Virtual displacement is an infinitesimal, hypothetical change in the configuration of a mechanical system that adheres to its geometric and kinematic constraints without the passage of time, distinguishing it from actual motion where time evolves.[1] This concept, denoted as \delta \mathbf{r}_i for the displacement of the i-th particle, satisfies the constraint equations with dt = 0, ensuring that constraint forces perform no virtual work.[2] It forms the cornerstone of the principle of virtual work, which posits that for a system in static or dynamic equilibrium, the total virtual work done by applied and inertial forces through such displacements is zero, enabling the derivation of equations of motion without explicit constraint forces.[3]The idea of virtual displacement emerged in Hellenistic mechanics around the 4th century BCE, as seen in the Mechanica Problemata attributed to Pseudo-Aristotle, where geometric arguments on levers implicitly used virtual velocities proportional to displacements for equilibrium analysis.[4] Medieval scholars like Jordanus de Nemore in the 13th century refined these notions in De ratione ponderis, introducing positional gravity and virtual displacements to explain balance in weighted systems.[4] The modern formulation crystallized in the 18th century: Johann Bernoulli articulated the principle of virtual work for statics around 1717, Jean le Rond d'Alembert extended it to dynamics in 1743 via d'Alembert's principle, and Joseph-Louis Lagrange systematized it in Mécanique Analytique (1788), integrating it with generalized coordinates to yield Lagrange's equations.[5]In analytical mechanics, virtual displacements facilitate the transition from Newtonian formulations to variational approaches, allowing analysis of complex systems with holonomic or nonholonomic constraints by projecting forces onto allowable directions.[1] They underpin applications in structural engineering, where they equate external virtual work to internal strain energy for displacement calculations, and in robotics for constraint-compatible motion planning.[3] Extensions to relativistic and quantum contexts, such as in field theories, further generalize the principle, maintaining its role in minimizing action or energy functionals.[5]
Introduction and Definition
Historical Origins
The concept of virtual displacement traces its roots to medieval mechanics, where early ideas of infinitesimal variations emerged in discussions of static equilibrium and balance in weighted systems. These notions built upon ancient Hellenistic mechanics, as seen in the Mechanica Problemata attributed to Pseudo-Aristotle around the 4th century BCE, which employed geometric arguments on levers that implicitly utilized virtual velocities proportional to displacements for equilibrium analysis.[4]Medieval scholars like Jordanus de Nemore in the 13th century refined these ideas in De ratione ponderis, introducing the concept of positional gravity and virtual displacements to explain balance in inclined planes and weighted systems. Jordanus demonstrated that equilibrium occurs when the virtual work associated with possible displacements is zero, providing a foundational approach to constraint-compatible variations without time evolution.[4]The principle gained more structured form in the late 17th century through Pierre Varignon's work on machines and equilibrium. In his 1687 Projet d'une nouvelle mécanique, presented to the Paris Academy of Sciences, Varignon developed the principle of virtual velocities, applying it to systems of pulleys and levers to determine equilibrium conditions via infinitesimal motions.[6] He described these virtual motions as possible displacements that a system could undergo without violating constraints, though they remain unrealized in actual motion, emphasizing their role in balancing forces instantaneously.[7] This approach shifted focus from finite displacements to hypothetical infinitesimal ones, enabling analytical treatment of complex mechanisms.Johann Bernoulli advanced the idea significantly in 1717, formalizing virtual displacements within the framework of the principle of least action during his correspondence with Varignon.[8] Bernoulli articulated that for a system in equilibrium, the virtual work associated with infinitesimal displacements must vanish, integrating the concept into variational methods for optimizing paths of motion.[9] This formalization connected virtual displacements to broader dynamical principles, paving the way for their use beyond statics.In the 18th century, virtual displacement played a pivotal role in the transition from statics to full dynamics, with key contributions from Leonhard Euler and Joseph-Louis Lagrange. Euler, in works like his 1736 Mechanica and 1744 calculus of variations papers, extended virtual velocities to continuous systems, incorporating them into equations for elastic bodies and rigid motion.[10] Lagrange further synthesized these ideas in his 1760s memoirs and 1788 Mécanique Analytique, deriving the equations of motion from d'Alembert's principle augmented by virtual displacements, thus establishing an analytical foundation for mechanics that unified constraints and dynamics without geometric constructions.[8] This evolution marked virtual displacement as a cornerstone of modern analytical mechanics.
Formal Definition
In analytical mechanics, a virtual displacement is defined as an idealized, infinitesimal variation \delta \mathbf{q} of the generalized coordinates \mathbf{q} of a mechanicalsystem at a fixed instant of time t, such that it remains consistent with the instantaneous constraints of the system but does not necessarily adhere to the actual time evolution of the motion.[11] This variation represents a hypothetical "possible" reconfiguration that respects the system's positional restrictions at that moment, without implying any real physical movement or associated kinetics.[1]Unlike an actual displacement d\mathbf{q}, which constitutes the total differential incorporating both spatial changes and the time derivative \dot{\mathbf{q}} \, dt along the system's trajectory, a virtual displacement excludes any temporal progression by setting dt = 0.[1] Consequently, d\mathbf{q} = \delta \mathbf{q} + \dot{\mathbf{q}} \, dt, highlighting that virtual displacements capture only the allowable infinitesimal shifts in configuration space independent of velocity or acceleration.For holonomic systems, where constraints are expressible as functions of coordinates and time without velocities, virtual displacements reside in the tangent space to the configuration manifold at the current position, forming the space of all admissible infinitesimal directions tangent to the constraint surface.[12]Formally, the virtual displacement arises as the limiting difference \delta \mathbf{q} = \mathbf{q}'(t) - \mathbf{q}(t), where \mathbf{q}'(t) denotes a nearby allowable configuration at the same fixed time t, taken in the limit as the separation between \mathbf{q}' and \mathbf{q} approaches zero while preserving constraintcompatibility.[13]
Notation and Formulation
Standard Notations
In classical mechanics, virtual displacements are conventionally denoted using the symbol δ prefixed to the relevant coordinate or vector, indicating an infinitesimal, hypothetical variation consistent with system constraints at a fixed instant in time. For position vectors of particles, the standard notation is δ\mathbf{r}_i, where i indexes the particle, representing the virtual change in position.[11][2] In formulations using generalized coordinates, such as those in Lagrangian mechanics, the virtual displacement is denoted δq_j, where j labels the coordinate, capturing variations in abstract parameters that describe the system's configuration.[1][11] For rotational degrees of freedom, the angular virtual displacement is typically written as δθ, denoting a small hypothetical rotation.[14][11]A key distinction exists between the virtual variation δ and other differential symbols: δ signifies an infinitesimal virtual displacement that occurs without time evolution (δt = 0), whereas d denotes the actual differential along the true path of motion, incorporating time dependence (dr = \dot{\mathbf{r}} dt).[11][1] The symbol Δ, in contrast, represents finite displacements, which are non-infinitesimal changes not restricted to virtual contexts.[11] This notation for virtual work is expressed as δW = \sum_i \mathbf{F}_i \cdot δ\mathbf{r}_i, where \mathbf{F}_i are the applied forces, quantifying the hypothetical work under such displacements.[11][1][15]The usage of these notations varies with the coordinate system. In Cartesian coordinates, virtual displacements are decomposed into components δx, δy, δz along the orthogonal axes, facilitating direct vector analysis in Newtonian frameworks.[11][15] In curvilinear coordinates, such as polar or spherical systems, the notations adapt to the local basis, employing δq for generalized curvilinear parameters or δθ for angular components, which better suit constrained or non-Euclidean geometries.[11]
Aspect
Lagrangian Mechanics (Generalized Coordinates)
Newtonian Mechanics (Cartesian Vectors)
Primary Symbol
δq_j (virtual variation in generalized coordinate q_j)
δ\mathbf{r}_i (virtual displacement vector for particle i)
Context
Abstract coordinates for constrained systems; used in variational principles like Euler-Lagrange equations
Direct position vectors in Euclidean space; emphasis on force vectors and equilibrium
Intuitive for unconstrained particles; aligns with component-wise force resolution
These notations provide the symbolic foundation for expressing virtual displacements in the formal definition, enabling consistent application across mechanical analyses.[11][1]
Mathematical Representation
In classical mechanics, virtual displacements are mathematically represented as elements of the tangent space T_q Q to the configuration space Q at a configuration point q. The configuration space Q is the manifold parameterizing all possible positions of the mechanical system, often reduced by holonomic constraints to a lower-dimensional submanifold. A virtual displacement \delta q at q is a vector in T_q Q, capturing allowable infinitesimal variations in the generalized coordinates q = (q^1, \dots, q^n) that respect the system's constraints without advancing time.[16][17]For holonomic constraints defined by equations f_i(q) = [0](/page/0), i = 1, \dots, m, the virtual displacements form a vector field \delta \mathbf{r}(x) tangent to the constraint surface, satisfying the linear compatibility conditions \sum_k a_{ik} \delta q_k = [0](/page/0), where a_{ik} = \partial f_i / \partial q_k are the components of the constraintJacobian. These conditions ensure \delta \mathbf{r} lies in the kernel of the constraintgradient, orthogonal to the normal directions defined by \nabla f_i. An explicit representation arises via orthogonal projection onto the tangent plane: for a single constraint with normal vector \mathbf{n} = \nabla f, the projected virtual displacement is given by\delta \mathbf{r} = d\mathbf{r} - \frac{ (\mathbf{n} \cdot d\mathbf{r}) \mathbf{n} }{ |\mathbf{n}|^2 },where d\mathbf{r} is an arbitrary infinitesimal displacement in ambient space, and the subtracted term removes the component along \mathbf{n}. This projection spans the allowable directions while ensuring infinitesimal magnitude |\delta \mathbf{r}| \to [0](/page/0), though the directions collectively generate the full tangent space.In the case of nonholonomic constraints, which impose velocity-level restrictions such as \sum_k a_{ik} \dot{q}_k = 0 that are not integrable to position constraints, virtual displacements \delta q are defined to satisfy the same linear conditions \sum_k a_{ik} \delta q_k = 0, restricting them to a subbundle (distribution) D_q \subset T_q Q. Unlike holonomic cases, this distribution is non-integrable, meaning paths generated by integrating \delta \mathbf{r} may leave the allowable configuration submanifold, but individual virtual displacements remain instantaneously compatible with the constraints at fixed time. The infinitesimal nature persists, with |\delta \mathbf{r}| \to 0, emphasizing directional variations over finite motions.[16]
Core Properties
Linearity and Homogeneity
Virtual displacements exhibit key algebraic properties that align with the structure of vector spaces, specifically linearity and homogeneity. Linearity implies that if \delta_1 \mathbf{r} and \delta_2 \mathbf{r} are valid virtual displacements consistent with the system's constraints, then their linear combination \alpha \delta_1 \mathbf{r} + \beta \delta_2 \mathbf{r} is also a valid virtual displacement for any scalars \alpha, \beta \in \mathbb{R}. This property arises because virtual displacements are infinitesimal variations in the configuration space that respect the constraints, and the transformation between Cartesian and generalized coordinates preserves linearity: \delta \mathbf{r}_i = \sum_j \frac{\partial \mathbf{r}_i}{\partial q_j} \delta q_j. Homogeneity follows similarly, ensuring that scaling a virtual displacement by a constant preserves its validity: if \delta \mathbf{r} is a virtual displacement, then k \delta \mathbf{r} is also one for any scalar k \in \mathbb{R}. These properties hold for holonomic constraints, where the relations are linear in the differentials.[18]A sketch of the proof for these properties relies on the form of holonomic constraints, typically expressed as f_l(\mathbf{r}_1, \dots, \mathbf{r}_N, t) = 0 for l = 1, \dots, m, which yield linear relations in the virtual differentials upon differentiation: \sum_k a_{lk} \delta q_k = 0, where a_{lk} are coefficients depending on the configuration and time. For linear combinations, substituting \delta q_k = \alpha \delta_1 q_k + \beta \delta_2 q_k into the constraint equation gives \sum_k a_{lk} (\alpha \delta_1 q_k + \beta \delta_2 q_k) = \alpha \sum_k a_{lk} \delta_1 q_k + \beta \sum_k a_{lk} \delta_2 q_k = 0, since each individual virtual displacement satisfies the constraints. Homogeneity follows by setting \beta = 0 and \alpha = k, confirming that scaled variations remain admissible. This linearity extends to the operation on coordinates: \delta(\alpha q + \beta p) = \alpha \delta q + \beta \delta p.[18]The implications of linearity and homogeneity are profound, as they establish that the set of all virtual displacements at a given configuration forms a vector space, often denoted as the tangent space to the constraint manifold. This structure allows virtual displacements to span the space of allowable variations, facilitating the use of basis vectors (e.g., corresponding to independent generalized coordinates) to parameterize all possible infinitesimal motions consistent with constraints. Such a vector space framework underpins the derivation of equations of motion in analytical mechanics, enabling efficient handling of complex systems through linear algebra.[18]
Compatibility with Constraints
In classical mechanics, virtual displacements must be compatible with the kinematic constraints of the system to ensure that they represent allowable infinitesimal changes in configuration without violating the imposed restrictions. For holonomic constraints, expressed as f(\mathbf{q}) = 0, where \mathbf{q} denotes the generalized coordinates, the compatibility condition requires that the virtual displacement \delta \mathbf{q} satisfies \delta f = \sum_i \frac{\partial f}{\partial q_i} \delta q_i = 0. This condition implies that \delta \mathbf{q} lies in the tangent space orthogonal to the gradients of the constraint functions, preserving the manifold defined by the constraints.[19][2]For non-holonomic constraints, typically formulated in Pfaffian form as \sum_j a_{ij} \dot{q}_j = 0, the virtual displacements \delta \mathbf{q} must similarly obey \sum_j a_{ij} \delta q_j = 0, ensuring consistency with the differential constraints on allowable velocities at the instantaneous configuration. This formulation extends the compatibility to velocity-dependent restrictions, where virtual displacements are confined to the allowable directions in the configuration space without integrating the constraints fully.[20]Under the assumption of ideal constraints, the forces of constraint perform no virtual work, leading to the condition \sum_i \lambda_i \delta f_i = 0, where \lambda_i are the Lagrange multipliers associated with each constraint f_i = 0. This property allows constraint forces to be eliminated from the equations of motion, as their contribution vanishes for any compatible virtual displacement.[2][21]Unlike actual motions, which evolve under the influence of inertia and time derivatives, virtual displacements are purely kinematic, ignoring dynamic effects and focusing solely on positional consistency with constraints at a fixed instant. This distinction enables their use in static and dynamic principles without accounting for temporal evolution.[22]Holonomic constraints are classified as scleronomic if time-independent, where virtual displacements align directly with the fixed constraint geometry, or rheonomic if time-dependent, introducing explicit temporal variations that virtual displacements must instantaneously respect.[23][24]
In the simplest case of classical mechanics, consider an unconstrained particle moving freely in three-dimensional Euclidean space \mathbb{R}^3, with its position described by the vector \mathbf{r} = (x, y, z). A virtual displacement \delta \mathbf{r} for this particle is defined as any infinitesimalvector in \mathbb{R}^3, expressed as \delta \mathbf{r} = (\delta x, \delta y, \delta z), where each component \delta x, \delta y, and \delta z can vary independently without restriction.[25] This arbitrariness arises because there are no constraints on the motion, allowing the particle to "pretend" to displace in any direction instantaneously, without the passage of time or change in the applied forces.[25]The space of all possible virtual displacements at position \mathbf{r} corresponds to the full tangent space T_{\mathbf{r}} \mathbb{R}^3, which is isomorphic to \mathbb{R}^3 itself due to the flat geometry of Euclidean space.[26] This tangent space structure underscores that virtual displacements are tangent vectors at \mathbf{r}, capturing infinitesimal variations in the configuration manifold. The system has three degrees of freedom, with all directions allowable, reflecting the complete freedom of motion in \mathbb{R}^3.[25]Although virtual displacements do not involve time evolution, they can be conceptually related to the particle's velocity \mathbf{v} by considering \delta \mathbf{r} \approx h \mathbf{v} for a small scalar h > 0, but without implying actual flow along the trajectory.[25] In this setup, the virtual work performed by an external force \mathbf{F} on the particle is given by \delta W = \mathbf{F} \cdot \delta \mathbf{r}, where the dot product allows \delta W to take any value depending on the arbitrary direction of \delta \mathbf{r}. This linearity permits virtual displacements to be any linear combination of the standard basis vectors \mathbf{e}_x, \mathbf{e}_y, \mathbf{e}_z.[25]
Constrained Particles on a Surface
In classical mechanics, virtual displacements for a particle constrained to move on a surface defined by a holonomic constraint f(\mathbf{r}) = 0 must satisfy the condition \nabla f \cdot \delta \mathbf{r} = 0, ensuring that the infinitesimal displacement \delta \mathbf{r} remains tangent to the surface.[2] This orthogonality to the surface normal \mathbf{n} = \nabla f / |\nabla f| implies that \delta \mathbf{r} lies within the tangent plane at the particle's position, which is typically spanned by two linearly independent directions orthogonal to \mathbf{n}.[2] Such constraints reduce the degrees of freedom from three (in unconstrained Euclidean space) to two, as the particle's motion is confined to the two-dimensional manifold of the surface.[27]A representative example is a particle constrained to the surface of a sphere satisfying \mathbf{r} \cdot \mathbf{r} = R^2, where \mathbf{r} is the position vector and R is the radius. Here, \nabla f = 2\mathbf{r}, so the virtual displacement must obey \mathbf{r} \cdot \delta \mathbf{r} = 0, meaning \delta \mathbf{r} is perpendicular to \mathbf{r} and has infinitesimal magnitude |\delta \mathbf{r}| \ll R.[27] This condition confines \delta \mathbf{r} to the tangent plane at \mathbf{r}, preserving the spherical constraint during the virtual variation.To parametrize these displacements, surface coordinates such as parameters u and v can be used, expressing the position as \mathbf{r} = \mathbf{r}(u, v). The virtual displacement then takes the form\delta \mathbf{r} = \frac{\partial \mathbf{r}}{\partial u} \delta u + \frac{\partial \mathbf{r}}{\partial v} \delta v,where \delta u and \delta v are independent infinitesimal variations in the parameters, spanning the tangent space.[24]In the principle of virtual work, only external forces contribute to the virtual work \delta W = \mathbf{F} \cdot \delta \mathbf{r}, as the constraint force \mathbf{N} (normal to the surface) performs no work due to \mathbf{N} \cdot \delta \mathbf{r} = 0.[28] This property aligns with the compatibility of virtual displacements, which ensures their orthogonality to constraint normals.[1]
Rigid Body with Fixed Point
In the context of a rigid body pivoting around a fixed point O, virtual displacements capture the allowable infinitesimal changes in configuration that preserve the body's rigidity and the fixed-point constraint. For any point P in the body, the virtual displacement \delta \mathbf{r}_P is expressed as \delta \mathbf{r}_P = \delta \boldsymbol{\theta} \times (\mathbf{r}_P - \mathbf{O}), where \delta \boldsymbol{\theta} denotes the infinitesimal rotation vector representing the virtual angular variation. This formulation ensures that the displacement at O vanishes (\delta \mathbf{r}_O = 0) and that all points move perpendicular to the line connecting them to O, consistent with pure rotation. The vector \delta \boldsymbol{\theta} is analogous to the angular velocity \boldsymbol{\omega} scaled by an infinitesimal time interval \delta t, but it is formulated in a time-independent manner to focus on geometric variations rather than actual motion.[29][30][31]The rigidity of the body imposes holonomic constraints that fix the distances between all pairs of points, ensuring no deformation under virtual displacements. For points i and j, the constraint \delta (|\mathbf{r}_i - \mathbf{r}_j|^2) = 0 implies (\mathbf{r}_i - \mathbf{r}_j) \cdot (\delta \mathbf{r}_i - \delta \mathbf{r}_j) = 0. Substituting the rotational form of the displacements yields (\mathbf{r}_i - \mathbf{r}_j) \cdot [\delta \boldsymbol{\theta} \times (\mathbf{r}_i - \mathbf{r}_j)] = 0, which holds identically due to the properties of the cross product, confirming compatibility. These constraints reduce the configuration space, with the virtual displacements spanning only the rotational variations around O.[30][31]Orientation changes can be parameterized using Euler angles \phi, \theta, \psi, where the virtual displacements correspond to independent infinitesimal variations \delta \phi, \delta \theta, \delta \psi that respect the coordinate ranges (e.g., avoiding gimbal lock singularities). Alternatively, quaternions provide a non-singular representation, with virtual variations in the four quaternion components \delta q = (\delta q_0, \delta q_1, \delta q_2, \delta q_3) subject to the unit norm constraint q \cdot \delta q = 0. In both cases, these variations map to the same three-dimensional space of \delta \boldsymbol{\theta}. For a more restrictive scenario, such as rotation about a fixed axis, the virtual rotation \delta \theta is confined to the direction of the axis, reducing the allowable variations to a single scalar parameter.[31][32]The set of all possible virtual displacements for a rigid body with a fixed point forms the tangent space to the special orthogonal group SO(3) at the current orientation, which is three-dimensional and corresponds to the three independent rotational degrees of freedom. This structure allows linear combinations of basis rotations (e.g., about x, y, z axes) to generate arbitrary infinitesimal rotations, underscoring the vector space nature of the tangent space.[31][29]
Applications in Classical Mechanics
Principle of Virtual Work
The principle of virtual work states that a mechanical system is in equilibrium if and only if the total virtual work done by all applied forces through any admissible virtual displacement is zero.[1] This condition is expressed mathematically as \delta W = \sum_i \mathbf{F}_i \cdot \delta \mathbf{r}_i = 0, where \mathbf{F}_i are the applied forces acting on each particle and \delta \mathbf{r}_i are the corresponding virtual displacements that satisfy the system's constraints.[25] Admissible virtual displacements are infinitesimal and instantaneous changes in position that respect the geometric constraints without involving actual motion or time evolution.[1]For systems where the applied forces are conservative and derivable from a potential energy function V, the principle links directly to the stationarity of the potential energy. In such cases, the virtual work \delta W = -\delta V, so equilibrium requires \delta V = 0 for all admissible virtual displacements.[33] This equivalence underscores that equilibrium configurations minimize or stationarize the potential energy under the given constraints.[33]In constrained systems, the total virtual work includes contributions from both external applied forces and constraint forces, yet the latter perform no virtual work due to their nature. Constraint forces, such as those from supports or ties, act perpendicular to the allowable directions of motion, ensuring their dot product with admissible virtual displacements vanishes.[25] Formally, for holonomic constraints defined by f(\mathbf{r}_i) = 0, the constraint forces take the form \mathbf{F}_i^c = \lambda \nabla f_i, where \lambda is a scalar multiplier. The equilibrium equation becomes \sum_i (\mathbf{F}_i^\text{ext} + \lambda \nabla f_i) \cdot \delta \mathbf{r}_i = 0. Since admissible \delta \mathbf{r}_i satisfy \delta f_i = \nabla f_i \cdot \delta \mathbf{r}_i = 0, the terms involving \lambda vanish, reducing the condition to \sum_i \mathbf{F}_i^\text{ext} \cdot \delta \mathbf{r}_i = 0.[1]This principle finds practical application in analyzing the equilibrium of structures like trusses, where virtual displacements along member elongations help compute internal forces without resolving all reactions; in levers, to balance moments through infinitesimal rotations; and in simple machines, to determine force ratios under static loads.[34]
D'Alembert's Principle and Dynamics
D'Alembert's principle extends the concept of virtual displacements to dynamic systems by incorporating inertial effects, treating the system as if it were in a state of dynamic equilibrium. Formulated originally by Jean le Rond d'Alembert in 1743, the principle states that for a system of particles, the virtual work done by the applied forces minus the virtual work done by the inertial forces vanishes for any virtual displacement consistent with the constraints: \sum_i (\mathbf{F}_i - m_i \mathbf{a}_i) \cdot \delta \mathbf{r}_i = 0, where \mathbf{F}_i are the applied forces, m_i \mathbf{a}_i are the inertial terms, and \delta \mathbf{r}_i are the virtual displacements.[35] This approach effectively adds fictitious inertial forces -\mathbf{m}_i \mathbf{a}_i to the applied forces, allowing the use of static equilibrium methods for dynamic problems.[36]The principle leads directly to the equations of motion by separating the inertial and applied force contributions: \sum_i \mathbf{F}_i \cdot \delta \mathbf{r}_i = \sum_i m_i \mathbf{a}_i \cdot \delta \mathbf{r}_i. In terms of kinetic energy T, the right-hand side can be expressed using the relation \sum_i m_i \mathbf{a}_i \cdot \delta \mathbf{r}_i = \sum_j \left[ \frac{d}{dt} \left( \frac{\partial T}{\partial \dot{q}_j} \right) - \frac{\partial T}{\partial q_j} \right] \delta q_j, where q_j are generalized coordinates and \delta q_j are the corresponding virtual displacements. For conservative forces derivable from a potential V, this bridges to the Lagrangian formulation, yielding \sum_j \left[ \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_j} \right) - \frac{\partial L}{\partial q_j} \right] \delta q_j = 0 with L = T - V, implying the Euler-Lagrange equations for each j.[37][36][38]D'Alembert's principle adeptly handles both holonomic and nonholonomic constraints through the requirement that virtual displacements respect the constraints, ensuring that constraint forces perform no virtual work. For holonomic constraints, which can be expressed as functions of positions and time, the number of independent generalized coordinates is reduced accordingly, and Lagrange multipliers may be introduced if needed: \frac{d}{dt} \left( \frac{\partial T}{\partial \dot{q}_j} \right) - \frac{\partial T}{\partial q_j} + \sum_k \lambda_k \frac{\partial g_k}{\partial q_j} = Q_j. For nonholonomic constraints g_k(\mathbf{q}, \dot{\mathbf{q}}, t) = 0, virtual displacements are chosen to satisfy \sum_j \frac{\partial g_k}{\partial \dot{q}_j} \delta q_j = 0, ensuring zero virtual work by constraint forces. The resulting equations of motion are \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_j} \right) - \frac{\partial L}{\partial q_j} = \sum_k \lambda_k \frac{\partial g_k}{\partial \dot{q}_j} + Q_j, where \lambda_k are Lagrange multipliers.[37][38][39]The principle also extends naturally to rheonomic systems, where constraints or potentials depend explicitly on time, by allowing time-dependent transformations in the generalized coordinates and including \partial V / \partial t terms in the Lagrangian if applicable, while the virtual displacements remain instantaneous and time-independent. This maintains the principle's validity for time-varying constraints, such as those in mechanisms with moving supports.[37]