformation. This kinematic model simplifies the three-dimensional deformation to a one-dimensional problem along the beam length, focusing on the transverse deflection of the . The field is defined with the transverse w(x) representing the deflection of the 's centerline in the perpendicular to the axis, while the axial varies linearly across the cross-section as u(x,y) = -y \frac{dw}{dx}, where x is the position along the and y is the distance from the in the height . There is no in the lateral , so v(x,y,z) = 0. This formulation captures the extension or contraction of fibers away from the due to bending rotation. The rotation of the cross-section \theta(x) is given by the slope of the deflected centerline, \theta(x) = \frac{dw}{dx}, which holds under the where \sin \theta \approx \tan \theta \approx \theta. This rotation implies that points on the cross-section move axially proportional to their distance from the , leading to the linear variation in axial displacement. The longitudinal normal strain \varepsilon_x arises from the derivative of the axial displacement and is expressed as \varepsilon_x = \frac{\partial u}{\partial x} = -y \frac{d^2 w}{dx^2}, indicating that only normal strains in the x-direction are considered, with shear strains neglected. This strain distribution is linear through the thickness, positive (tensile) above the neutral axis for concave-up deflection and compressive below it. The theory focuses exclusively on this longitudinal strain component for bending analysis. The \kappa(x) quantifies the bending deformation and is defined as the second of the transverse , \kappa(x) = \frac{d^2 w}{dx^2}, which for small deflections approximates the change in slope per unit . This measure directly relates to via \varepsilon_x = -y \kappa, providing a geometric interpretation of how the beam's induces straining.

Governing Equations

Static equilibrium derivation

The derivation of the governing equation for static equilibrium in Euler–Bernoulli beam theory relies on applying force and balances to a differential element of the beam, assuming small deflections and neglecting axial forces for cases. Consider a beam segment of dx subjected to a transverse distributed load q(x) acting in the direction of deflection w(x). The internal V(x) at position x and V(x + dx) at x + dx, along with the external load, must balance in the transverse direction. Summing forces yields V(x) - V(x + dx) + q(x) \, dx = 0, which simplifies to the equilibrium relation \frac{dV}{dx} = q(x). Moment equilibrium for the same element, taking moments about the right face and neglecting higher-order terms, gives M(x) - M(x + dx) + V(x) \, dx = 0, leading to \frac{dM}{dx} = V(x), where M(x) is the internal . These relations connect the load, , and moment without considering time-dependent inertia effects. To relate these to deflection, the theory employs the constitutive moment-curvature relation derived from and the kinematic assumption that plane sections remain plane and perpendicular to the . The beam curvature \kappa(x) is approximated as \kappa(x) = \frac{d^2 w}{dx^2} for small slopes. The is then M(x) = EI \frac{d^2 w}{dx^2}, where E is and I is the second moment of area about the . Substituting into the moment equilibrium equation produces V(x) = \frac{dM}{dx} = EI \frac{d^3 w}{dx^3} for constant EI. Differentiating again and using the shear equilibrium yields \frac{dV}{dx} = EI \frac{d^4 w}{dx^4} = q(x), or the governing differential equation EI \frac{d^4 w}{dx^4} = q(x). For beams with spatially varying flexural rigidity EI(x), successive differentiation generalizes this to \frac{d^2}{dx^2} \left( EI(x) \frac{d^2 w}{dx^2} \right) = q(x). This fourth-order equation describes the static deflection under transverse loading, solved subject to boundary conditions.

Dynamic equilibrium equation

The dynamic equilibrium equation in Euler–Bernoulli beam theory incorporates inertial effects to describe the time-dependent transverse motion of the beam under distributed loading. This equation is derived by applying Newton's second law or to the force balance on a differential beam element, extending the static relations by including the mass acceleration term. Consider an infinitesimal beam element of length dx at position x, with transverse displacement w(x,t). The shear forces acting on the element are V(x) and V(x + dx), the external distributed load per unit length is q(x,t), and the inertial term accounts for acceleration in the direction of positive w. The transverse force balance yields: V(x) - V(x + dx) + q(x,t) \, dx = \rho A \frac{\partial^2 w}{\partial t^2} dx Dividing through by dx and taking the limit as dx \to 0 gives the relation between the shear force gradient and the net transverse loading, including inertia: -\frac{\partial V}{\partial x} + q(x,t) = \rho A \frac{\partial^2 w}{\partial t^2} or equivalently, \frac{\partial V}{\partial x} = q(x,t) - \rho A \frac{\partial^2 w}{\partial t^2}. This treats the inertial term \rho A \frac{\partial^2 w}{\partial t^2} consistently with the external loading direction. To obtain the governing , relate the V to the M via moment equilibrium: V = \frac{\partial M}{\partial x}. Under the Euler–Bernoulli assumptions, the moment is linked to the by M = E I \frac{\partial^2 w}{\partial x^2}, where E is and I the second moment of area (assumed constant for simplicity). Substituting and differentiating twice with respect to x results in the full equation: \frac{\partial^2}{\partial x^2} \left( [E](/page/E!) I \frac{\partial^2 [w](/page/W)}{\partial x^2} \right) + \rho A \frac{\partial^2 [w](/page/W)}{\partial t^2} = [q](/page/Q)(x,t). For uniform beams where [E](/page/E!) and I are constant, this simplifies to [E](/page/E!) I \frac{\partial^4 [w](/page/W)}{\partial x^4} + \rho A \frac{\partial^2 [w](/page/W)}{\partial t^2} = [q](/page/Q)(x,t). This fourth-order governs the linear transverse vibrations and forced responses of the . The derivation relies on key linearity assumptions, including small-amplitude deflections to ensure the remains \frac{\partial^2 w}{\partial x^2} without geometric nonlinearities, and neglect of effects in the basic formulation (viscous or structural can be added later as c \frac{\partial w}{\partial t}). These maintain the validity of the uncoupled, linear relations between , , and in the theory.

Vibration Theory

Free vibration analysis

Free vibration analysis of beams within the Euler–Bernoulli framework involves solving the homogeneous dynamic equation under zero external loading, yielding the natural frequencies and mode shapes that characterize the system's undamped oscillatory behavior. The process begins with the assumption of a separable solution for the transverse displacement w(x,t) = \phi(x) T(t), where \phi(x) represents the spatial mode shape and T(t) the temporal component. Substituting this form into the governing partial differential equation \frac{\partial^2}{\partial x^2} \left( EI \frac{\partial^2 w}{\partial x^2} \right) + \rho A \frac{\partial^2 w}{\partial t^2} = 0 and dividing through by \phi(x) T(t) leads to the separation of variables, resulting in \frac{EI \phi''''(x)}{\rho A \phi(x)} = -\frac{\ddot{T}(t)}{T(t)} = \omega^2, where \omega is the constant natural frequency (separation parameter). The temporal equation \ddot{T}(t) + \omega^2 T(t) = 0 yields harmonic solutions T(t) = \cos(\omega t + \psi), while the spatial ordinary differential equation simplifies to \phi''''(x) - \beta^4 \phi(x) = 0, with \beta^4 = \frac{\rho A \omega^2}{EI}. The general solution to this fourth-order equation is \phi(x) = A \cos(\beta x) + B \sin(\beta x) + C \cosh(\beta x) + D \sinh(\beta x), where the coefficients A, B, C, D are determined by applying the specific boundary conditions at the beam ends. Imposing the boundary conditions on \phi(x) and its derivatives results in a homogeneous system of four equations in the coefficients, which has nontrivial solutions only when the of the vanishes, yielding the (or frequency) equation that determines the eigenvalues \beta_n L (for beam length L) and thus the natural frequencies \omega_n = \beta_n^2 \sqrt{\frac{EI}{\rho A}} for each mode n. The corresponding eigenfunctions \phi_n(x) form a complete orthogonal set with respect to the weighting function \rho A(x), satisfying \int_0^L \rho A \phi_m(x) \phi_n(x) \, dx = 0 for m \neq n, enabling modal decomposition for more complex analyses. For approximate estimation of natural frequencies without solving the exact , the provides an upper-bound variational method: \omega^2 \approx \frac{\int_0^[L](/page/L') EI (\phi''(x))^2 \, dx}{\int_0^[L](/page/L') \rho A \phi^2(x) \, dx}, where \phi(x) is a function satisfying the geometric boundary conditions; this approach, originally developed for acoustic problems, yields progressively accurate results with refined functions.

Cantilever beam modes

In the context of Euler–Bernoulli beam theory, the beam configuration features a fixed at one end and a free end at the other, making it a common model for structures like diving boards or aircraft wings. The boundary conditions reflect this setup: at the fixed end (x = 0), the transverse and are zero, given by w(0, t) = 0 and \frac{\partial w}{\partial x}(0, t) = 0; at the free end (x = ), the and vanish, expressed as \frac{\partial^2 w}{\partial x^2}([L](/page/L'), t) = 0 and \frac{\partial^3 w}{\partial x^3}([L](/page/L'), t) = 0. Applying these boundary conditions to the general solution for free vibration yields the \cos(\beta L) \cosh(\beta L) = -1, where \beta^4 = \frac{\rho A \omega^2}{EI}. This has infinitely many roots \beta_n L, with the first few being approximately 1.875104 for n=1, 4.694091 for n=2, and 7.854757 for n=3; for higher modes (n ≥ 4), \beta_n L \approx (2n - 1)\pi / 2. The corresponding natural frequencies are \omega_n = (\beta_n L)^2 \sqrt{\frac{EI}{\rho A L^4}}, where E is the , I the second moment of area, \rho the material density, and A the cross-sectional area. The mode shapes, representing the spatial deflection patterns, are \phi_n(x) = \cosh(\beta_n x) - \cos(\beta_n x) - \sigma_n [\sinh(\beta_n x) - \sin(\beta_n x)], with \sigma_n = \frac{\cosh(\beta_n L) + \cos(\beta_n L)}{\sinh(\beta_n L) + \sin(\beta_n L)}; these are often normalized such that \int_0^L \phi_n^2(x) \, dx = L for in . Physically, the first mode shape exhibits smooth with maximum deflection at the free end and no nodes along the length, akin to fundamental bending. Higher modes introduce additional nodes (one for the second mode, two for the third), with deflection patterns showing increasing and steeper gradients near the fixed end, all governed by flexural deformation without or rotary inertia effects in the Euler–Bernoulli .

Simply supported beam modes

For a simply supported of length L, the boundary conditions impose zero transverse displacement and zero at both ends: w(0) = 0, w''(0) = 0, w(L) = 0, and w''(L) = 0. These conditions yield exact analytical solutions for the free vibration modes via . The wave numbers are \beta_n = n\pi / L for the nth , where n = 1, 2, 3, \dots. The corresponding shapes are \phi_n(x) = \sin(n\pi x / L), which are sinusoidal and symmetric about the beam's midpoint. The natural frequencies follow as \omega_n = (n\pi / L)^2 \sqrt{EI / \rho A}, where E is the , I is the area , \rho is the , and A is the cross-sectional area. The mode shapes exhibit a harmonic progression, with each successive introducing one additional half-wavelength along the length. Nodal points, where the is zero, occur at x = (k/n)L for integers k = 1, 2, \dots, n-1, resulting in n-1 nodes between the supports. This structure contrasts with asymmetric boundary conditions, such as in beams, where modes involve more complex transcendental functions. For higher modes, approximate methods like the Rayleigh-Ritz technique or finite element analysis provide good estimates for low-order frequencies but show increasing deviation from exact values as n grows, often requiring finer discretization or higher-order trial functions for accuracy. For instance, in Rayleigh-Ritz approximations using polynomial basis functions, errors in \omega_n can exceed 5% for n > 10 without sufficient terms.

Stress and Strain Analysis

Bending stress formula

In Euler–Bernoulli beam theory, the normal stress in the longitudinal direction, \sigma_x, resulting from pure bending is distributed linearly across the beam's cross-section according to the formula \sigma_x = -\frac{M y}{I} where M is the internal bending moment at the section, y is the perpendicular distance from the neutral axis to the point of interest, and I is the second moment of area of the cross-section about the neutral axis. This expression assumes the beam is subjected to transverse loads that produce bending without significant shear deformation, consistent with the theory's kinematic hypotheses. The derivation begins with the definition of the bending moment as the first moment of the normal stress distribution over the cross-sectional area: M = -\int_A \sigma_x y \, dA The negative sign accounts for the conventional sign convention where positive M causes compression above the neutral axis. To satisfy equilibrium under pure bending, the net axial force must be zero: \int_A \sigma_x \, dA = 0. The theory's assumption that plane cross-sections remain plane and perpendicular to the deformed beam axis after bending implies a linear variation of longitudinal strain, and by Hooke's law for linear elastic materials, \sigma_x = E \epsilon_x, this leads to a linear stress distribution: \sigma_x = -k y, where k is a constant of proportionality determined from kinematics and material properties. Substituting into the moment equation gives M = k I, so k = M / I, yielding the stress formula. The passes through the of the cross-section for homogeneous, isotropic beams, as the antisymmetric linear stress profile ensures zero and locates the zero-stress line at the geometric center. For a positive (typically causing the beam to sag), fibers above the neutral axis experience (\sigma_x < 0), while those below experience tensile stress (\sigma_x > 0). The maximum normal stress occurs at the extreme fibers of the cross-section, where |y| reaches its maximum value c (the distance from the neutral axis to the outermost fiber): \sigma_{\max} = \frac{M c}{I} This quantity, often called the flexural stress, is critical for assessing the beam's capacity to resist bending without yielding or failure.

Strain-curvature relationship

In the Euler–Bernoulli beam theory, the strain-curvature relationship arises from the kinematic assumptions that plane sections remain plane and perpendicular to the neutral axis after deformation, leading to a linear variation of longitudinal strain across the beam's cross-section. The axial strain \epsilon_x at a distance y from the neutral axis is given by \epsilon_x = -y \kappa, where \kappa is the curvature of the beam's deflected shape. This formula indicates that fibers above the neutral axis (positive y) experience compression (\epsilon_x < 0) while those below undergo elongation (\epsilon_x > 0), with the magnitude proportional to the distance y from the neutral axis. For small deflections, the \kappa approximates the second of the transverse deflection w(x) with respect to the axial coordinate x, such that \kappa \approx \frac{d^2 w}{dx^2}. Equivalently, since the or slope \theta \approx \frac{dw}{dx} for small angles, \kappa \approx \frac{d\theta}{dx}. This geometric approximation holds under the theory's assumption of negligible deformation and small slopes, ensuring that the strain distribution remains purely due to without transverse shear contributions. The basic Euler–Bernoulli formulation neglects transverse strains, such as those arising from the Poisson effect, treating the beam as a one-dimensional where deformation is dominated by axial extension or contraction along the length. This simplification focuses on the longitudinal strain field and omits lateral contractions, which is valid for slender beams where bending effects prevail over volumetric changes. The strain-curvature link connects local deformation to global deflection: integrating the curvature along the beam length yields the slope \theta(x) = \int \kappa \, dx + C_1, and further integration gives the deflection w(x) = \int \theta \, dx + C_2 x + C_3, where constants C_1, C_2, C_3 are determined by boundary conditions. This integration path underscores the theory's reliance on curvature as the fundamental measure of bending kinematics.

Material constitutive relations

In Euler–Bernoulli beam theory, the material constitutive relations establish the connection between internal stresses and strains, relying on the linear elastic behavior of the beam material. The core relation is , which for an isotropic, homogeneous material states that the axial normal stress \sigma_x is proportional to the axial normal strain \epsilon_x: \sigma_x = E \epsilon_x where E is the , a material property representing the under uniaxial or . This relation assumes small deformations within the elastic limit, where the material returns to its original shape upon load removal./03%3A_Development_of_Constitutive_Equations_of_Continuum%2C_Beams_and_Plates/3.04%3A_Hook%E2%80%99s_Law_in_Generalized_Quantities_for_Beams) When combined with the kinematic assumption of the theory—that plane cross-sections remain plane and perpendicular to the after deformation—the strain distribution across the beam height is linear, given by \epsilon_x = -y \kappa, where \kappa = \frac{d^2 w}{dx^2} is the and y is the from the . Substituting this into yields the stress distribution: \sigma_x = -E y \frac{d^2 w}{dx^2} This expression highlights how varies linearly through the thickness, with maximum values at the outer fibers. The assumption of ensures uniform properties in all directions within the plane of , and linearity precludes effects like yielding or permanent deformation. To relate this to the internal M, the is integrated over the cross-sectional area A: M = \int_A \sigma_x y \, dA = -E \frac{d^2 w}{dx^2} \int_A y^2 \, dA = EI \frac{d^2 w}{dx^2} Here, I = \int_A y^2 \, dA is the second moment of area about the , and EI represents the , a key quantifying the beam's resistance to . This moment-curvature relation derives directly from the constitutive law and geometry, without considering shear stresses, which are neglected in the theory. The formulation holds under the assumptions of , excluding , , or temperature-dependent effects that could alter the stress-strain response.

Conditions and Applications

Boundary conditions

In Euler–Bernoulli beam theory, boundary conditions define the kinematic and static constraints at the beam's ends, enabling the determination of unique solutions to the governing differential equation for deflection. These conditions reflect physical support types and directly influence the beam's structural response under loading. The fixed or clamped boundary condition constrains both transverse deflection and rotation at the end, expressed mathematically as w=0 and \frac{dw}{dx}=0, where w(x) is the transverse deflection. This setup models rigid encastrement, providing reaction forces and moments to enforce zero displacement and slope. The pinned or simply supported boundary condition prevents transverse deflection but allows rotation, given by w=0 and M=0, where the bending moment M = -EI \frac{d^2 w}{dx^2} = 0 (with E as and I as the ), simplifying to \frac{d^2 w}{dx^2}=0. Such supports transmit vertical reaction forces but no moments, common in bridge girders or floor beams. The free boundary condition imposes no kinematic constraints, requiring zero bending moment and shear force at the end: M=0 or \frac{d^2 w}{dx^2}=0, and V=0 or \frac{d^3 w}{dx^3}=0, where shear V = -EI \frac{d^3 w}{dx^3}. This represents an unsupported end, as in cantilever tips, with no reactions present. The guided or roller boundary condition restricts rotation but permits transverse deflection, specified as \frac{dw}{dx}=0 and V=0 or \frac{d^3 w}{dx^3}=0. It models scenarios like a vertically sliding support, transmitting moment reactions but no shear forces. These conditions dictate the forces and moments at supports—for instance, fixed ends sustain both, while free ends sustain neither—and govern the deflection profile's shape, such as linear slopes near pinned ends versus curved profiles near clamped ends. They are applied to the general solution of the beam equation to yield specific deflection curves.

Loading configurations

In Euler–Bernoulli beam theory, loading configurations primarily involve transverse loads that induce , with axial and loads playing supplementary roles in the overall response. The transverse load, denoted as q(x), represents the distributed per acting to the beam's and is central to the governing for static deflection. Common forms include uniform loading, where q(x) is constant along the beam , such as self-weight or evenly spread ; triangular loading, where q(x) varies linearly from zero to a maximum, often modeling accumulation or gradients; and point loading, characterized by a concentrated P applied at a specific x = a, which can be idealized as an impulsive contribution to q(x). Axial loads, represented by N(x), act along the beam's longitudinal axis and can be tensile (positive) or compressive (negative), influencing rather than primary in the linear Euler–Bernoulli framework. Compressive axial loads reduce the beam's resistance to transverse deflection through second-order effects, potentially leading to when the load exceeds a , though the theory's core assumptions prioritize small deflections from . Moment loads, denoted as m(x), describe applied couples or distributed torques that directly introduce rotational effects without net force, such as those from eccentric connections or thermal gradients. These are typically concentrated at points or vary along the length, altering the internal bending moment distribution in the beam. Due to the linearity of Euler–Bernoulli beam theory, the allows the total response—such as deflection or —from multiple simultaneous loads to be obtained by summing the individual responses from each load acting alone, facilitating analysis of complex configurations.

Cantilever beam examples

Cantilever beams, fixed at one end and free at the other, serve as fundamental examples for applying Euler–Bernoulli beam theory to static loading scenarios, illustrating how distributed and concentrated loads produce deflections and internal stresses. These analyses assume small deflections, linear material behavior, and neglect deformation, focusing on effects alone. Consider a cantilever of length L, flexural EI, subjected to a concentrated point load P applied transversely at the free end. The deflection w(x) along the beam, measured from the fixed end at x=0, is derived by integrating the governing EI \frac{d^4 w}{dx^4} = 0 with appropriate boundary conditions (w(0) = 0, \frac{dw}{dx}(0) = 0, \frac{d^2 w}{dx^2}(L) = 0, EI \frac{d^3 w}{dx^3}(L) = -P) and load term. The resulting expression is: w(x) = \frac{P x^2}{6 EI} (3L - x) The maximum deflection occurs at the free end (x = L) and equals \frac{P L^3}{3 EI}. The corresponding shear force diagram is linear, starting at -P at the fixed end and zero at the free end, while the bending moment diagram is linear, increasing from 0 at the free end to -P L at the fixed end. The deflection curve is cubic, reflecting the integration of the moment distribution. For a uniformly distributed load q (force per unit length) applied along the entire length, the governing equation becomes EI \frac{d^4 w}{dx^4} = q, with the same boundary conditions as above but adjusted at the free end (EI \frac{d^3 w}{dx^3}(L) = 0). The deflection is: w(x) = \frac{q x^2}{24 EI} (6 L^2 - 4 L x + x^2) The maximum deflection at the free end is \frac{q L^4}{8 EI}. Here, the diagram is parabolic, varying from -q L at the fixed end to 0 at the free end, the moment diagram is cubic, reaching a maximum of -\frac{q L^2}{2} at the fixed end, and the deflection follows a quartic profile. Bending stresses in these cantilever configurations arise from the normal strain due to curvature, given by the constitutive relation \sigma_x = -E z \frac{d^2 w}{dx^2}, which simplifies to the flexure formula \sigma = \frac{M y}{I} at any section, where M is the bending moment, y is the distance from the neutral axis, and I is the moment of inertia. For the point load case, the maximum stress occurs at the fixed end (M_\max = P L) on the outer fiber (y = c, the distance to the extreme fiber), yielding \sigma_\max = \frac{P L c}{I}. This stress is tensile on one side and compressive on the other, highlighting the theory's prediction of linear stress variation across the cross-section.

Three-point bending setup

The three-point bending setup consists of a simply supported at its two ends over a length L, with a concentrated vertical load P applied at the . The reactions at each support are equal and opposite to half the applied load, P/2, due to and static under Euler-Bernoulli assumptions of small deflections and plane sections remaining . This configuration induces a maximum at the center while minimizing effects away from the supports, making it ideal for isolating flexural behavior in material characterization tests. Under this loading, the beam exhibits a symmetric deflection profile governed by the Euler-Bernoulli equation EI \frac{d^4 w}{dx^4} = 0 in the unloaded segments, with boundary conditions of zero deflection at the supports (w(0) = w(L) = 0) and continuity of deflection and slope at the center. For $0 \leq x \leq L/2, the deflection w(x) is w(x) = \frac{P x}{48 EI} (3L^2 - 4x^2), where E is the Young's modulus and I is the area moment of inertia of the cross-section. The maximum deflection occurs at the load point x = L/2: w\left( \frac{L}{2} \right) = \frac{P L^3}{48 EI}. This central deflection is obtained by double integration of the moment-curvature relation M(x) = EI \frac{d^2 w}{dx^2}, using the piecewise moment distribution M(x) = (P/2) x for $0 \leq x \leq L/2. The corresponding maximum normal stress develops at the top and bottom fibers at midspan, where the bending moment peaks at M_{\max} = PL/4. For a rectangular cross-section of width b and depth h, the extreme fiber stress \sigma_{\max} is \sigma_{\max} = \frac{3 P L}{2 b h^2}, derived from the flexure formula \sigma = \frac{M y}{I} with maximum distance y = h/2 from the neutral axis and I = b h^3 / 12. This stress distribution assumes linear elastic material response and neglects shear contributions, consistent with slender beam approximations in Euler-Bernoulli theory. In applications, the three-point bending setup is extensively used to determine the elastic modulus by measuring the linear load-deflection slope and applying E = \frac{P L^3}{48 I w(L/2)}, providing a direct assessment of flexural stiffness for engineering materials. It also plays a key role in fracture mechanics, where the load at crack initiation or peak load in notched specimens quantifies fracture toughness and energy, particularly for brittle or composite materials under controlled crack propagation. These tests enable reliable evaluation of failure modes and mechanical reliability in structural design.

Advanced Topics

Large deflection extensions

When deflections in beams become large relative to the beam's thickness, the linear assumptions of classical Euler–Bernoulli theory, which neglect higher-order geometric effects, no longer suffice, necessitating nonlinear extensions to capture the coupling between and . These extensions account for moderate to large rotations while retaining the core kinematic hypothesis that plane sections remain plane and perpendicular to the . A key advancement is the von Kármán approximation for nonlinear , which incorporates a quadratic term arising from the transverse deflection to model the axial stretching induced by bending. The axial ε_x at a point along the length x and height y is given by \begin{equation} \varepsilon_x = \frac{du}{dx} + \frac{1}{2} \left( \frac{dw}{dx} \right)^2 - y \frac{d^2 w}{dx^2}, \end{equation} where u(x) is the axial displacement and w(x) is the transverse deflection. This formulation, originally developed for plates but adapted to beams, enables the prediction of behaviors such as geometric stiffening, where large deflections lead to increased effective stiffness due to induced membrane forces. For more precise modeling of extreme deflections, the exact κ replaces the d²w/dx². The is defined as κ = dθ/ds, where θ is the local rotation angle and s is the along the deformed beam, with ds = √[1 + (dw/dx)²] dx. For analytical tractability, this yields the approximation \begin{} \kappa \approx \frac{d^2 w / dx^2}{\left[1 + (dw/dx)^2 \right]^{3/2}}, \end{} which accounts for the nonlinear of the centerline without assuming small slopes. This expression is essential for problems involving significant rotations, as it directly relates the M to EIκ, where EI is the . The foundational framework for inextensional large bending—assuming no axial extension—is Euler's elastica theory, which treats the beam as a flexible rod under end loads, leading to a nonlinear solved via elliptic integrals. Developed by Leonhard Euler in 1744, this theory describes shapes as elastica curves, applicable to buckled or heavily loaded beams where deflections approach the beam length. Analytical solutions exist for specific boundary conditions, but general cases require numerical methods, such as shooting techniques or finite element discretizations, to integrate the governing equations. These extensions reveal important limitations: the geometric nonlinearity induces an apparent increase in , altering load-deflection responses and potentially delaying onset compared to linear predictions, while failure criteria must be adjusted to include membrane stresses that can cause premature yielding at the surfaces.

cases

Statically indeterminate cases occur in Euler–Bernoulli beam theory when the support conditions introduce redundant reactions, exceeding the number of available equations from alone. Examples include fixed-fixed beams, where both ends are clamped, and continuous beams spanning multiple supports, such as a two-span beam with three supports. In these configurations, the beam's deformation must satisfy both force and geometric conditions derived from the beam's deflection , \frac{d^2v}{dx^2} = \frac{M(x)}{EI}, where v is the transverse deflection, M(x) is the , E is the of elasticity, and I is the . Two primary solution approaches address these cases: the force method () and the displacement method. The force method releases the redundant supports to create a statically determinate primary , computes the displacements at the release points due to applied loads using Euler–Bernoulli techniques, and then imposes by applying redundant forces such that the net displacement at those points is zero. This leverages the principle of superposition inherent in the linear Euler–Bernoulli assumptions. The displacement method, exemplified by the slope-deflection equations, assumes unknown rotations and displacements at beam ends or joints, expresses end moments in terms of these via M = \frac{2EI}{L} (2\theta_A + \theta_B - 3\psi) (where \theta are end rotations, \psi is the chord rotation, and L is the length), and solves the resulting system using joint equilibrium. A representative example is the propped cantilever , a fixed-free with an additional vertical at the free end, rendering it to the first degree. For a uniform distributed load q over length L, the prop reaction R is determined by ensuring zero deflection at the prop location through : the downward deflection due to q alone, \delta_q = \frac{qL^4}{8EI}, is counteracted by the upward deflection due to R, \delta_R = \frac{RL^3}{3EI}, yielding R = \frac{3qL}{8}. This superposition is solved by integrating the Euler–Bernoulli moment-curvature relation twice for each loading component. For computational efficiency in the force method, the moment-area method or conjugate beam analogy can be integrated to evaluate compatibility displacements without full double integration. The moment-area method computes the tangential deviation as the first moment of the M/EI diagram about a point, while the conjugate beam treats the real 's M/EI as a load on a conjugate structure with boundary conditions mirroring the original, where and in the conjugate yield slope and deflection in the real . These graphical aids simplify analysis for complex indeterminate layouts while remaining grounded in Euler– .

References

  1. [1]
    The 100th Anniversary of the Timoshenko–Ehrenfest Beam Model
    Sep 30, 2022 · The first viable theory was suggested by Daniel Bernoulli (1700–1882) and Leonhard Euler (1707–1783), famous Swiss mathematicians. This is the ...
  2. [2]
    introduction to beam theory - Welcome to AE Resources
    This set of equations is a relationship between strain and displacement. Constitutive law is the out put of solving a two dimensional problem on cross section.Missing: fundamentals | Show results with:fundamentals
  3. [3]
    [PDF] the Euler-Bernoulli Beam model, the Rayleigh Beam model, and
    Dec 18, 2017 · The Euler-Bernoulli beam model normally produces an overestimate of the natural fre- quencies of a vibrating beam. In 1877, in order to improve ...
  4. [4]
    History of Engineering Mechanics - Harvard SEAS
    1750. The Euler-Bernoulli Beam Theory relates a beam's deflection to its applied load. Two major assumptions of Euler-Bernoulli Beam Theory are that plane ...
  5. [5]
    [PDF] Euler-Bernoulli Beams - Professor Terje Haukaas
    Aug 21, 2023 · The Euler-Bernoulli beam theory was established around 1750 with contributions from. Leonard Euler and Daniel Bernoulli.
  6. [6]
    [PDF] Euler-Bernoulli Beams: Bending, Buckling, and Vibration
    Feb 9, 2004 · Euler-Bernoulli Beam Theory: Displacement, strain, and stress distributions. Beam theory assumptions on spatial variation of displacement ...Missing: fundamentals | Show results with:fundamentals
  7. [7]
    Terms Explained - Engineering Beam Theory
    Apr 6, 2024 · Assumptions and limitations · The cross-section of the beam is considered small compared to its length meaning that the beam is long and thin.
  8. [8]
    Euler-Bernoulli Beam Theory - an overview | ScienceDirect Topics
    This first-order shear deformation theory relaxes the normality assumption of the Euler–Bernoulli beam theory but assumes a constant transverse shear strain ( ...<|control11|><|separator|>
  9. [9]
    [PDF] Nonlinear bending models for beams and plates - LSU Math
    Aug 20, 2014 · It is known that the solution for large deflections of a beam cannot be obtained from the elementary linear EB theory. This theory neglects the ...Missing: limitations | Show results with:limitations
  10. [10]
    Beam theory: Bending - JPE
    Euler-Bernoulli beam theory is only valid with the following assumptions: Generally, these criteria are met when the beam is a slender beam with small ...
  11. [11]
    [PDF] 2.5. THEORIES OF STRAIGHT BEAMS | TAMU Mechanics
    Mar 2, 2017 · In the Euler–Bernoulli beam theory (EBT), where one assumes that L >> 2h and ν = 0, we have u1 = 0, and u3 is given by. uEBT. 3. (x1, x3) ...
  12. [12]
    Euler Bernoulli Beam - Engineering at Alberta Courses
    Compute the bending moment, the shear force, the stress distribution, and the strain distribution in a beam by solving the differential equation of equilibrium.
  13. [13]
    Euler-Bernoulli Beam Equation Derivation and Assumptions
    The Euler-Bernoulli beam equation is derived from four segments of beam theory: kinematics, constitutive, resultants, and equilibrium. The Euler-Bernoulli ...
  14. [14]
    [PDF] Vibrational Control of a Cantilever Beam Using Sensor and Actuator ...
    May 10, 2024 · This paper covers the vibrational control of cantilever beams. I will begin by deriving the governing equations of a fixed-free cantilever beam ...<|control11|><|separator|>
  15. [15]
    [PDF] Impacts of various boundary conditions on beam vibrations - CORE
    There are three assumptions for the Euler-Bernoulli beam theory ... two rotational springs under a moving delta load using Euler-Bernoulli beam theory and.Missing: limitations | Show results with:limitations
  16. [16]
    Vibration of Continuous Systems | Wiley Online Books
    Dec 13, 2006 · Vibration of Continuous Systems ; Author(s):. Singiresu S. Rao, ; First published:13 December 2006 ; Print ISBN:9780471771715 | ; Online ISBN: ...
  17. [17]
    [PDF] comparison between the dynamic response of selective laser
    Dec 6, 2016 · For a cantilever beam, the corresponding values are as follows: 𝛽1l = 1.875104, 𝛽2l = 4.694091, 𝛽3l = 7.854757. Figure 3.3 shows the ...
  18. [18]
    A Note on the Use of Approximate Solutions for the Bending ...
    Mar 17, 2010 · Let us consider a simply supported Euler–Bernoulli beam with length L and constant height H ⁠, cross section area A ⁠, and moment of inertia I ⁠ ...
  19. [19]
  20. [20]
    Natural Frequencies and Normal Modes of a Spinning Timoshenko ...
    Expressions for computing natural frequencies and mode shapes are given. Numerical simulation studies show that the simply-supported beam possesses very ...
  21. [21]
    Analytical solution for modal analysis of Euler-Bernoulli and ... - Extrica
    Sep 30, 2018 · Table 2 shows the frequencies of the E-B and Timoshenko beams with constant thicknesses and simply supported (SS) boundary conditions. The ...
  22. [22]
    Exact expressions for numerical evaluation of high order modes of ...
    Dec 1, 2018 · Technical note. Exact expressions for numerical evaluation of high order modes of vibration in uniform Euler-Bernoulli beams.
  23. [23]
    [PDF] Flexural Stresses In Beams (Derivation of Bending Stress Equation)
    Beam has a longitudinal plane of symmetry and the bending moment lies within this plane. 3. Beam is subjected to pure bending (bending moment does not change ...
  24. [24]
    [PDF] CHAP 3 FINITE ELEMENT ANALYSIS OF BEAMS AND FRAMES
    BEAM THEORY cont. • Euler-Bernoulli Beam Theory cont. – Plane sections normal to the beam axis remain plane and normal to the axis after deformation (no ...Missing: fundamentals | Show results with:fundamentals
  25. [25]
    [PDF] 10. Beams: Flexural and shear stresses x
    Euler- Bernoulli definitions and kinematic assumptions for thin beams. Consider the following assumptions related to the geometry and loading of a beam: • The ...
  26. [26]
    Bernoulli Beam - an overview | ScienceDirect Topics
    The Euler-Bernoulli beam is the simplest out-of-plane bending element. Although this element is often considered as part of matrix structural analysis.<|control11|><|separator|>
  27. [27]
    [PDF] Force/Deflection relationships - Rice University
    Here, the relationship between moment m(x) and curvature c(x) is the proxy for “stress”-“strain”. For a beam element the Bernoulli-Euler simple beam theory ...
  28. [28]
    Classical Beam Theories of Structural Mechanics - Google Books
    Jun 13, 2021 · This book provides a systematic and thorough overview of the classical bending members based on the theory for thin beams (shear-rigid) ...<|control11|><|separator|>
  29. [29]
    Euler–Bernoulli Beam Elements - ResearchGate
    Oct 19, 2025 · This chapter provides the fundamental description of one-dimensional thin beam members. Based on the three basic equations of continuum ...
  30. [30]
    [PDF] Appendix A: Exact Analytical Solutions of Straight Bars and Beams
    Table A3.1: Conventional boundary conditions for the Bernoulli–Euler beam theory. Type of support. Geometric B.C.. Force B.C..Missing: guided | Show results with:guided
  31. [31]
    Euler Beam - an overview | ScienceDirect Topics
    Euler–Bernoulli Beam Theory. Treated are external loads, cross-sections, and ... Reference work. Encyclopedia of Vibration. 2001, Encyclopedia of Vibration. R.A. ...Missing: original | Show results with:original
  32. [32]
    [PDF] Euler-Bernoulli Beam ∫ ∫ ∫ ∫ - Rose-Hulman
    The boundary conditions for solving are slope (Neumann) on the boundary (courtesy integration by parts) and deflection (Dirichlet) on the boundary.
  33. [33]
    [PDF] Euler-Bernoulli and Timoshenko beam theory - TAMU Mechanics
    Euler-Bernoulli beam theory. ➢ Kinematics. ➢ Equilibrium equations. ➢ Governing equations in terms of the displacements. Timoshenko beam theory. ➢ Kinematics.
  34. [34]
  35. [35]
    [PDF] Beams with Axial Force - Professor Terje Haukaas
    Mar 4, 2019 · In this document the Euler-Bernoulli beam theory is extended with P-delta effects, i.e., the influence of axial force on the bending ...
  36. [36]
    [PDF] 18. Column buckling
    Comments on boundary conditions and buckling analysis. Recall that a pinned-end condition corresponds to boundary conditions of zero bending moment at the pin.Missing: roller | Show results with:roller
  37. [37]
    [PDF] 7.4 The Elementary Beam Theory
    The term beam has a very specific meaning in engineering mechanics: it is a component that is designed to support transverse loads, that is, loads that act ...
  38. [38]
    [PDF] Beams Deflections (Method of Superposition)
    The method of superposition states that beam deflections from multiple loads are the sum of individual loads, and is useful for statically indeterminate beams.
  39. [39]
    [PDF] Discussion for indeterminate beams
    The enforcement of displacement and rotation boundary conditions produce the equations needed to solve for the external reaction loads. Once the external ...Missing: clamped | Show results with:clamped
  40. [40]
    Vibrations of Cantilever Beams:
    One method for finding the modulus of elasticity of a thin film is from frequency analysis of a cantilever beam.
  41. [41]
    New Page 1
    Jan 6, 2002 · Problem: Determine the end deflection and root bending stress of a steel cantilever beam modeled as a 2D problem. ... Mc/I, c = h/2 ! A ...
  42. [42]
    [PDF] Lecture 5: Solution Method for Beam Deflections
    The concentrated load P can be treated as a special case of the distributed load q(x) = Pδ(x - x0), where δ is the Dirac delta function.
  43. [43]
    [PDF] Bending Test – Formulas
    Elastic Modulus. Formula. Ef. Flexural Stress (3-point). Of. = Of. 3FL. 2bd2. = FL. Flexural Stress (4-point). = L3m. 4bd3. Rectangular. Circular. πR3. Flexural ...Missing: maximum | Show results with:maximum
  44. [44]
    [PDF] ANALYSIS OF THE THREE-POINT-BEND TEST FOR MATERIALS ...
    An analysis capability is described for the three-point-bend test applicable to materials of linear but unequal tensile and compressive stress-strain ...
  45. [45]
    [PDF] Fracture Energy for Three-Point Bend Tests on Single-Edge ... - DTIC
    ABSTRACT Three test series of single-edge notched beams in three-point bending were conducted to evaluate the fracture energy of concrete.
  46. [46]
    Construction of beam elements considering von Kármán nonlinear ...
    The von Kármán nonlinear strains based formulation introduced in the paper is applicable for such class of beam problems undergoing large transverse deflection ...
  47. [47]
    [PDF] On the Nonlinear Deformation Geometry of Euler-Bernoulli Beams
    Nonlinear expressions are developed to relate the orientation of the deformed-beamcross section, torsion, local componentsof bending curvature,.
  48. [48]
    [PDF] A STUDY OF LARGE DEFLECTION OF BEAMS AND PLATES
    Examples include the effects of rotatory inertia, shear distortion, body couples, and the effect of axial deformation and axial forces. Including all of these ...<|control11|><|separator|>
  49. [49]
    Large-Deflection Nonlinear Mechanics of Curved Cantilevers Under ...
    Aug 19, 2021 · Solution begins with the Euler–Bernoulli beam theory, where the local curvature κ(s) = dθ/ds at a given point s along a beam, with initial ...
  50. [50]
    [PDF] Euler's Original Derivation of Elastica Equation - Scholarly Commons
    Sep 4, 2025 · Euler derived the differential equations of elastica by the variational method in 1744, but his original derivation has never been properly ...
  51. [51]
    A variational approach for large deflection of ends supported ...
    The exact curvature corresponding to the elastica theory is used to obtain the large deflection configurations of a nanorod, while in the linear beam theory ...
  52. [52]
    [PDF] ELEMENTS OF ARCHITECTURAL STRUCTURES
    There are two general methods for analysis of statically indeterminate structures; the force or flexibility method, and the stiffness or displacement method.
  53. [53]
    [PDF] Deflection and Supporting Force Analysis of a Slender Beam ... - DTIC
    propped cantilever beam has been derived using Euler-Bernoulli beam theory for a ... ' Deflection shape of the statically-indeterminate propped cantilever beam.