Fact-checked by Grok 2 weeks ago

Stress functions

Stress functions are scalar, vector, or tensor potentials employed in the theory of to parametrize the components of the tensor in such a way that the equations of static are satisfied identically, thereby reducing the complexity of solving boundary value problems in . These functions transform the equilibrium conditions—expressed as the of the tensor being zero in the absence of body forces—into constraints on the potentials themselves, often leading to higher-order partial differential equations like the for two-dimensional cases. In two-dimensional plane stress or plane strain problems, the most prominent example is the Airy stress function φ(x₁, x₂), introduced by in 1862, where the components are defined as σ₁₁ = ∂²φ/∂x₂², σ₂₂ = ∂²φ/∂x₁², and σ₁₂ = -∂²φ/∂x₁∂x₂, ensuring without body forces and requiring φ to satisfy ∇⁴φ = 0 for in isotropic materials. This approach is particularly effective for problems with prescribed boundary tractions, such as beams under loading or cracks in . For three-dimensional elasticity, several formulations exist to achieve the minimal number of three independent functions needed to satisfy the three equations while addressing the six compatibility conditions derived from strain relations via . The Maxwell stress functions, developed by James Clerk Maxwell in 1870, represent the tensor through a whose yields the stresses, originally motivated by reciprocal diagrams in . In contrast, the Beltrami stress functions (1892) use symmetric tensor potentials with six components, while the Morera stress functions (also 1892) employ three functions, where the is the curl of the potential tensor, providing greater flexibility but requiring solutions to Beltrami-Michell equations for compatibility in the absence of body forces. These methods, such as the Boussinesq-Neuber representation for isotropic bodies, enable the expression of both stresses and displacements in terms of potentials, facilitating analytical solutions for geometries. Beyond classical elasticity, stress function concepts have influenced higher-dimensional generalizations, including Einstein's 1915 tensor potentials for four-dimensional gravitational fields and Lanczos potentials (1949–1962) linking elasticity to variational principles in , though their primary applications remain in problems like and material .

Introduction to Stress Functions

Definition and Purpose

In , the deformation of solid bodies under external loads is analyzed assuming small displacements and strains, where the tensor is linearly related to the strain tensor via . This framework governs the behavior of isotropic or anisotropic materials, focusing on reversible deformations without . functions serve as scalar or tensor potentials that express the components of the six-element tensor in terms of their derivatives, thereby ensuring that the equilibrium equations are satisfied identically throughout the body. In this approach, the stresses are derived from the potentials such that the of the tensor vanishes in the absence of body forces, simplifying the governing system. The primary purpose of is to reduce the number of unknowns in value problems of , from the full set of six independent components to a smaller number of potential functions—typically one in two dimensions or three functions in three dimensions—while automatically incorporating . This reduction is particularly effective for problems involving surface tractions or body forces that can be derived from a , as the potentials then satisfy higher-order partial differential equations, such as biharmonic equations in two dimensions or coupled systems in three dimensions, derived from conditions. By focusing the solution process on these fewer functions and their conditions, facilitate analytical or numerical solutions to complex problems.

Historical Background

The development of stress functions in elasticity theory began in the mid-19th century as a means to simplify the solution of equilibrium problems in elastic solids, initially assuming no body forces with later extensions possible. In 1863, introduced the Airy stress function specifically for two-dimensional problems under conditions, representing the stress components in terms of second derivatives of a single to automatically satisfy the equilibrium equations in the absence of body forces. Building on this foundation, James Clerk Maxwell contributed to three-dimensional formulations in 1870 by proposing stress functions in the form of a diagonal for isotropic elastic media, which allowed for a representation of the stress tensor through potentials that ensured equilibrium while accommodating body forces. Eugenio Beltrami extended these ideas to general three-dimensional cases in 1892, generalizing the approach with a stress tensor potential that incorporated off-diagonal elements, providing a more versatile framework for arbitrary stress states in bodies. In 1892, Giuseppe Morera further refined the three-dimensional representation by introducing an off-diagonal tensor form for the stress functions, emphasizing antisymmetric components to address specific compatibility requirements in elasticity. A notable specialization occurred in 1903 when adapted function methods for torsion problems in prismatic bars, developing a that satisfied both and under twisting loads, famously linked to his membrane analogy for intuitive geometric interpretation. These 19th- and early 20th-century innovations, rooted in , laid the groundwork for function methods, which evolved through the 20th century into integral components of for simulating complex elastic deformations in applications.

Fundamental Concepts in Elasticity

Equilibrium Equations

In the theory of , the equations describe the conditions under which a deformable solid remains in static balance, assuming small deformations and neglecting inertial forces due to the absence of time-dependent accelerations. These equations arise from the requirement that the on any within the must be zero, balancing surface tractions from stresses with internal body forces. The fundamental form of the static equilibrium equations in tensor notation is given by \frac{\partial \sigma_{ij}}{\partial x_j} + f_i = 0, where \sigma_{ij} are the components of the symmetric , x_j denote the spatial coordinates, repeated indices imply (Einstein ), and f_i represent the components per unit volume, such as gravitational or electromagnetic forces. This formulation ensures force balance in three dimensions for a under linear assumptions. In the common case of negligible body forces (f_i = 0), the equations simplify to \frac{\partial \sigma_{ij}}{\partial x_j} = 0, meaning the divergence of the stress tensor vanishes, which imposes that the stress field is solenoidal or divergence-free. In Cartesian coordinates (x, y, z), the equilibrium equations expand into their component-wise form. For the x-direction, it reads \frac{\partial \sigma_{xx}}{\partial x} + \frac{\partial \sigma_{xy}}{\partial y} + \frac{\partial \sigma_{xz}}{\partial z} + f_x = 0, with analogous expressions for the y- and z-directions obtained by cyclic permutation of the indices. Without body forces, these become \frac{\partial \sigma_{xx}}{\partial x} + \frac{\partial \sigma_{xy}}{\partial y} + \frac{\partial \sigma_{xz}}{\partial z} = 0, and similarly for the other components. These partial differential equations must hold throughout the domain of the elastic body to maintain mechanical equilibrium. Stress functions provide a powerful approach to solving these equations by parameterizing the stress components in a way that automatically enforces . Typically, the stresses \sigma_{ij} are expressed as curls or higher-order derivatives of scalar or vector potentials, ensuring that the divergence \frac{\partial \sigma_{ij}}{\partial x_j} identically equals zero (or balances f_i if included). This construction transforms the equilibrium conditions into identities, allowing focus on boundary conditions and compatibility requirements, such as those in the Beltrami-Michell equations for deformation consistency.

Beltrami-Michell Compatibility Equations

The Beltrami-Michell equations represent the necessary and sufficient conditions for a in a linearly , isotropic to correspond to a compatible , ensuring that the strains can be integrated to yield single-valued, continuous displacements throughout the domain. These equations arise from transforming the geometric requirements of the tensor into conditions on the tensor, incorporating the material's constitutive response via . The foundational strain compatibility conditions, known as Saint-Venant's equations, are expressed in tensor notation as \varepsilon_{ij,kl} + \varepsilon_{kl,ij} - \varepsilon_{ik,jl} - \varepsilon_{jl,ik} = 0, where \varepsilon_{ij} denotes the components of the infinitesimal strain tensor, and commas indicate partial differentiation with respect to Cartesian coordinates x_k, x_l. These six independent equations (in three dimensions) guarantee that the strain field derives from a unique via \varepsilon_{ij} = \frac{1}{2}(u_{i,j} + u_{j,i}), preventing discontinuities or multi-valued displacements that would violate the continuum assumption. To obtain the stress-based form, substitute for an isotropic material, which relates strains to stresses as \varepsilon_{ij} = \frac{1+\nu}{E} \sigma_{ij} - \frac{\nu}{E} \delta_{ij} \sigma_{kk}, where E is , \nu is , \delta_{ij} is the , and \sigma_{kk} is the of the stress tensor. Differentiating this relation and inserting into the Saint-Venant equations, while accounting for the equilibrium equations \sigma_{ji,j} + f_i = 0 (with f_i the components per unit volume), yields the Beltrami-Michell equations: \sigma_{ij,kk} + \frac{1}{1+\nu} \sigma_{kk,ij} = -\frac{1+\nu}{E} (f_{i,j} + f_{j,i}). This derivation involves applying the Laplacian operator (via ,kk) to the stress components and cross-differentiating to eliminate displacement terms, resulting in a system of six equations that must hold alongside the three equilibrium equations for a complete stress formulation. In the absence of body forces (f_i = 0), the equations simplify to the homogeneous form originally proposed by Beltrami in 1892: \sigma_{ij,kk} + \frac{1}{1+\nu} \sigma_{kk,ij} = 0. Michell extended this in 1900 to include body forces, completing the compatibility framework for general loading. These equations play a critical role in verifying that an equilibrated stress field—satisfying force balance—is geometrically admissible, as integrated strains must produce consistent displacements without gaps or overlaps in the deformed configuration. Stress functions, such as those introduced later, are designed to automatically satisfy these compatibility conditions by reducing them to partial differential equations on scalar or tensor potentials.

General Three-Dimensional Stress Functions

Beltrami Stress Functions

The Beltrami stress functions provide a general representation of the stress tensor in three-dimensional using a symmetric second-order tensor potential \Phi_{kl}, where k, l = 1, 2, 3. The stress components are given by \sigma_{ij} = \epsilon_{ikm} \epsilon_{jln} \frac{\partial^2 \Phi_{kl}}{\partial x_m \partial x_n}, with \epsilon denoting the . This formulation originates from Eugenio Beltrami's work on solving equilibrium problems in elastic solids without body forces. This double-curl structure ensures that the equilibrium equations \partial_j \sigma_{ij} = 0 are satisfied identically for each component i, as the divergence of a double curl vanishes. The potential \Phi_{kl} must be symmetric (\Phi_{kl} = \Phi_{lk}) and sufficiently smooth to represent arbitrary symmetric stress fields meeting . In the isotropic case, this approach focuses on materials governed by with Poisson's ratio \nu. To ensure compatibility of the strains derived from the stresses via the constitutive relations, the potential \Phi_{kl} must be four-times continuously differentiable and satisfy higher-order partial differential equations obtained by substituting the stress expression into the Beltrami-Michell compatibility equations. These equations take the form \nabla^2 \sigma_{ij} + \frac{1}{1+\nu} \frac{\partial^2 \sigma_{kk}}{\partial x_i \partial x_j} = 0 (in the absence of body forces), leading to biharmonic-like conditions on \Phi_{kl} that incorporate \nu. The Beltrami approach offers advantages in handling general three-dimensional problems in isotropic elasticity by reducing the problem to finding a suitable \Phi_{kl} that matches boundary conditions, and it extends naturally to anisotropic materials through modified constitutive tensors. For explicit computation, the stress components are \sigma_{xx} = \frac{\partial^2 \Phi_{yy}}{\partial z^2} + \frac{\partial^2 \Phi_{zz}}{\partial y^2} - 2\frac{\partial^2 \Phi_{yz}}{\partial y \partial z}, with cyclic permutations for \sigma_{yy} and \sigma_{zz}, and shear components involving mixed derivatives of the off-diagonal potentials with appropriate factors of 2.

Maxwell Stress Functions

The Maxwell stress functions provide a simplification of the general Beltrami stress functions for solving three-dimensional problems in isotropic , by assuming the Beltrami stress tensor is diagonal with three scalar potentials A(x,y,z), B(x,y,z), and C(x,y,z) along the diagonal components \Phi_{xx} = A, \Phi_{yy} = B, and \Phi_{zz} = C, while the off-diagonal components are zero. This assumption aligns the coordinate axes orthogonally with the principal directions of the function tensor, enabling the representation of stresses without body forces. The stress components are derived from these potentials as follows: \sigma_{xx} = \frac{\partial^2 B}{\partial z^2} + \frac{\partial^2 C}{\partial y^2} \sigma_{yy} = \frac{\partial^2 A}{\partial z^2} + \frac{\partial^2 C}{\partial x^2} \sigma_{zz} = \frac{\partial^2 A}{\partial y^2} + \frac{\partial^2 B}{\partial x^2} \sigma_{xy} = -\frac{\partial^2 A}{\partial x \partial y} \sigma_{xz} = -\frac{\partial^2 B}{\partial x \partial z} \sigma_{yz} = -\frac{\partial^2 C}{\partial y \partial z} These expressions identically satisfy the equations in the absence of body forces, reducing the problem to determining the potentials that meet conditions and . The form originates from reciprocal diagrams approach, extended to elasticity, where each potential acts analogously to an in a mutually orthogonal . To ensure compatibility with the strain field in isotropic materials, the potentials must satisfy the Beltrami-Michell equations, which for this diagonal case yield a system of coupled partial differential equations involving the biharmonic operator and \nu. For instance, in the absence of body forces, they can be expressed as a set of three equations (cyclic permutations): \nabla^4 A - \nu^2 (\nabla^4 B + \nabla^4 C) + \left( \frac{\partial^2 A}{\partial x^2} + \frac{\partial^2 B}{\partial y^2} + \frac{\partial^2 C}{\partial z^2} \right) = 0, with similar forms for B and C, reflecting the material's lateral contraction behavior. This formulation is applied to general three-dimensional elasticity problems without body forces, such as stress analysis in rectangular prisms or components with prescribed surface tractions orthogonal to the axes. In two dimensions, it reduces to the Airy stress function; for example, setting B = C = 0 yields a plane problem in the xy-plane with A serving as the Airy potential, where \sigma_{xx} = \partial^2 A / \partial y^2, \sigma_{yy} = \partial^2 A / \partial x^2, and \sigma_{xy} = -\partial^2 A / \partial x \partial y. The approach is limited to cases where the orthogonality assumption holds, restricting its use to problems where the principal stress directions align with the chosen Cartesian coordinates; otherwise, the full six-component Beltrami tensor is required.

Specialized Three-Dimensional Stress Functions

Morera Stress Functions

Morera stress functions provide a representation of the stress tensor in three-dimensional linear elasticity using an off-diagonal potential tensor \Phi_{ij}, where the diagonal components are zero. This approach assumes \Phi_{xy} = A(y,z), \Phi_{zx} = B(z,x), and \Phi_{yx} = C(x,y), with the tensor being antisymmetric such that \Phi_{ji} = -\Phi_{ij}. The formulation automatically satisfies the equilibrium equations in the absence of body forces. The components of the stress tensor are expressed in terms of second partial derivatives of these potentials. Specifically, \sigma_{xx} = -2 \frac{\partial^2 A}{\partial y \partial z}, \quad \sigma_{yy} = -2 \frac{\partial^2 B}{\partial z \partial x}, \quad \sigma_{zz} = -2 \frac{\partial^2 C}{\partial x \partial y}, and for the shear stresses, \sigma_{yz} = \frac{\partial^2 A}{\partial y^2} + \frac{\partial^2 A}{\partial z^2} + \frac{\partial^2 B}{\partial x^2}, \quad \sigma_{zx} = \frac{\partial^2 B}{\partial z^2} + \frac{\partial^2 B}{\partial x^2} + \frac{\partial^2 C}{\partial y^2}, \quad \sigma_{xy} = \frac{\partial^2 C}{\partial x^2} + \frac{\partial^2 C}{\partial y^2} + \frac{\partial^2 A}{\partial z^2}. These expressions derive from the general form \sigma_{ij} = \epsilon_{ikl} \epsilon_{jmn} \frac{\partial^2 \Phi_{lm}}{\partial x_k \partial x_n}, adapted for the off-diagonal choice. To ensure compatibility with the Beltrami-Michell equations, substitution of these stress components yields a system of coupled biharmonic-like equations for A, B, and C, incorporating \nu. For instance, one such equation is \nabla^4 A + (1-2\nu) \left( \frac{\partial^4 A}{\partial y^4} + \frac{\partial^4 A}{\partial z^4} + \cdots \right) + cross terms involving B and C = 0, with similar forms for the others. These equations couple the potentials, reflecting the interdependence in three dimensions. This representation is particularly advantageous for problems involving prescribed stresses on boundaries or when working in non-Cartesian coordinate systems, as the off-diagonal structure naturally accommodates shear-dominated fields. Additionally, it forms the basis for the Prandtl stress function in torsional problems by specializing to two dimensions with appropriate assumptions on the potentials.

Relations Between Three-Dimensional Functions

The Beltrami stress functions represent the most general approach among the three-dimensional formulations, utilizing a symmetric second-order tensor with six independent components to express the tensor in isotropic . In contrast, the Maxwell stress functions constitute a restriction of the Beltrami representation by setting the off-diagonal components of the tensor to zero, thereby employing only three scalar functions primarily associated with normal stress contributions. Similarly, the Morera stress functions restrict the Beltrami tensor by nullifying its diagonal components, relying on the three off-diagonal elements to capture effects. This hierarchical structure ensures that the Beltrami formulation encompasses both Maxwell and Morera as special cases, providing a unified for solving the and compatibility equations in three dimensions. Equivalence between these representations allows for straightforward embedding: any solution obtained via Maxwell functions can be incorporated into the Beltrami framework simply by assigning zero values to the off-diagonal tensor components, while a Morera solution embeds by setting the diagonal components to zero. Transformations between them are achieved by selectively activating or deactivating specific tensor components, potentially augmented by adding biharmonic particular solutions to ensure completeness for non-self-equilibrated fields or multiply connected regions. Such conversions preserve the satisfaction of the Beltrami-Michell compatibility equations, though care must be taken to maintain equilibrium. For instance, transitioning from a Maxwell to a full Beltrami representation involves introducing non-zero off-diagonal terms that adjust shear components without altering the original normal stresses. Selection criteria for these functions depend on the problem's characteristics: representations are preferred for axisymmetric problems or scenarios dominated by normal stresses, such as pressurized vessels, due to their simplicity in handling diagonal-dominated tensors. Morera functions suit applications emphasizing shear or torsional effects, like twisting shafts, where off-diagonal terms dominate. The full Beltrami tensor is essential for general cases, including anisotropic materials or complex loading where both normal and shear interactions are significant, offering greater flexibility despite increased computational demands. All three incorporate ν through the Beltrami-Michell conditions, which couple the Laplacian of the tensor to the trace via factors like (1+ν)/(1-ν), but the Beltrami form accommodates broader material laws by allowing independent variation across tensor components.

Two-Dimensional Stress Functions

Airy Stress Function

The Airy stress function is a scalar potential employed in two-dimensional linear elasticity to solve plane stress and plane strain problems, where deformations and stresses exhibit no variation in the z-direction and shear stresses σ_{xz} and σ_{yz} vanish. In plane stress, the out-of-plane normal stress σ_{zz} is zero, while in plane strain, σ_{zz} = ν(σ_{xx} + σ_{yy}), with ν denoting Poisson's ratio. This approach, introduced by George Biddell Airy in his 1863 analysis of beam strains, reduces the problem to finding a single function that satisfies both equilibrium and compatibility conditions. The stresses are derived from the ϕ(x, y), which emerges as a special case of the three-dimensional Maxwell stress functions by setting the vector components to zero except for the scalar part C = ϕ(x, y). The in-plane stress components are then given by \begin{align*} \sigma_{xx} &= \frac{\partial^2 \phi}{\partial y^2}, \\ \sigma_{yy} &= \frac{\partial^2 \phi}{\partial x^2}, \\ \sigma_{xy} &= -\frac{\partial^2 \phi}{\partial x \partial y}. \end{align*} These relations automatically fulfill the two-dimensional equations ∂σ_{xx}/∂x + ∂σ_{xy}/∂y = 0 and ∂σ_{xy}/∂x + ∂σ_{yy}/∂y = 0 in the absence of body forces. Strain compatibility requires the to satisfy the \nabla^4 \phi = \frac{\partial^4 \phi}{\partial x^4} + 2\frac{\partial^4 \phi}{\partial x^2 \partial y^2} + \frac{\partial^4 \phi}{\partial y^4} = 0 for problems without body forces; for conservative body forces derived from a potential V (f_x = -∂V/∂x, f_y = -∂V/∂y), the equation becomes ∇⁴ φ = ∇² V. In polar coordinates (r, θ), the biharmonic operator becomes more complex, facilitating solutions for axisymmetric geometries. Boundary tractions are enforced by specifying ϕ and its normal ∂ϕ/∂n along the : the tangential is proportional to ∂ϕ/∂n, while the normal relates to ∂ϕ/∂s, where s is the , allowing direct matching to prescribed loads. For instance, a stress-free requires both ϕ = constant and ∂ϕ/∂n = 0. Representative solutions include cubic polynomials for beams under transverse end loads, such as ϕ = (P/6EI) (x^3 y - 3 x y^3 + 2 y^3 l), which yield parabolic and linear stresses, and or logarithmic terms for infinite plates with circular under remote uniaxial , revealing stress concentrations up to three times the applied at the equator.

Prandtl Stress Function

The Prandtl stress function, introduced by in 1903, provides a mathematical framework for determining the distribution in prismatic bars under pure torsion, particularly for non-circular cross-sections where the simple circular shaft solution does not apply. This function simplifies the Saint-Venant torsion problem by automatically satisfying the equilibrium equations in the cross-section while incorporating compatibility through a governing . It is especially valuable for applications involving complex geometries, such as elliptical or rectangular sections, enabling computation of torsional rigidity and maximum shear stresses without directly solving for the warping function. In the formulation, the Prandtl stress function \phi(x, y) relates to the non-zero components as follows: \tau_{xz} = \frac{\partial \phi}{\partial y}, \quad \tau_{yz} = -\frac{\partial \phi}{\partial x}. These expressions ensure that the condition \frac{\partial \tau_{xz}}{\partial x} + \frac{\partial \tau_{yz}}{\partial y} = 0 is satisfied identically within the cross-section. Substituting into the compatibility requirements for isotropic linear elastic materials yields : \nabla^2 \phi = \frac{\partial^2 \phi}{\partial x^2} + \frac{\partial^2 \phi}{\partial y^2} = -2 G \theta, where G is the shear modulus and \theta is the angle of twist per unit length. For the free lateral surface of the bar, the boundary condition requires that the shear stress vector be tangent to the boundary, leading to \phi = C (a constant, often set to zero for simply connected solid sections without loss of generality) along the cross-section contour. The torsional moment (torque) M_t is obtained by integrating over the cross-sectional area A: M_t = 2 \iint_A \phi \, dx \, dy. This relation connects the applied torque directly to the stress function, allowing computation of the torsional constant (or polar moment of inertia analog) as J = M_t / (G \theta) = -2 \iint_A \phi \, dx \, dy / \theta. The magnitude of the shear stress at any point is \tau = |\nabla \phi|, with the maximum typically occurring at the boundary. Prandtl's key insight was the membrane analogy, which visualizes the solution to the Poisson equation as the deflection w(x, y) of a thin elastic under uniform transverse pressure p = 2 G \theta / T, where T is membrane tension. In this analogy, \phi corresponds to the volume under the deflected membrane surface (up to scaling), the contours of \phi match the membrane slope lines representing directions, and the is twice the enclosed volume. This physical interpretation facilitates qualitative understanding and approximate solutions for irregular sections using soap films or numerical analogs, though exact solutions require solving the . Exact closed-form solutions exist for certain geometries. For an elliptical cross-section with semi-axes a (along x) and b (along y), the stress is \phi = -\frac{G \theta a^2 b^2}{a^2 + b^2} \left( \frac{x^2}{a^2} + \frac{y^2}{b^2} - 1 \right), yielding M_t = \frac{\pi a^3 b^3 G \theta}{a^2 + b^2} and maximum \tau_{\max} = \frac{2 M_t}{\pi a b^2} at the ends of the minor axis. For rectangular sections, solutions involve infinite series, but the membrane analogy provides effective approximations for thin or narrow bars. Extensions to multiply connected domains (e.g., hollow sections) require adjusting the constant on inner boundaries to enforce zero . This remains a cornerstone for analytical and numerical torsion analyses in .

References

  1. [1]
    [PDF] The stress functions of the three-dimensional continuum and elastic ...
    The stress function tensor, whose divergence is prescribed by the equilibrium conditions and whose rotation is prescribed by the compatibility conditions, ...
  2. [2]
    Elastic solutions - 5.2 Airy functions - Applied Mechanics of Solids
    The Airy Stress Function approach works best for problems where a solid is subjected to prescribed tractions on its boundary, rather than prescribed ...
  3. [3]
    [PDF] AIRY, BELTRAMI, MAXWELL, MORERA, EINSTEIN AND ... - HAL
    ABSTRACT. The main purpose of this paper is to revisit the well known potentials, called stress functions, needed in order to study the parametrizations of ...
  4. [4]
    [PDF] A treatise on the mathematical theory of elasticity - HAL
    Apr 28, 2016 · ... treatise on the mathematical theory of elasticity. Augustus Love. To ... stress, stress-strain relations, the strength of materials, and ...
  5. [5]
    [PDF] 3.2 The Stress Function Method
    Feb 3, 2010 · By using the Airy stress function representation, the problem of determining the stresses in an elastic body is reduced to that of finding a ...
  6. [6]
    [PDF] DETERMINATION OF STRESSES IN ELASTIC SOLIDS USING ...
    The classical equations of elasticity i n terms of the stress components are reduced to a set of three equations in terms of the three Maxwell stress.
  7. [7]
    Development of an Airy stress function of general applicability in one ...
    Jan 1, 1990 · Stress function. REFERENCES. 1. G. B. Airy, Br. Assoc. Adv. Sci. Rep. (1862). 2. S. Timoshenko and J. N. Goodier, Theory of Elasticity (McGraw ...Missing: original | Show results with:original
  8. [8]
    (PDF) Airy, Beltrami, Maxwell, Morera, Einstein and Lanczos ...
    The main purpose of this paper is to revisit the well known potentials, called stress functions, needed in order to study the parametrizations of the stress ...
  9. [9]
    [PDF] Ludwig Prandtl - ICTP
    Oct 6, 2009 · [40] Prandtl, L., Zur Torsion von prismatischen Stäben. Phys. Zeitschr. 4, 1903, p. 758. (In [50]: p. 79). Page 194. 194. [41] Prandtl, L., Zur ...
  10. [10]
    Equilibrium - Continuum Mechanics
    An object is said to be in equilibrium when the right hand sides (RHS) of the equations are zero. 3-D Equilibrium. Stress Equilibrium. The process in 3-D is the ...
  11. [11]
    [PDF] Continuum Mechanics Summary - Quickfem
    This Appendix summarizes the basic relations of three-dimensional continuum mechanics for linear elastic solids. These include strain-displacement, constitutive ...
  12. [12]
    [PDF] Introduction to Linear Elasticity - UCI Mathematics
    Mar 25, 2024 · Mathematically: σ · n = t on ΓN , where σ is the stress tensor, n is the unit outward normal vector to the boundary, and t is a given function.
  13. [13]
    [PDF] Continuum Mechanics - Solids - HAL
    Dec 19, 2018 · These are the Beltrami-Michell equations (1900); they are the necessary conditions of compatibility written in terms of stresses and forces.
  14. [14]
    [PDF] linearized elasticity
    Jul 7, 2023 · LINEARIZED ELASTICITY and define the Airy stress function Φ(r, θ) such that σrr = 1 r. ∂Φ. ∂r. +. 1 r2. ∂2Φ. ∂θ2. + Vf , σθθ = ∂2Φ. ∂r2. + Vf ...
  15. [15]
    [PDF] A note on the compatibility equations for three-dimensional ...
    This is illustrated for both the Saint-Venant and the Beltrami– Michell compatibility equations.
  16. [16]
  17. [17]
    [PDF] Canonical forms for static, steady, and quasistatic equations - arXiv
    Jul 12, 2020 · For instance, the symmetric second order stress field σ, satisfying ∇ · σ = 0 is the double curl of the Beltrami stress tensor [9], still a ...
  18. [18]
    [PDF] Stress Formulation in Three-Dimensional Elasticity
    Stress with traditional methods must be recovered by backcalculation: differentiating either the displacement or the stress function. We have addressed the ...
  19. [19]
    [PDF] Completed Beltrami-Michell Formulation for Analyzing Mixed ...
    A finite elastic continuum consists of a field and a boundary. Stresses and strains must satisfy equilibrium equations and compatibility conditions both in the ...
  20. [20]
    On Reciprocal Diagrams in Space, and their relation to Airy's ...
    J. Clerk Maxwell, M.A., F.R.SS.L. and E.; On Reciprocal Diagrams in Space, and their relation to Airy's Function of Stress, Proceedings of the London Mathe.
  21. [21]
  22. [22]
    Stress Functions in Elasticity Theory - NASA ADS
    The article is devoted to the study of stress functions that make it possible to identically satisfy the equilibrium equations of the classical elasticity ...
  23. [23]
    Three-dimensional stress functions - ScienceDirect
    Four types of stress functions are known for solving elasticity problems: the components of the displacement vector, the components of the Galerkin vector, ...
  24. [24]
    A generalization of the Beltrami stress functions in continuum ...
    1945Kuzmin, R. O.: On Maxwell's and Morera's formulae in the theory of elasticity. C. R. Acad. Sci. URSS (Doklady)49, 326–328. Google Scholar. 1948Weber, C ...
  25. [25]
    [PDF] A GENERALIZATION OF THE BELTRAMI STRESS ... - DTIC
    Thus the Beltrami solution, and hence also the Maxwell and Morera solutions, are incomplete for periphractic regions.
  26. [26]
    [PDF] Completed Beltrami-Michell Formulation in Polar Coordinates
    The stress formulation has two distinct components: the field equations and the boundary conditions. (2) Strain formulation as the coefficient of variational ...
  27. [27]
    IV. On the strains in the Interior of beams - Journals
    On the strains in the Interior of beams. George Biddell Airy. Google Scholar ... Smith C and Gilbert M (2022) The stress function basis of the upper ...<|separator|>
  28. [28]
    [PDF] Module 4 Boundary value problems in linear elasticity - MIT
    Show that the particular choice of stress function given automatically satisfies the stress equilibrium equations in 2D for the case of no body forces. Concept ...<|control11|><|separator|>
  29. [29]
    [PDF] Airy Stress Function For Two Dimensional Inclusion Problems
    It establishes a single governing equation for the plane stress and plane strain conditions by reducing the Navier equation to a form from which the Airy stress ...
  30. [30]
    Approximate expression of the Prandtl membrane analogy in linear ...
    In this work, an approximate general method to obtain a mathematical expression of the linear elastic pure torsion problem is presented. The proposed method is ...
  31. [31]
    None
    ### Summary of Torsion and Prandtl Stress Function (Chapter 10, Pages 292-338, Third Edition)
  32. [32]
    [PDF] Unit 10 St. Venant Torsion Theory
    We satisfy equation (10 - 8) by introducing a Torsion (Prandtl). Stress Function φ (x, y) where: ... see Timoshenko for other relations (Ch. 11). Note: there ...