Fact-checked by Grok 2 weeks ago

Virtual work

Virtual work is a fundamental concept in that involves calculating the work done by forces acting through hypothetical infinitesimal displacements, known as , which are consistent with the geometric constraints of a but do not correspond to actual motion over time. These are instantaneous and independent of time, allowing for the analysis of without considering . The principle of virtual work states that a in static will have zero total virtual work performed by all applied forces for any admissible , providing a powerful method to derive equilibrium equations by eliminating constraint forces that do no virtual work. The origins of the principle trace back to the in , with early formulations appearing in discussions of and levers, evolving through medieval and Latin mechanics during the . During the , varied statements of virtual work laws emerged as distinct principles of , setting the stage for further refinement. In the 18th century, systematized the concept using Leibnizian infinitesimal displacements, while formalized it in his Mécanique Analytique (1788), integrating it into and extending it to dynamics via . By the , the French school applied versions to , solidifying its role in modern engineering and physics. In and , the principle equates external virtual work—done by applied loads, body forces, and surface tractions through virtual displacements—to internal virtual work—arising from stresses and virtual strains within deformable bodies, enabling the computation of deflections and reactions without solving full force systems. For particle systems or under workless constraints (where constraint forces perform no virtual work), it reduces to the condition that the virtual work of applied forces alone is zero, simplifying problems for systems with multiple . This method underpins the in , where discretized virtual displacements lead to stiffness matrices and load vectors for solving complex structures. Applications extend to nonlinear elasticity, , and even metamaterials , highlighting its versatility across scales from particles to continua.

Introduction

Overview

The principle of virtual work refers to the work performed by applied forces on a mechanical system during an virtual displacement that adheres to the system's geometric constraints, without considering time-dependent motion. This concept allows for the analysis of system behavior by imagining small, hypothetical changes in position that respect boundaries or linkages, enabling the evaluation of force effects in a constrained environment. The formal principle of virtual work emerged in the as a tool to streamline the solution of problems, building on ancient and medieval precursors and shifting focus from direct force balancing to energy-like considerations in scenarios. It provides a general framework for handling complex where traditional methods, such as resolving individual forces, become cumbersome due to multiple constraints. In practice, the principle derives conditions by requiring that the total virtual work vanish for any admissible in a constrained , thereby confirming without explicitly solving for constraint reactions. This approach connects directly to static analysis, offering an alternative to Newtonian force equations. Today, the principle underpins variational principles across physics and engineering, forming the basis for advanced techniques like and finite element methods used in and simulation.

Historical Development

The principle of virtual work traces its conceptual roots to ancient discussions of in mechanical systems, particularly through 's analysis of levers in the 4th century BCE. In his Physics and the pseudo-Aristotelian Mechanical Problems, Aristotle conceptualized "power" as the product of and , explaining the balance of a by the inverse proportionality of weights to their velocities, which anticipates the idea of displacements in conditions. This qualitative approach, elaborated in the pseudo-Aristotelian Mechanical Problems, linked to circular motions and ratios, laying an early groundwork for later quantitative formulations without invoking explicit virtual displacements. Medieval scholars advanced these ideas toward a more systematic treatment of , with Jordanus de Nemore's contributions in the 13th century marking a pivotal step. In works such as De ratione ponderis (composed before 1260), Jordanus introduced the concept of "positional gravity," where the effective weight of a body varies with its position on an , effectively employing a precursor to virtual work by considering motions to determine . His demonstrations, including proofs of the through virtual displacements, represented the first mathematical application of such principles in , drawing on decrypted Hellenistic sources and emphasizing the balance of moments in constrained systems. The formalization of the principle emerged in the , beginning with Johann Bernoulli's 1717 letter to Pierre Varignon, where he introduced the notion of "virtual displacements" as infinitesimal variations compatible with system constraints. Bernoulli posited that a body is in equilibrium if the sum of the products of applied forces and their corresponding virtual displacements equals zero, providing a general criterion for static systems beyond simple levers. This , later published in Varignon's Nouvelle mécanique (1725), shifted the focus from actual motions to hypothetical ones, enabling broader applications in . Leonhard Euler extended these ideas in the 1760s through his work on the mechanics of , incorporating displacements into variational principles and demonstrating their utility for systems with multiple . A key milestone came with Joseph-Louis Lagrange's generalization in Mécanique Analytique (1788), which unified into a comprehensive framework for both and , treating constraints via multipliers and deriving from conditions. Earlier, had transitioned the principle to in his Traité de dynamique (1743), applying displacements to by balancing inertial forces with external ones, effectively reducing dynamic problems to static .

Fundamental Concepts

Definition and Basic Principles

Virtual work is a foundational concept in that facilitates the analysis of systems in by considering hypothetical rather than actual motions. Real displacements refer to the actual, finite changes in that a mechanical system undergoes during its physical motion over time, governed by the of applied forces and constraints. In contrast, virtual displacements are infinitesimal, imaginary variations in the system's that are compatible with the existing constraints but do not correspond to evolution; they are "frozen" in time, meaning no actual movement or energy transfer occurs. The of virtual work states that a mechanical system is in the total virtual work performed by all s acting on the system is zero for any admissible . This principle applies to both particles and rigid bodies, providing a scalar condition that simplifies equilibrium analysis without requiring resolutions of s. Virtual work is computed as the of a real with a , or equivalently, a virtual with a real displacement, though the former is standard for equilibrium problems. Forces in the context of virtual work are categorized into applied forces, such as or external loads, which generally contribute to the virtual work, and constraint forces, arising from ideal constraints like rigid links or smooth surfaces that enforce kinematic restrictions. Ideal constraints are assumed to perform no virtual work, meaning the constraint forces are perpendicular to the allowable virtual displacements, ensuring they do not dissipate or input energy in these hypothetical motions. The principle relies on specific assumptions about the system's constraints: they must be , meaning they can be expressed as functions of the without involving velocities, thereby reducing the system's to a set of independent coordinates. Additionally, the constraints are scleronomic, indicating they are time-independent and do not explicitly vary with time, which ensures that virtual displacements remain consistent across instantaneous configurations.

Virtual Displacements

Virtual displacements are infinitesimal, hypothetical changes in the position of a system or its components that occur instantaneously without any passage of time or actual motion, serving as a kinematic in the analysis of . Denoted typically as \delta \mathbf{r} for a particle's \mathbf{r}, these displacements are arbitrary in magnitude and direction but must be consistent with the geometric constraints of the at its current configuration. A key property of virtual displacements is their compatibility with the system's constraints, ensuring that they do not violate any imposed restrictions such as supports, joints, or surfaces. For holonomic constraints defined by equations f(\mathbf{r}, t) = 0, compatibility requires that the virtual displacement satisfies \delta f = \nabla f \cdot \delta \mathbf{r} = 0, meaning \delta \mathbf{r} is perpendicular to the normal vector \mathbf{n} = \nabla f of the constraint surface. This condition guarantees that the displacement remains kinematically admissible, preserving the integrity of the configuration during the hypothetical variation. Kinematically, virtual displacements represent tangent vectors to the configuration manifold of the system, which is the space of all allowable configurations defined by the s. In this geometric framework, the set of all possible virtual displacements at a given point forms the , capturing the instantaneous directions of permissible motion without altering the constraint equations. For a two-dimensional rigid body, such as a beam pivoted at one end, virtual displacements consist of infinitesimal rotations \delta \theta about the pivot and translations perpendicular to any additional constraints, like a fixed support that prohibits linear motion at the pivot point. In the case of a ladder leaning against a wall, a compatible virtual displacement might involve a small angular variation \delta \phi that adjusts the contact points while maintaining surface adherence. The collection of virtual displacements spans the allowable motion space of the system, with their corresponding to the , which quantify the number of independent parameters needed to specify the . For instance, a in three dimensions has three , and its virtual displacements fill the full three-dimensional , whereas a constrained particle on a surface has two, restricted to the . This spanning property allows virtual displacements to systematically explore conditions within the reduced dimensionality imposed by constraints.

Mathematical Formulation

General Expression for Virtual Work

The general expression for virtual work in a mechanical describes the infinitesimal work performed by forces acting through compatible virtual displacements. For a consisting of N particles, the virtual work \delta W is given by \delta W = \sum_{i=1}^N \mathbf{F}_i \cdot \delta \mathbf{r}_i, where \mathbf{F}_i denotes the on the i-th particle and \delta \mathbf{r}_i is its , which must be consistent with the 's kinematic constraints. This summation extends naturally to multi-body systems, where it accounts for all particles within rigid bodies or interconnected components, treating rigid bodies as collections of particles with internal constraints that contribute no net virtual work. In continuous media, such as deformable solids, the virtual work principle equates the external virtual work to the internal virtual work. The external virtual work is \delta W_\text{ext} = \int_V \mathbf{b} \cdot \delta \mathbf{u} \, dV + \int_S \mathbf{t} \cdot \delta \mathbf{u} \, dS, where \mathbf{b} is the density, \mathbf{t} is the surface traction, \delta \mathbf{u} is the field, V is the , and S is the surface. The internal virtual work is \delta W_\text{int} = \int_V \boldsymbol{\sigma} : \delta \boldsymbol{\epsilon} \, dV, where \boldsymbol{\sigma} is the tensor and \delta \boldsymbol{\epsilon} is the tensor derived from the virtual displacement field. Virtual work is inherently a scalar , representing the first-order to the actual work integral along an infinitesimal path in the configuration space, obtained by linearizing the displacement about the current position. For a system in equilibrium, the total virtual work vanishes for any admissible virtual displacement, expressed as \delta W = \delta W_\text{applied} + \delta W_\text{constraint} = 0, where \delta W_\text{applied} arises from external and body forces, and \delta W_\text{constraint} from reaction forces at supports or joints. The units of virtual work are those of energy, such as joules in the International System of Units (SI).

Static Equilibrium Applications

In static equilibrium, the principle of virtual work states that for a system at rest under the action of forces, the total virtual work performed by all applied forces through any admissible virtual displacement is zero. This condition, derived from the general expression for virtual work by setting it to zero in the absence of motion, ensures that the system remains balanced. Mathematically, for a system of particles or rigid bodies, it is expressed as \sum \mathbf{F} \cdot \delta \mathbf{r} = 0, where \mathbf{F} represents the applied forces and \delta \mathbf{r} are infinitesimal virtual displacements consistent with the system's constraints. This equation implies the balance of both forces and moments, as virtual displacements can be chosen as pure translations (yielding \sum \mathbf{F} = 0) or infinitesimal rotations (yielding \sum \mathbf{M} = 0). A key advantage of this approach is the reduction of equations by eliminating unknown forces. By selecting virtual displacements \delta \mathbf{r} that are orthogonal to the directions of forces—meaning they satisfy the geometric constraints without violating them—the contributions from reactions (such as forces or tensions) vanish, as their with \delta \mathbf{r} is zero. This leaves only the applied s in the equations, significantly simplifying the analysis for systems with multiple constraints. For instance, in the case of a particle resting on a plane under and a , an admissible virtual displacement might include a small vertical component \delta y and \delta x. The virtual work is then mg \delta y - N \delta y + F \delta x = 0, where N is the ; choosing \delta y = 0 (horizontal displacement only) isolates the horizontal balance, while vertical shows N = mg directly, canceling the without solving for it explicitly. Compared to traditional free-body diagrams, which require isolating each body and solving for all reaction components, the virtual work method handles complex constraints more efficiently by focusing solely on applied forces and compatible displacements. It is particularly useful for interconnected rigid bodies, where drawing complete free-body diagrams becomes cumbersome due to numerous unknowns. However, the principle assumes ideal, workless constraints without and scleronomic (time-independent) geometry, limiting its direct application to systems involving dissipative forces or moving boundaries.

Classical Applications in Statics

Constraint Forces

In the principle of virtual work applied to static , constraint forces arising from ideal perform no virtual work for any compatible \delta \mathbf{r}. This follows from the condition, where the constraint \mathbf{F}_c is to the allowable virtual displacements, yielding \mathbf{F}_c \cdot \delta \mathbf{r} = 0. Such ideal , common in scleronomic systems without or other dissipative effects, allow the virtual work to simplify by eliminating constraint forces entirely, focusing only on applied forces. Constraint forces can be identified as residing in the to the of admissible displacements. In geometric terms, if the constraints define a manifold, the displacements \delta \mathbf{r} are to this manifold, and \mathbf{F}_c lies normal to it, ensuring zero . This underpins the efficiency of the work method in reducing the for analysis. A practical example occurs in analyzing beams supported at multiple points, where virtual rotations can isolate reaction moments. Consider a fixed beam with a reaction moment M_A at support A; imposing a small virtual rotation \delta \theta about A (while keeping other points fixed) results in the virtual work equation M_A \delta \theta + \delta W_{\text{applied}} = 0, solving directly for M_A = -\delta W_{\text{applied}} / \delta \theta, as other constraint forces contribute zero work under this specific displacement. For non-ideal constraints, such as those involving dissipation (e.g., ), constraint forces may perform non-zero virtual work, complicating the analysis; however, the virtual work principle typically assumes constraints to maintain simplicity. In practical computations, specific virtual displacements \delta \mathbf{r} are selected to nullify the work of all forces except the desired constraint force, effectively isolating it within the equation. This targeted choice, often a unit displacement in the direction of the unknown, facilitates solving for individual reactions without full system resolution.

Law of the Lever

The classical law of the lever describes the condition for a rigid bar pivoted at a , with two point masses m_1 and m_2 attached at horizontal distances d_1 and d_2 from the pivot, respectively, under the influence of . In this setup, the bar remains horizontal in when the weights balance about the , and the principle of virtual work provides a direct method to derive this condition without resolving individual forces. To apply the principle, consider a consisting of an \delta\theta of the bar about the , consistent with the kinematic constraints. This produces vertical virtual displacements \delta y_1 = -d_1 \delta\theta for the first mass (downward) and \delta y_2 = +d_2 \delta\theta for the second mass (upward), assuming small angles where the vertical component approximates the . The corresponding virtual work done by is then \delta W = m_1 g \delta y_1 + m_2 g \delta y_2 = -m_1 g d_1 \delta\theta + m_2 g d_2 \delta\theta. For , the total virtual work must vanish for any such admissible \delta\theta, yielding -m_1 g d_1 + m_2 g d_2 = 0, or equivalently, m_1 d_1 = m_2 d_2. This equilibrium relation, m_1 d_1 = m_2 d_2, is precisely , which states that two magnitudes are in equilibrium at distances reciprocally proportional to their weights, as proven geometrically in his work On the Equilibrium of Planes (Propositions 6 and 7). ' formulation, dating to around 250 BCE, predates the principle of virtual work but serves as a key precursor, later formalized through virtual displacements by eighteenth-century mechanicians like and to encompass broader static systems. The derivation extends naturally to unequal-arm levers, where d_1 \neq d_2, maintaining the balance condition m_1 d_1 = m_2 d_2 as an expression of torque equilibrium about the (\tau_1 = \tau_2, with \tau = m g d). This torque interpretation underscores the lever's role in , where a smaller at greater balances a larger load at shorter , without altering the virtual work approach. The reaction at the contributes no virtual work, as the virtual displacement there is zero.

Gear Trains

In gear trains, a series of meshed transmits while maintaining static under applied loads, assuming frictionless operation and no slip at the contact points. The setup involves with pitch radii r_i, where the virtual angular displacements \delta \theta_i between consecutive satisfy \frac{\delta \theta_i}{\delta \theta_{i+1}} = -\frac{r_{i+1}}{r_i}, reflecting the geometric constraint that the arc lengths at the pitch circles are equal in magnitude but opposite in direction. The principle of virtual work applied to such systems states that for equilibrium, the total virtual work done by all external torques is zero: \sum \tau_i \delta \theta_i = 0, where \tau_i are the applied torques on each gear. Substituting the kinematic relations between the \delta \theta_i yields the equilibrium condition that torque ratios are inverse to the speed ratios, with the magnitude of the torque amplification equal to the gear ratio defined by the number of teeth N; for a simple pair, \frac{\tau_1}{\tau_2} = -\frac{N_1}{N_2}. Consider a simple two-gear train where gear 1 (driver, with N_1 teeth) meshes with gear 2 (driven, with N_2 teeth), and an input \tau_1 is applied to gear 1. A compatible \delta \theta_1 of gear 1 induces \delta \theta_2 = -\frac{N_1}{N_2} \delta \theta_1 on gear 2. The becomes \tau_1 \delta \theta_1 + \tau_2 \delta \theta_2 = 0, leading to \tau_2 = \tau_1 \frac{N_2}{N_1} (magnitude), ensuring the output torque balances the input through the gear ratio. This analysis holds under the idealization of rigid gears with instantaneous point contact, neglecting any energy losses.

Dynamic Extensions

Dynamic Equilibrium for Rigid Bodies

In dynamic equilibrium, the principle of extends to rigid bodies undergoing accelerated motion by requiring that the total virtual work performed by both applied forces and forces vanishes for any admissible consistent with the kinematic s of the system. This formulation accounts for the body's nonzero , differing from static cases where only applied forces contribute to zero virtual work. The terms effectively balance the applied loads during motion, enabling analysis without explicit resolution of constraint forces. For a single rigid body, the configuration space consists of 6 degrees of freedom: 3 for translational motion of the center of mass and 3 for rotational orientation. Virtual displacements δr and δθ are thus defined within this space, ensuring rigid body constraints (constant distances between points) are preserved, such that the virtual work of internal constraint forces is zero. The translational component of the virtual work equation is given by \delta W_\text{trans} = \sum (\mathbf{F} - m \mathbf{a}) \cdot \delta \mathbf{r} = 0, where the sum is over the body's mass elements or equivalently the net applied force F, total mass m, acceleration a of the center of mass, and compatible virtual displacement δr. Similarly, the rotational component is \delta W_\text{rot} = \sum (\boldsymbol{\tau} - \mathbf{I} \boldsymbol{\alpha}) \cdot \delta \boldsymbol{\theta} = 0, with τ representing net applied torques, I the inertia tensor about the center of mass, α the angular acceleration, and δθ the virtual angular displacement. In multi-body systems, such as chains of connected rigid elements (e.g., linkages or robotic arms), the principle applies by summing the virtual work contributions over all bodies, incorporating joint constraints that couple their motions. The overall becomes a generalized form aggregating translational and rotational terms across the n bodies, yielding 6n equations that describe the in terms of . This approach eliminates the need to compute individual constraint reactions at , as their virtual work is zero by construction. This virtual work-based dynamic equilibrium is mathematically equivalent to Newton's laws of motion for rigid bodies but offers a constraint-free perspective, projecting the equations onto the independent degrees of freedom and simplifying analysis for complex geometries or mechanisms.

D'Alembert's Principle

D'Alembert's principle extends the concept of virtual work from statics to dynamics by incorporating inertia forces as fictitious forces that enable the treatment of dynamic systems as if they were in equilibrium. The principle states that for a system of particles in dynamic equilibrium, the total virtual work done by the applied forces and the inertia forces is zero: \sum_i (\mathbf{F}_i - m_i \mathbf{a}_i) \cdot \delta \mathbf{r}_i = 0, where \mathbf{F}_i is the applied force on the i-th particle, m_i is its mass, \mathbf{a}_i is its acceleration, and \delta \mathbf{r}_i is the virtual displacement consistent with the constraints. The term -m_i \mathbf{a}_i represents the inertia force, which balances the applied forces in the virtual work calculation. This formulation was originally presented by in his 1743 work Traité de dynamique, dans lequel les loix de l'équilibre & du mouvement des corps sont réduites au plus petit nombre possible, where he sought to unify the laws of and motion under a single framework inspired by earlier ideas on displacements. D'Alembert's approach emphasized reducing the complexity of dynamic problems by analogy to , avoiding direct appeals to Newton's second law in constrained systems. A key advantage of is that it transforms dynamic problems into equivalent static equilibrium problems by including inertia forces, which simplifies the analysis of systems with constraints since constraint forces do no virtual work and can often be eliminated from the equations. This method is particularly useful for systems involving multiple or non-Cartesian coordinates, as it allows the use of virtual displacements to derive without explicitly solving for constraint reactions. For example, in Atwood's machine with two masses M > m connected by a over a , assuming inextensible string constraint, a δs downward for M corresponds to -δs upward for m. The virtual work is [(M g) δs + (m g) (-δs) - (M + m) a δs] = 0, where a is the acceleration magnitude, yielding a = g (M - m)/(M + m). The tensions in the do no virtual work due to the constraint-compatible displacements and are eliminated from the equation. This illustrates how the principle incorporates dynamics via inertia while treating the system as equilibrated, without needing to solve for constraint forces. D'Alembert's principle serves as a direct precursor to the development of Lagrange's , providing the foundational virtual work framework that later generalized using and the function in his 1788 Mécanique Analytique.

Generalized Inertia Forces

In the dynamic analysis of rigid bodies using the principle of virtual work, generalized inertia forces account for the inertial effects that arise during motion, extending the static equilibrium condition to include acceleration-dependent terms. These forces are incorporated such that the total virtual work, including contributions from applied forces and , vanishes for admissible virtual displacements. This approach, rooted in , treats as equivalent to additional forces in a quasi-static framework. For a single , the force at any point includes the translational component -m \mathbf{a}_G, where m is the and \mathbf{a}_G is the of of , along with rotational contributions such as the centripetal term -\boldsymbol{\omega} \times (\boldsymbol{\omega} \times (\mathbf{r} - \mathbf{r}_G)), where \boldsymbol{\omega} is the and \mathbf{r} - \mathbf{r}_G is the relative to of . However, in the work formulation, these are aggregated through the with displacements: the translational work is -m \mathbf{a}_G \cdot \delta \mathbf{r}_G, and rotational effects manifest as torques acting through angular displacements \delta \boldsymbol{\theta}. This ensures that the principle captures both linear and angular inertial effects without decomposing into separate particle motions. When employing q_j to describe the 's configuration, the virtual work due to forces takes the form \delta W_{\text{inertia}} = \sum_j Q_j^{\text{in}} \delta q_j, where Q_j^{\text{in}} = -\frac{d}{dt} \left( \frac{\partial T}{\partial \dot{q}_j} \right) + \frac{\partial T}{\partial q_j} and T is the total of the . This expression arises from expressing the accelerations in terms of the generalized coordinates and their derivatives, projecting onto the virtual displacements, and combining with applied generalized forces to yield the . For rigid bodies, T includes terms like \frac{1}{2} m v_G^2 + \frac{1}{2} \boldsymbol{\omega} \cdot \mathbf{I} \boldsymbol{\omega}, where \mathbf{I} is the tensor, allowing efficient computation even for complex geometries. In multi-body systems, such as those connected by joints in linkages, the generalized inertia forces exhibit between bodies due to shared constraints and kinematic dependencies. Jacobians relating Cartesian velocities to generalized coordinate rates propagate inertial effects across joints, resulting in a that couples the \ddot{q}_j terms in the dynamic equations. For instance, in a planar linkage, the inertia contribution from one link's affects the translational inertia of adjacent links through revolute or prismatic joints. A representative example is the slider-crank mechanism, where the crank angle \theta serves as a generalized coordinate. The inertia effects arise from the kinetic energy T of the , connecting rod, and slider, leading to generalized inertia forces Q_\theta^{\text{in}} = -\frac{d}{dt} \left( \frac{\partial T}{\partial \dot{\theta}} \right) + \frac{\partial T}{\partial \theta}, with T encompassing rotational inertia of the (J \dot{\theta}^2 / 2) and coupled translational terms for the rod and slider. This coupling produces inertia torques that vary with \theta and \dot{\theta}, influencing the input required at the . For rigid bodies analyzed in non-inertial reference , such as rotating or accelerating attached to a moving component, additional fictitious forces must be included in the virtual work. These comprise centrifugal forces m \boldsymbol{\omega} \times (\boldsymbol{\omega} \times \mathbf{r}'), Coriolis forces -2m \boldsymbol{\omega} \times \mathbf{v}' (where primed quantities are relative to the ), and Euler forces -m \dot{\boldsymbol{\omega}} \times \mathbf{r}', contributing virtual work terms analogous to the inertial ones: \delta W_{\text{fict}} = \sum - \mathbf{F}_{\text{fict}} \cdot \delta \mathbf{r}'. This ensures the principle remains valid by treating fictitious effects as effective forces in the .

Deformable Bodies

Principle for Deformable Systems

The principle of virtual work for deformable systems generalizes the formulation by incorporating the effects of internal deformations, enabling the analysis of structures where bodies experience straining under loads. This extension, rooted in the works of Lagrange, allows for the of continuous media to be expressed through energy balances involving stresses and strains. In deformable systems, the total virtual work \delta W comprises external and internal contributions. The external virtual work arises from forces and surface tractions acting through a field \delta \mathbf{u}(\mathbf{x}): \delta W_\text{ext} = \int_V \boldsymbol{\rho} \mathbf{b} \cdot \delta \mathbf{u} \, dV + \int_S \mathbf{t} \cdot \delta \mathbf{u} \, dS, where \boldsymbol{\rho} is the , \mathbf{b} the per unit , \mathbf{t} the surface traction vector, V the volume of the , and S its surface. The internal virtual work accounts for the stresses within the : \delta W_\text{int} = -\int_V \boldsymbol{\sigma} : \delta \boldsymbol{\epsilon} \, dV, with \boldsymbol{\sigma} the Cauchy stress tensor and \delta \boldsymbol{\epsilon} the virtual strain tensor derived from \delta \mathbf{u}. Equilibrium holds when the total virtual work vanishes for all admissible virtual displacement fields \delta \mathbf{u} that are kinematically compatible, meaning they satisfy the essential boundary conditions and allow computation of compatible virtual strains \delta \boldsymbol{\epsilon} = \frac{1}{2} (\nabla \delta \mathbf{u} + (\nabla \delta \mathbf{u})^T): \delta W = \delta W_\text{ext} + \delta W_\text{int} = 0. This condition ensures that the real stress field is in balance with the applied loads without requiring pointwise enforcement of differential equations. The virtual displacement field \delta \mathbf{u}(\mathbf{x}) must be sufficiently smooth and compatible with the deformation kinematics of the system, such as continuity across element boundaries in discretized models, to guarantee that the virtual strains represent possible infinitesimal changes in shape. Applications of this principle are central to , particularly for analyzing under and , trusses with axial deformations, and general in finite element formulations. For instance, in a simply supported , applying admissible virtual displacements yields the governing for deflection, while in trusses, it facilitates efficient computation of member elongations and joint displacements. In settings, it underpins the weak form of the equations used in numerical simulations of bodies. The principle assumes small deformations, where virtual displacements do not significantly alter the body's geometry or the definitions of stress and strain measures; linear elasticity is not strictly required, as the formulation applies to nonlinear materials provided the virtual fields remain consistent with the kinematics.

Principle of Virtual Displacements

The principle of virtual displacements provides a kinematic formulation of the virtual work principle specifically for deformable bodies in equilibrium, extending the general approach for deformable systems by focusing on admissible virtual displacements. In this method, an arbitrary virtual displacement field \delta \mathbf{u} is selected that satisfies the kinematic boundary conditions of the problem, such as fixed displacements on relevant surfaces, ensuring compatibility with the constraints of the deformable body. The principle states that for equilibrium, the total virtual work done by external and internal forces through these virtual displacements must vanish: \delta W_{\text{ext}} + \delta W_{\text{int}} = 0. The external work \delta W_{\text{ext}} arises from body forces and surface tractions acting through the virtual displacements, expressed as \delta W_{\text{ext}} = \int_V \mathbf{f} \cdot \delta \mathbf{u} \, dV + \int_{\Gamma_t} \mathbf{t} \cdot \delta \mathbf{u} \, d\Gamma, where \mathbf{f} are body forces per unit volume, \mathbf{t} are tractions on the traction boundary \Gamma_t, V is the volume of the body, and \Gamma denotes the surface. The internal virtual work \delta W_{\text{int}} accounts for the stresses within the deformable material deforming through the compatible virtual \delta \boldsymbol{\epsilon}, given by \delta W_{\text{int}} = -\int_V \boldsymbol{\sigma} : \delta \boldsymbol{\epsilon} \, dV, where \boldsymbol{\sigma} is the and the colon denotes the . This enforces in a variational sense, integrating over the rather than . Applying yields the weak form of the equations, which reduces the order of derivatives required compared to forms and naturally incorporates conditions. This weak form serves as the foundational framework for displacement-based finite element methods, where the displacements are approximated by shape functions within elements to solve value problems numerically. A representative example is the analysis of a cantilever beam under a tip load, modeled using Euler-Bernoulli theory. By choosing a virtual displacement field corresponding to a unit rotation at the free end (satisfying the fixed-end boundary condition), the principle equates the external virtual work from the tip load to the internal virtual work from bending stresses, yielding the deflection equation \delta = \frac{P L^3}{3 E I}, where P is the load, L the length, E the modulus, and I the moment of inertia. This approach directly computes deflections without solving differential equations./03%3A_Analysis_of_Statically_Indeterminate_Structures/08%3A_Deflections_of_Structures-_Work-Energy_Methods/8.01%3A_Virtual_Work_Method) The kinematic nature of this makes it particularly suited for value problems, as it directly uses functions akin to the primary variables (), facilitating straightforward in methods that prescribe conditions on while treating conditions (tractions) variationally.

Principle of Virtual Forces

The principle of forces, also known as the principle of complementary work, provides a static formulation for ensuring strain-displacement in deformable bodies under static loading. It posits that for a body with given real displacements \mathbf{u} and corresponding strains \boldsymbol{\varepsilon}, the internal complementary work performed by these strains on any admissible \boldsymbol{\delta \sigma} equals the external complementary work performed by the real displacements on the associated tractions \boldsymbol{\delta t} and body forces \boldsymbol{\delta b}. Admissible fields must satisfy conditions: \nabla \cdot \boldsymbol{\delta \sigma} + \boldsymbol{\delta b} = \mathbf{0} in the volume V and \boldsymbol{\delta t} = \boldsymbol{\delta \sigma} \cdot \mathbf{n} on the surface S, where \mathbf{n} is the . The governing is \int_V \boldsymbol{\delta \sigma} : \boldsymbol{\varepsilon} \, dV = \int_S \boldsymbol{\delta t} \cdot \mathbf{u} \, dS + \int_V \boldsymbol{\delta b} \cdot \mathbf{u} \, dV, which holds for all such equilibrated virtual fields. This approach is particularly advantageous when stress or traction boundary conditions are prescribed, as it directly incorporates them without requiring kinematic assumptions. As the adjoint to the principle of virtual displacements, the principle of virtual forces shifts focus from kinematic compatibility to static equilibrium of virtual fields, enabling solutions where displacements are harder to parameterize. In practice, virtual stress fields \boldsymbol{\delta \sigma} are constructed to be self-equilibrated, often via finite element approximations or analytical patterns that satisfy the homogeneous equilibrium equations. For instance, in linear elasticity, assuming Hooke's law \boldsymbol{\sigma} = \mathbf{C} \boldsymbol{\varepsilon}, the principle enforces the inverse relation through variation, ensuring the real stress state derives from compatible strains. This duality facilitates hybrid methods in computational mechanics, where one principle handles equilibrium and the other compatibility. A representative application appears in the analysis of statically indeterminate trusses, where the force method employs virtual force patterns to resolve redundant member forces. The structure is first reduced to a statically determinate primary system by removing redundant members or supports; compatibility conditions are then enforced using the principle, with virtual unit forces applied along the redundant directions to generate equilibrated internal force patterns \delta N_k in each member. The flexibility coefficients are computed as f_{ij} = \sum \frac{\delta N_i \delta N_j L}{A E}, where L, A, and E are member length, area, and modulus, respectively; solving \sum f_{ij} X_j = -\Delta_i^0 yields the redundant forces X_j, revealing the full member force distribution. This reveals how virtual force patterns directly contribute to determining the actual internal forces by balancing compatibility with the primary equilibrium solution./01%3A_Chapters/1.10%3A_Force_Method_of_Analysis_of_Indeterminate_Structures) The Hellinger-Reissner extends this framework into a mixed for linear elastostatics, treating displacements \mathbf{u} and stresses \boldsymbol{\sigma} as independent variables within a single functional. The principle derives from combining the virtual forces and displacements principles, yielding the stationary condition \Pi(\mathbf{u}, \boldsymbol{\sigma}) = \int_V \left[ \boldsymbol{\sigma} : (\nabla^s \mathbf{u} - \mathbf{C}^{-1} \boldsymbol{\sigma}) \right] dV - \int_{S_u} \mathbf{t} \cdot (\mathbf{u} - \overline{\mathbf{u}}) dS - \int_{S_t} \overline{\mathbf{t}} \cdot \mathbf{u} \, dS = 0, where \nabla^s denotes the symmetric , \mathbf{C}^{-1} is the tensor, and S_u, S_t are - and traction-prescribed boundaries. Variation with respect to \boldsymbol{\sigma} recovers the principle of virtual forces (compatibility), while variation with respect to \mathbf{u} recovers the principle of virtual s (equilibrium). This extension is foundational for mixed finite element methods, avoiding locking in incompressible materials. In limit analysis and , the principle underpins the static (lower-bound) , where equilibrated virtual stress fields \boldsymbol{\delta \sigma} (scaled to the collapse load factor) that nowhere violate the yield criterion provide a safe estimate of the ultimate load. For rigid-plastic materials, admissible stress fields satisfying and yield bounds f(\boldsymbol{\sigma}) \leq 0 \lambda \geq \lambda_c, with equality at the exact collapse mechanism. Applications include shakedown analysis for cyclic loading, ensuring long-term structural integrity without excessive plastic deformation, and optimizing plastic in frames or .

Advanced Formulations

Equivalence to Equilibrium Equations

The principle of virtual displacements states that for a deformable in , the virtual work done by internal stresses equals the virtual work done by external forces and forces for any admissible virtual displacement field δu compatible with the boundary conditions. To establish its equivalence to the strong form of the equilibrium equations, consider the virtual work expression: \delta W = \int_V \boldsymbol{\sigma} : \delta \boldsymbol{\varepsilon} \, dV - \int_V \mathbf{b} \cdot \delta \mathbf{u} \, dV - \int_{S_t} \mathbf{t} \cdot \delta \mathbf{u} \, dS = 0, where \boldsymbol{\sigma} is the , \delta \boldsymbol{\varepsilon} = \sym(\nabla \delta \mathbf{u}) is the virtual strain tensor, \mathbf{b} are body forces per volume, and \mathbf{t} are prescribed surface tractions on the traction boundary S_t. Applying the and to the internal virtual work term yields: \int_V \boldsymbol{\sigma} : \nabla \delta \mathbf{u} \, dV = \int_V \delta \mathbf{u} \cdot (\div \boldsymbol{\sigma}) \, dV + \int_S (\boldsymbol{\sigma} \cdot \mathbf{n}) \cdot \delta \mathbf{u} \, dS, assuming \boldsymbol{\sigma} is symmetric (as required by angular momentum balance). Substituting back into the virtual work equation and collecting terms gives: \int_V \delta \mathbf{u} \cdot (\div \boldsymbol{\sigma} + \mathbf{b}) \, dV + \int_{S_t} \delta \mathbf{u} \cdot (\boldsymbol{\sigma} \cdot \mathbf{n} - \mathbf{t}) \, dS + \int_{S_u} \delta \mathbf{u} \cdot (\boldsymbol{\sigma} \cdot \mathbf{n}) \, dS = 0, where S_u is the displacement boundary (with \delta \mathbf{u} = 0 there). Since \delta \mathbf{u} is arbitrary within the space of smooth, kinematically admissible fields (typically C^1 continuous and vanishing on S_u), the integrands must vanish pointwise: \div \boldsymbol{\sigma} + \mathbf{b} = 0 in the volume V (Cauchy's equilibrium equation) and \boldsymbol{\sigma} \cdot \mathbf{n} = \mathbf{t} on S_t. The assumptions include sufficiently smooth fields for integration by parts to hold, such as C^1 continuity for \delta \mathbf{u} and twice-differentiable \boldsymbol{\sigma}. Dually, the principle of virtual forces (or complementary virtual work) recovers the strain equations. In this formulation, virtual stress fields \delta \boldsymbol{\sigma} in (satisfying \div \delta \boldsymbol{\sigma} + \delta \mathbf{b} = 0 and compatible boundary tractions) are applied to real strains \boldsymbol{\varepsilon}, yielding \int_V \boldsymbol{\varepsilon} : \delta \boldsymbol{\sigma} \, dV = \int_V \delta \mathbf{b} \cdot \mathbf{u} \, dV + \int_{S_t} \delta \mathbf{t} \cdot \mathbf{u} \, dS. on this expression, under similar smoothness assumptions (e.g., C^1 for strains and virtual stresses), leads to the condition that \boldsymbol{\varepsilon} = \sym(\nabla \mathbf{u}) must hold to ensure compatibility, preventing interpenetration and maintaining in the deformation. Rigorous proofs establishing these equivalences in the framework of and , including handling of boundary conditions and field regularity, were advanced by researchers such as Eric Reissner in the mid-20th century through variational theorems that unified displacement and stress formulations.

Alternative Forms and Variations

One prominent alternative formulation of virtual work arises in the variational context, where it connects to through the condition that the variation of integral vanishes: \delta \int L \, dt = 0, with L denoting the . This leads directly to the Euler-, providing a foundational framework for deriving in conservative systems and highlighting virtual work as a discrete instantiation of broader variational . Such a unifies virtual work with principles, enabling applications in fields like optimal control and field theories. For systems subject to non-holonomic constraints, which cannot be expressed as time-independent position relations, Gauss's principle of least constraint offers a key extension of virtual work. This principle posits that the actual motion minimizes a quadratic form involving the deviations of accelerations from unconstrained values, weighted by masses, under virtual displacements compatible with the constraints; equivalently, it minimizes the virtual work associated with inertia and constraint forces. Formulated originally by Carl Friedrich Gauss in 1829, it applies to nonlinear non-holonomic systems by adjusting the virtual work to account for velocity-dependent constraints, yielding equations of motion without explicit Lagrange multipliers in some cases. In , virtual work is extended to a covariant form using four-vectors, where requires the Minkowski inner product of the and an infinitesimal four-displacement to vanish: F^\mu \delta x_\mu = 0. The , defined as the proper-time of the , K^\mu = \frac{dP^\mu}{d\tau}, ensures Lorentz invariance, with the principle adapting classical virtual work to account for relativistic effects like in particle dynamics. This formulation is particularly useful in high-energy physics for analyzing constrained motions in systems or relativistic continua. Computational implementations of work often employ discrete variants in multibody dynamics software, discretizing virtual displacements over time steps to generate algebraic equations for simulating complex assemblies of rigid and flexible components. These methods, rooted in variational integrators, preserve and in numerical schemes, facilitating analysis in tools like Adams or Simscape Multibody for automotive and . By formulating joint forces via discrete virtual power, such approaches handle large-scale systems efficiently without continuous . Despite its versatility, the principle of virtual work faces limitations in dissipative systems, where non-ideal constraints like or perform non-zero virtual work, invalidating the assumption that constraint forces contribute nothing to the total virtual work. In such scenarios, the principle fails to directly yield correct equations unless modified, as seen in frictional contacts or viscoelastic materials. Alternatives include energy-based methods, such as the integrated into , which accounts for rates and provides a more robust framework for non-conservative dynamics. These extensions, often drawing on extended Noether theorems, better capture irreversible processes while maintaining variational structure.

References

  1. [1]
    [PDF] Physics 5153 Classical Mechanics Principle of Virtual Work
    The concept of virtual work is centered on the idea of calculating the amount of work done on a system of particles through a virtual displacement. We will ...
  2. [2]
    [PDF] The Principle of Virtual Work - Duke People
    Virtual work is the work done by a real force acting through a virtual displace- ment or a virtual force acting through a real displacement. A virtual ...
  3. [3]
    History of Virtual Work Laws - SpringerLink
    The book begins with the first documented formulations of laws of virtual work in the IV century BC in Greece and proceeds to the end of the XIX century AD in ...
  4. [4]
    [PDF] Lecture 4 - The Principle of Virtual Work - MIT OpenCourseWare
    The principle of virtual work states that for any compatible virtual displacement field imposed on the body in its state of equilibrium, the total internal ...
  5. [5]
    [PDF] The Origins of Analytic Mechanics in the 18th century - HAL-SHS
    Analytical mechanics arose from non-Newtonian principles like least action, shifting from geometric to analytical, and was developed by Lagrange in the 18th ...
  6. [6]
  7. [7]
    [PDF] Equilibrium and Forces: from Aristotle to Lagrange - Amazon AWS
    “This insight is, indeed, the seed from which will come out, through a twenty century development, the powerful ramifications of the Principle of virtual ...<|separator|>
  8. [8]
    [PDF] Who Was the First to Formulate the Principle of Virtual Work? - HAL
    Abstract In this chapter, we try to reconstruct the history of the Principle of Virtual. Work (PVW) in mechanics to find its original formulation.
  9. [9]
    History of Virtual Work Laws | Request PDF - ResearchGate
    This chapter is devoted to the debate in Italy on the principle of virtual velocities as presented in Lagrange's Méchanique analitique of 1788.
  10. [10]
    [PDF] The Emergence of the Principle of Virtual Velocities
    The historical path to Lagrange's statement of the principle of virtual velocities has been two-millennium long, a facet of what Benvenuto [BEN 91] calls “ ...
  11. [11]
    [PDF] History of classical mechanics
    His statics is based on the principle of virtual work. While this principle may be traced from antiquity through the work of Jordanus, and while it had been ...
  12. [12]
    Confusion with Virtual Displacement - Physics Stack Exchange
    Jan 26, 2019 · So in a nutshell, a finite virtual displacement is a displacement of position that doesn't violate the constraints and is frozen in time. See ...What exactly is a virtual displacement in classical mechanics?Virtual displacements - classical mechanics - Physics Stack ExchangeMore results from physics.stackexchange.com
  13. [13]
    Virtual displacements and Virtual Work - Physics Forums
    Aug 3, 2011 · A 'virtual displacement' is said to be consistent with the constraints on the system, occurs without the passage of time, is infinitesimal.Virtual displacement vs. differential displacement - Physics ForumsVirtual Displacement: Definition & Meaning - Physics ForumsMore results from www.physicsforums.com
  14. [14]
    What is the difference between variation in displacement and virtual ...
    Apr 6, 2019 · A virtual displacement is any field of displacements, which satisfy the boundary conditions, but is imaginary (do not actually occur).
  15. [15]
    [PDF] Principle of Virtual Work - Department of Civil Engineering, IIT Bombay
    Principle of Virtual Work: • If a particle is in equilibrium, the total virtual work of forces acting on the particle is zero for any virtual displacement.
  16. [16]
    [PDF] CHAPTER 6 LAGRANGE'S EQUATIONS (Analytical Mechanics)
    Need to define concept of workless constraints. 6.4 Virtual Work. Definition: A workless constraint is any constraint such that the virtual work (work done ...<|control11|><|separator|>
  17. [17]
    Principle of virtual work - Stéphane Caron
    Sep 28, 2021 · The principle of virtual work is not always applicable: it only works with ideal constraints where forces do not work along virtual ...
  18. [18]
    [PDF] Lecture 4: Constraints, Virtual Work, etc. - LIGO-Labcit Home
    Holonomic constraints are further classified as time independent or scleronomic if time does not appear in Eq. (1), and time dependent or rheonomic if time ...
  19. [19]
    [PDF] Classical Mechanics Virtual Work & d'Alembert's Principle
    Aug 15, 2016 · d'Alembert's principle, developed from an idea originally due to Bernoulli ... We will now derive the Euler-Lagrange equation from d'Alembert's ...
  20. [20]
    On Virtual Displacement and Virtual Work in Lagrangian Dynamics
    It is observed that for holonomic, scleronomous constraints, the virtual displacements are the displacements allowed by the constraints. However, this is not so ...
  21. [21]
    The Principle of Virtual Work - Engineering at Alberta Courses
    The statement of the principle of virtual work usually involves the phrase “virtual displacement field,” which is designed to engage the intuition by attempting ...
  22. [22]
    virtual work - Welcome to AE Resources
    PRINCIPLE OF VIRTUAL WORK​​ For a system at equilibrium in which constraint forces are normal to the virtual displacements, the principle of virtual work ...
  23. [23]
    [PDF] 8.5 Virtual Work
    May 8, 2013 · The principle of virtual work (or principle of virtual displacements) I: if a particle is in equilibrium under the action of a number of ...
  24. [24]
    [PDF] Principle of Virtual Displacements in Structural Dynamics
    In other words, the displacements δR and δr are admissible to the kinematic constraints. Call δF and δf a set of any arbitrary “virtual” forces in equilibrium.
  25. [25]
    [PDF] Chapter 11: Virtual Work
    Principle of Virtual Work. The principle of virtual work states that if a body is in equilibrium, then the algebraic sum of the virtual work done by all the ...
  26. [26]
    [PDF] Chapter 1 Principle of virtual work
    Mar 1, 2014 · 1.3 Principle of virtual work. The modern approach to a statics problem is to apply the two conditions that the total force and the total ...
  27. [27]
  28. [28]
    [PDF] Analytical Dynamics of Discrete Systems
    mk üik − fik = 0, i = 1, ··· , 3 ! If the virtual work equation is satisfied for any displacement ... Principle of Virtual Work for a System of N Particles.
  29. [29]
    Governing eqs - 2.4 Work - Applied Mechanics of Solids
    In this section, we derive formulas that enable you to calculate the work done by stresses acting on a solid. In addition, we prove the principle of virtual ...
  30. [30]
    [PDF] arXiv:physics/0510204v2 [physics.ed-ph] 2 Jun 2006
    The concept of virtual displacement and the principle of zero virtual work by constraint forces are central to both Lagrange's method of undetermined ...
  31. [31]
    Video: Virtual Work - JoVE
    Sep 22, 2023 · The principle of virtual work states that if a body is in static and dynamic equilibrium, then the sum of all the virtual work done by all ...
  32. [32]
    [PDF] Learning The Virtual Work Method In Statics - ASEE PEER
    ̇ Principle of virtual work​​ Recall that bodies considered here are rigid bodies or systems of pin-connected rigid bodies. The term “force system” denotes a ...Missing: expression seminal
  33. [33]
    [PDF] Ch10: Method of Virtual Work - Amin Fakhari
    The method of virtual work is particularly effective for solving equilibrium problems that involve an ideal system of several connected rigid bodies. The ...
  34. [34]
    [PDF] Physically Based Modeling: Principles and Practice Constrained ...
    Figure 2: In the case of a point-on-circle constraint, the principle of virtual work simply requires the constraint force to lie in a direction normal to the ...
  35. [35]
    [PDF] On the foundations of analytical dynamics
    referred to as ideal constraints, and the assumption that they do no work under virtual displacements is referred to as D'Alembert's principle [8]. However ...Missing: orthogonality | Show results with:orthogonality
  36. [36]
    [PDF] 2006-823: LEARNING THE VIRTUAL WORK METHOD IN STATICS
    Statics is a course aimed at developing in students the concepts and skills related to the analysis and prediction of conditions of bodies under the action of ...Missing: expression seminal
  37. [37]
    [PDF] NONIDEAL CONSTRAINTS AND LAGRANGIAN DYNAMICS
    Jan 1, 2000 · the forces of constraint do no work under virtual displace- ments. Such constraints are often referred to as ideal con- straints and seem to ...Missing: dissipative | Show results with:dissipative
  38. [38]
    [PDF] The principle of virtual work, counterfactuals, and the avoidance of ...
    The first principles of mechanics, with historical and practical illustrations. Cambridge: L. and L.L. Deighton. Whewell, W. (1874). History of the ...
  39. [39]
    Archimedes' Law of the Lever
    Archimedes' Law of the Lever states that magnitudes are in equilibrium at distances reciprocally proportional to their weights.Missing: derivation | Show results with:derivation
  40. [40]
    The Feynman Lectures on Physics Vol. I Ch. 4: Conservation of Energy
    This approach is called the principle of virtual work, because in order to apply this argument we had to imagine that the structure moves a little—even though ...
  41. [41]
    [PDF] Method of Virtual Work | Seismic Consolidation
    The principle of virtual work for a particle states that if a parti- cle is in equilibrium, the total virtual work of the forces acting on the particle is zero ...
  42. [42]
    [PDF] Chapter 1 D'Alembert's principle and applications
    Feb 1, 2014 · The first term in equation (1.2.22) is the virtual inertial work that would occur if all of the mass were concentrated at the center of mass.
  43. [43]
    [PDF] D'Alembert's Principle - Craig Fraser - University of Toronto
    Introduction. In 1743 the young French geometer Jean d'Alembert published his work Treatise on Dynamics, in which the Laws of Equilibrium and.
  44. [44]
    [PDF] Multibody dynamics
    Generalized coordinates. • Virtual work and generalized forces. • Lagrangian dynamics for mass points. • Lagrangian dynamics for a rigid body.
  45. [45]
    None
    ### Summary of Virtual Work and Inertia Torques for Slider-Crank Mechanism
  46. [46]
    [PDF] Work and energy in inertial and non inertial reference frames - arXiv
    Nov 30, 2008 · does not produce a virtual work in the sense of D'Alembert ... such a theorem behaves in a non inertial frame. It is worth pointing ...
  47. [47]
    Solid Mechanics - 2.5 principle of virtual work
    We obtain a principle stating the equality of these two works known as the Principle of virtual work. This result is obtained by reformulating the differential ...
  48. [48]
    [PDF] MEEN 618: ENERGY AND VARIATIONAL METHODS
    Mar 8, 2017 · The principle of virtual displacements can also be used, in addition to deriving equations of equilibrium, to directly determine reaction ...
  49. [49]
    [PDF] 6 the principle of virtual displacements (pvd)
    The principle of virtual displacements is one of the most pow- erful ... (virtualв work done by the external loads (body force and trac- tionв is equal ...
  50. [50]
    [PDF] The Finite Element Method for One-Dimensional Problems
    The weak form which is the principle of virtual displacements. The weak form can be derived from the strong form or found from the principle of minimum ...
  51. [51]
    [PDF] The principle of virtual forces - Ruhr-Universität Bochum
    If you apply infinitisemal small, virtual forces (stresses) on a field, the external virtual work is equal to the whole inner virtual work. The principle of ...
  52. [52]
    [PDF] VARIATIONAL PRINCIPLES IN THE LINEAR THEORY OF ... - DTIC
    The object of this paper is to supply generalizations to linear quasi-static viscoelasticity theory of certain variational principles which characterize the ...
  53. [53]
    Limit Analysis Method - an overview | ScienceDirect Topics
    The limit analysis constitutes now a classical part of the theory of plasticity with a broad spectrum of applications in structural mechanics, geotechnical ...<|control11|><|separator|>
  54. [54]
    [PDF] Introduction to Finite Element Analysis
    ... Principle of virtual work ... [div σ]i = σij,j. (1.38). Inserting (1.37) into (1.35) and bringing all volume integrals onto the same side of the equation.
  55. [55]
    [PDF] VARIATIONAL FORMULATIONS - Docenti.unina.it
    into an integral, or “weak”, form represented by the specialization of the ... The Principle of Virtual Displacements represents the theoretical basis of the.
  56. [56]
    [PDF] Energy Theorems and Structural Analysis - Computational Mechanics
    The second fundamental principle is developed in Section 6. We call it the principle of virtual forces or complementary virtual work. Here we consider a state ...
  57. [57]
    [PDF] arXiv:2410.02960v2 [math.SG] 3 Feb 2025
    Feb 3, 2025 · ... virtual work principle ... ([0,T],Q). Proposition 1.1. Hamilton's variational principle holds if and only if the Euler–Lagrange equation.Missing: ∫ | Show results with:∫
  58. [58]
    [PDF] From virtual work principle to least action principle for stochastic ...
    Although MEP, as a variational method, is actually almost a doctrine for many and used often for equilibrium as well as for nonequilibrium system, the ...
  59. [59]
    Nonholonomic Constraints and Gauss's Principle of Least Constraint
    Jan 1, 1972 · A modified form of Gauss's principle of least constraint is used to derive the equations of motion for systems with nonlinear, nonholonomic ...Missing: virtual | Show results with:virtual
  60. [60]
    Dynamics of Multibody Systems Using Virtual Work and Symbolic ...
    Virtual work and virtual power methods are used to develop the dynamic equations in terms of joint coordinates. These dynamic equations are reduced to a minimal ...Missing: discrete | Show results with:discrete
  61. [61]
    Dynamics of Flexible Multibody Systems Using Virtual Work and ...
    Aug 9, 2025 · By combining linear graph theory with the principle of virtual work, we obtain a dynamic formulation that extends graph-theoretic modelling ...
  62. [62]
    When is the principle of virtual work valid? - Physics Stack Exchange
    Feb 19, 2012 · The principle of virtual work says that forces of constraint don't do net work under virtual displacements that are consistent with constraints.Missing: dissipative limitations
  63. [63]
    [PDF] The extrema of an action principle for dissipative mechanical systems
    Oct 1, 2013 · This Lagrangian could also be de- rived from the virtual work principle[16]. In this formulation of LAP for dissipative systems, the three major ...
  64. [64]
    On the Nature of Constraints for Continua Undergoing Dissipative ...
    Jul 28, 2005 · We now restrict ourselves to systems for which the net virtual work of forces of constraint is zero. We have seen that this condition holds ...