Fact-checked by Grok 2 weeks ago

Generalized coordinates

In , generalized coordinates are a minimal set of independent parameters, such as lengths, angles, or other variables, that completely and unambiguously specify the configuration of a mechanical system at any given time, allowing for a more efficient description than the full set of Cartesian coordinates, particularly in systems with constraints. This concept was introduced by in his seminal 1788 work Mécanique Analytique, where he reformulated in a purely analytical , emphasizing principles over Newtonian forces to derive . The number of generalized coordinates equals the system's , which is typically 3N for N unconstrained particles in but reduced by the number of . In , these coordinates q_i and their time derivatives \dot{q}_i are used to express the T and V, forming the L = T - V, from which Lagrange's equations \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = Q_i (where Q_i are generalized forces) yield the system's dynamics without explicitly resolving constraint forces. This approach offers significant advantages, including coordinate invariance, simplified handling of complex geometries like pendulums or multi-body systems, and applicability to non-inertial frames or continuous systems in field theory.

Fundamentals

Definition and motivation

In classical mechanics, generalized coordinates refer to a minimal set of independent parameters q_1, q_2, \dots, q_n that fully describe the configuration of a possessing n , eliminating redundancy in the representation of the system's state. These coordinates serve as a flexible framework, distinct from the fixed Cartesian , by parameterizing positions in a way that aligns directly with the system's intrinsic and constraints. The concept originated in the late through the pioneering work of , who introduced it in his seminal 1788 treatise Mécanique Analytique to advance the field of . Lagrange's formulation shifted the focus from force-based Newtonian methods to a , using these coordinates to derive systematically without relying on geometric intuitions. The primary motivation for generalized coordinates lies in their ability to streamline the analysis of constrained mechanical systems by reducing the number of variables from the 3N Cartesian coordinates (where N is the number of particles) to just n, incorporating constraints implicitly and avoiding the need to solve extraneous equations. This simplification is particularly advantageous in Lagrangian and Hamiltonian mechanics, where it facilitates the derivation of equations of motion by excluding non-contributory forces, such as those perpendicular to the motion. Furthermore, they promote computational efficiency for intricate systems, such as those involving rigid bodies, and enable the adoption of curvilinear systems like polar or spherical coordinates, which better capture rotational or symmetric dynamics.

Relation to Cartesian coordinates

In classical mechanics, the positions of the particles in a system of N particles are typically described in Cartesian coordinates as \mathbf{r}_i = (x_i, y_i, z_i) for i = 1, \dots, N. To employ generalized coordinates q_1, q_2, \dots, q_s, where s \leq 3N is the number of , each Cartesian component is expressed as a function of these generalized coordinates and possibly time: x_i = x_i(q_1, \dots, q_s, t), y_i = y_i(q_1, \dots, q_s, t), z_i = z_i(q_1, \dots, q_s, t). This transformation maps the full 3N-dimensional Cartesian space to a reduced s-dimensional space, incorporating any constraints through the choice of the q_k. The nature of this transformation depends on whether the underlying constraints are time-independent or time-dependent. In scleronomic systems, the constraints do not explicitly involve time, so the position functions simplify to x_i = x_i(q_1, \dots, q_s), without the t argument. In contrast, rheonomic systems have time-dependent constraints, leading to explicit time reliance in the : x_i = x_i(q_1, \dots, q_s, t). This distinction affects the form of the dynamical equations, as time dependence in rheonomic cases introduces additional terms. The set of generalized coordinates q_1, \dots, q_s parameterizes the configuration space, which is an s-dimensional manifold representing all possible configurations of the system accessible under the given constraints. Each point in this space corresponds to a unique specification of the system's position, with the manifold's geometry determined by the transformation to ; for , the configuration space is time-independent, while for , it evolves with time. The velocities in Cartesian coordinates are obtained by differentiating the position functions with respect to time, yielding the transformation \mathbf{v}_i = \dot{\mathbf{r}}_i = \sum_{k=1}^s \frac{\partial \mathbf{r}_i}{\partial q_k} \dot{q}_k + \frac{\partial \mathbf{r}_i}{\partial t}, where the partial derivatives form the columns of the matrix \mathbf{J} with elements J_{\alpha k} = \partial x_{\alpha i}/\partial q_k for the \alpha-th Cartesian component. This relates the generalized velocities \dot{q}_k to the Cartesian velocities \mathbf{v}_i, and its nonzero ensures the transformation is locally invertible, preserving the structure of the configuration space.

Constraints and Degrees of Freedom

Holonomic constraints

Holonomic constraints are those that can be expressed in the form f_j(q_1, \dots, q_n, t) = 0 for j = 1, \dots, m, where the q_i are generalized coordinates and t is time, allowing the system's to be described on a lower-dimensional manifold parameterized by the remaining independent coordinates./13%3A_Lagrangian_Mechanics/13.03%3A_Holonomic_Constraints) These constraints arise from integrable relations, such as geometrical conditions like a fixed between particles in a , which restrict motion without depending on velocities. For a system of N particles initially with $3N Cartesian , m independent reduce the number of generalized coordinates to n = 3N - m. The integrability of stems from their representation as forms that are , enabling the elimination of dependent variables to yield explicit of coordinates and time. Specifically, a in the form \sum_i a_i \, dq_i + a_t \, dt = 0 is if there exists an making it the total of some f(q, t), thus df = 0 implies f = \text{constant}. Constraints without explicit time dependence, known as scleronomic, simplify to f_j(q_1, \dots, q_n) = 0, further restricting the system to a fixed configuration space. In the Lagrangian formulation, holonomic constraints can be enforced either by directly substituting the relations to reduce variables or, more generally, by introducing Lagrange multipliers \lambda_j to augment the as \mathcal{L}' = \mathcal{L} - \sum_j \lambda_j f_j, preserving all coordinates while accounting for the constraints through the . This approach ensures the generalized coordinates effectively parameterize the admissible configurations on the constraint manifold.

Non-holonomic constraints

Non-holonomic constraints are velocity-dependent relations that cannot be expressed as functions of the generalized coordinates q_i and time t alone, but instead take the form of linear differential equations known as constraints: \sum_{k=1}^n a_{jk}(q, t) \, dq_k + a_{j0}(q, t) \, dt = 0, where j = 1, \dots, m labels the s, and the coefficients a_{jk} and a_{j0} depend on the coordinates and time. These are not exact differentials, meaning they cannot be integrated to yield a solely on the variables, distinguishing them from that reduce the system's space. The non-integrability of such constraints is determined by the Frobenius theorem, which provides necessary and sufficient conditions for a of fields (or equivalently, a system) to be integrable. Specifically, for the constraint forms to be integrable, the exterior derivatives of the Pfaffian equations must lie within the ideal generated by the constraints themselves; if this involutivity condition fails, the constraints are non-holonomic, restricting allowable fields without confining the accessible configuration space. A classic example is the rolling without slipping of a or disk on a , where the imposes \dot{x} = r \dot{\theta} \cos \phi and \dot{y} = r \dot{\theta} \sin \phi, with x, y as the contact point coordinates, \theta as the rotation angle, \phi as the heading angle, and r the ; these link velocities but do not integrate to a position , allowing the disk to reach any (x, y) while following curved paths. Another instance is a rolling on a , where similar velocity relations prevent sliding but permit full positional freedom in the . In systems with N particles, non-holonomic constraints do not reduce the dimension of the configuration space, which remains $3N (or n generalized coordinates), but they limit the admissible trajectories in that space, constraining the dynamics to a sub-bundle of the . The number of for configuration is thus n = 3N - 0 (no reduction), but the effective dynamical freedom is n - m, where m is the number of independent constraints. Addressing non- systems poses challenges because the constraints cannot be used to eliminate coordinates directly, as in the case; instead, formulations like quasi-coordinates—parameters proportional to velocities rather than positions—or Appell's equations, which incorporate the constraints into higher-order terms involving accelerations, are required to derive the . These methods, originally developed by Appell in 1899, allow for the systematic inclusion of non-integrable constraints without reducing the coordinate set.

Physical Quantities

Kinetic energy

In classical mechanics, the kinetic energy T of a system consisting of N particles, each with mass m_i and position vector \mathbf{r}_i, is fundamentally expressed in Cartesian coordinates as T = \frac{1}{2} \sum_{i=1}^N m_i \left\| \frac{d\mathbf{r}_i}{dt} \right\|^2, where \left\| \frac{d\mathbf{r}_i}{dt} \right\|^2 = \dot{\mathbf{r}}_i \cdot \dot{\mathbf{r}}_i represents the squared speed of the i-th particle. To express this in generalized coordinates q_1, \dots, q_n, the position vectors are written as \mathbf{r}_i = \mathbf{r}_i(q_1, \dots, q_n, t), reflecting the system's configuration space parametrized by the q_k. The of each particle then follows from the : \dot{\mathbf{r}}_i = \sum_{k=1}^n \frac{\partial \mathbf{r}_i}{\partial q_k} \dot{q}_k + \frac{\partial \mathbf{r}_i}{\partial t}. For scleronomic systems, where the between generalized and Cartesian coordinates is time-independent (i.e., \frac{\partial \mathbf{r}_i}{\partial t} = 0), the simplifies to \dot{\mathbf{r}}_i = \sum_{k=1}^n \frac{\partial \mathbf{r}_i}{\partial q_k} \dot{q}_k. Substituting this into the expression and expanding the yields \left\| \dot{\mathbf{r}}_i \right\|^2 = \sum_{k=1}^n \sum_{l=1}^n \dot{q}_k \dot{q}_l \left( \frac{\partial \mathbf{r}_i}{\partial q_k} \cdot \frac{\partial \mathbf{r}_i}{\partial q_l} \right). Thus, the total becomes T = \frac{1}{2} \sum_{k=1}^n \sum_{l=1}^n g_{kl}(q, t) \dot{q}_k \dot{q}_l, where the coefficients g_{kl} form the elements of the (or ) on the configuration space, defined by g_{kl} = \sum_{i=1}^N m_i \left( \frac{\partial \mathbf{r}_i}{\partial q_k} \cdot \frac{\partial \mathbf{r}_i}{\partial q_l} \right). This g_{kl} is symmetric (g_{kl} = g_{lk}) and, for physically realizable systems, positive definite, ensuring T > 0 for any nonzero \dot{q}. The resulting form of T is a homogeneous in the generalized velocities \dot{q}_k, which endows the configuration space with a Riemannian metric structure induced by the system's ; this geometric underscores how g_{kl} quantifies the "inertial " between different directions in coordinate space. In rheonomic systems (time-dependent transformations), cross terms involving \dot{q}_k and \frac{\partial \mathbf{r}_i}{\partial t} may appear, but the quadratic structure in \dot{q} persists as the dominant feature. For systems subject to constraints, the expression for T adapts to the choice of coordinates. In the presence of , which can be integrated to reduce the number of independent coordinates to the n, the generalized coordinates q_k are selected to automatically satisfy the constraints, allowing T to be directly formulated in the reduced set as above, with the g_{kl} reflecting the constrained configuration manifold. For non-holonomic constraints, which impose velocity-dependent restrictions without integrable position constraints, a larger set of coordinates is typically used; the kinetic energy is expressed in these coordinates, but the allowable velocities \dot{q} are projected onto the orthogonal to the constraint gradients (e.g., via the constraint ), ensuring compatibility while preserving the quadratic form of T.

Generalized momentum

In Lagrangian mechanics, the generalized momentum p_k conjugate to the generalized coordinate q_k is defined as the partial derivative of the L with respect to the corresponding generalized velocity \dot{q}_k, given by p_k = \frac{\partial L}{\partial \dot{q}_k}, where L = T - V is the , with T denoting the and V the of the system. Since the potential energy V generally does not depend on the velocities, this simplifies to p_k = \frac{\partial T}{\partial \dot{q}_k}. For systems where the takes a quadratic form in the generalized velocities, as expressed through the g_{kl} from the kinetic energy section, the generalized momentum assumes the linear form p_k = \sum_l g_{kl} \dot{q}_l. A significant property of the generalized momentum arises when the is independent of the coordinate q_k, rendering q_k a cyclic or ignorable coordinate. In such cases, the Euler-Lagrange equation yields \frac{d p_k}{dt} = -\frac{\partial L}{\partial q_k} = 0, implying that p_k is conserved throughout the motion. This conservation corresponds to symmetries in the system; for instance, in polar coordinates describing motion under a central , the angular coordinate \theta is cyclic, and its conjugate p_\theta = m r^2 \dot{\theta} represents the , which remains constant. Such conserved generalized momenta provide integrals of motion that simplify the analysis of dynamical systems. The generalized momentum relates to the linear momenta in Cartesian coordinates through the coordinate transformation. Specifically, if the position vector in Cartesian coordinates is expressed as a function of the generalized coordinates, the conjugate momentum is p_k = \sum_i \frac{\partial x_i}{\partial q_k} p_{x_i} + \sum_i \frac{\partial y_i}{\partial q_k} p_{y_i} + \sum_i \frac{\partial z_i}{\partial q_k} p_{z_i}, where p_{x_i}, p_{y_i}, p_{z_i} are the components of the linear momenta. This expression demonstrates that each generalized momentum is a weighted sum of the Cartesian linear momentum components, with weights given by the partial derivatives of the Cartesian positions with respect to q_k; in the special case where the generalized coordinates are Cartesian, p_k directly coincides with the linear momentum. In the framework of , the generalized momenta serve as independent variables alongside the coordinates, facilitating a description of the dynamics. The function H is obtained from the via the Legendre transform, H = \sum_k p_k \dot{q}_k - L, typically resulting in H = T + V when expressed in terms of q_k and p_k. The time evolution of the system is then dictated by Hamilton's equations: \dot{q}_k = \frac{\partial H}{\partial p_k}, \quad \dot{p}_k = -\frac{\partial H}{\partial q_k}. This formulation highlights the symmetry between coordinates and momenta, enabling powerful techniques for solving mechanical problems and extending to .

Examples

Bead on a wire

A classic example of using generalized coordinates involves a of m constrained to slide without along a curved wire in . The wire's shape is arbitrary but fixed, and the bead's position can be parameterized by the s measured along the wire from a reference point, or, in the special case of a circular wire of R, by the angle \theta from a reference direction. This setup enforces a holonomic constraint, reducing the system's from three (in Cartesian coordinates) to one. The single generalized coordinate is s, the distance traveled along the wire, which fully specifies the bead's position given the wire's geometry. For a circular wire, s = R \theta, so \theta serves equivalently as the generalized coordinate. The kinetic energy T of the bead is then T = \frac{1}{2} m \dot{s}^2, reflecting the bead's speed solely along the wire's tangent, as the constraint eliminates motion perpendicular to it. The potential energy V due to gravity (assuming the wire lies in a vertical plane with height increasing along s) is V = m g h(s), where h(s) is the vertical height as a function of s. The for the system is formed as L = T - V = \frac{1}{2} m \dot{s}^2 - m g h(s). Applying the Euler- for the (with no non-conservative generalized forces), \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{s}} \right) - \frac{\partial L}{\partial s} = 0, yields the equation of motion. This simplifies to m \ddot{s} = -\frac{\partial V}{\partial s}, or equivalently m \ddot{s} = -m g \frac{d h}{d s}, where the constraint forces (normal to the wire) are handled implicitly without explicit computation. This approach demonstrates a key advantage of generalized coordinates: the three-dimensional problem is reduced to a single in one degree of freedom, avoiding the need for Lagrange multipliers to enforce the constraints explicitly.

Simple pendulum

The simple consists of a point m attached to a massless of fixed l, constrained to swing freely in a vertical under the influence of , with the pivot point fixed in space. This system has one degree of freedom due to the constraint enforcing the fixed rod length. The natural choice for the generalized coordinate is the angle \theta measured from the downward vertical. In terms of Cartesian coordinates with the pivot at the origin and the positive y-axis upward, the position of the mass is given by x = l \sin \theta, \quad y = -l \cos \theta. The corresponding velocity components are \dot{x} = l \dot{\theta} \cos \theta, \quad \dot{y} = l \dot{\theta} \sin \theta. These transformations express the constrained motion solely in terms of \theta and \dot{\theta}. The T is the radial form derived from the speed squared \dot{x}^2 + \dot{y}^2 = l^2 \dot{\theta}^2, yielding T = \frac{1}{2} m l^2 \dot{\theta}^2. The V, taking the reference at the lowest point, is V = -m g l \cos \theta, where g is the . The L = T - V then becomes L = \frac{1}{2} m l^2 \dot{\theta}^2 + m g l \cos \theta. Applying Lagrange's equation \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{\theta}} \right) - \frac{\partial L}{\partial \theta} = 0 to this system produces the nonlinear differential equation of motion \ddot{\theta} + \frac{g}{l} \sin \theta = 0. This equation captures the pendulum's oscillatory dynamics, reducible to for small \theta.

Double pendulum

The double pendulum consists of two point masses, m_1 and m_2, attached to massless rods of lengths l_1 and l_2, respectively, with the second rod connected to the end of the first, allowing motion in a under . This system has two , making it a classic example of coupled oscillators where the motion of one influences the other. The generalized coordinates are the \theta_1 and \theta_2, both measured from the downward vertical. The position of the first is given by Cartesian coordinates x_1 = l_1 \sin \theta_1, y_1 = -l_1 \cos \theta_1, assuming the pivot is at the and the positive y-axis points upward. For the second , the position is x_2 = x_1 + l_2 \sin \theta_2 = l_1 \sin \theta_1 + l_2 \sin \theta_2, y_2 = y_1 - l_2 \cos \theta_2 = -l_1 \cos \theta_1 - l_2 \cos \theta_2. The T of the system is derived from the velocities of both masses: T = \frac{1}{2} (m_1 + m_2) l_1^2 \dot{\theta}_1^2 + \frac{1}{2} m_2 l_2^2 \dot{\theta}_2^2 + m_2 l_1 l_2 \dot{\theta}_1 \dot{\theta}_2 \cos(\theta_1 - \theta_2). This expression captures the coupling term m_2 l_1 l_2 \dot{\theta}_1 \dot{\theta}_2 \cos(\theta_1 - \theta_2), which arises from the relative motion between the pendulums. The V is V = -(m_1 + m_2) g l_1 \cos \theta_1 - m_2 g l_2 \cos \theta_2, leading to the L = T - V. Applying the Euler-Lagrange equations to the yields two coupled nonlinear ordinary differential equations for \ddot{\theta}_1 and \ddot{\theta}_2: (m_1 + m_2) l_1 \ddot{\theta}_1 + m_2 l_2 \cos(\theta_1 - \theta_2) \ddot{\theta}_2 + m_2 l_2 \sin(\theta_1 - \theta_2) \dot{\theta}_2^2 + (m_1 + m_2) g \sin \theta_1 = 0, m_2 l_2 \ddot{\theta}_2 + m_2 l_1 \cos(\theta_1 - \theta_2) \ddot{\theta}_1 - m_2 l_1 \sin(\theta_1 - \theta_2) \dot{\theta}_1^2 + m_2 g \sin \theta_2 = 0. These equations highlight the nonlinear coupling, which can lead to chaotic behavior for certain initial conditions and energies.

Spherical pendulum

A spherical pendulum consists of a point mass m attached to a light, inextensible of fixed length l, suspended from a fixed point and allowed to swing freely in under the influence of , with the constraint that the mass moves on the surface of a sphere of radius l. This system has two degrees of freedom, making it an ideal example for illustrating generalized coordinates in Lagrangian mechanics. The motion is conveniently described using spherical coordinates, where the generalized coordinates are the polar angle \theta (measured from the downward vertical axis) and the azimuthal angle \phi (measuring around the vertical). The position of the mass in Cartesian coordinates is given by the transformations: \begin{align*} x &= l \sin \theta \cos \phi, \\ y &= l \sin \theta \sin \phi, \\ z &= -l \cos \theta, \end{align*} assuming the suspension point is at the origin and the z-axis points upward. The T of the is derived from its velocity components in spherical coordinates: T = \frac{1}{2} m l^2 \left( \dot{\theta}^2 + \sin^2 \theta \, \dot{\phi}^2 \right), while the V, due to , is V = - m g l \cos \theta, where g is the . The \mathcal{L} = T - V thus becomes \mathcal{L} = \frac{1}{2} m l^2 \left( \dot{\theta}^2 + \sin^2 \theta \, \dot{\phi}^2 \right) + m g l \cos \theta.[35] Applying the Euler-Lagrange equation to the \theta-coordinate yields the equation of motion: m l^2 \ddot{\theta} = m l^2 \sin \theta \cos \theta \, \dot{\phi}^2 - m g l \sin \theta, which simplifies to \ddot{\theta} = \sin \theta \cos \theta \, \dot{\phi}^2 - \frac{g}{l} \sin \theta.[35] The \phi-coordinate is cyclic in the , implying that the conjugate generalized p_\phi = \frac{\partial \mathcal{L}}{\partial \dot{\phi}} = m l^2 \sin^2 \theta \, \dot{\phi} is conserved, representing the constant about the vertical axis. This conservation allows reduction of the system to a single effective equation in \theta, highlighting the utility of generalized coordinates for capturing symmetries in constrained systems.

Virtual Work in Generalized Coordinates

Principle of virtual work

The principle of virtual work states that for a system in equilibrium, the total virtual work performed by all applied forces through any virtual displacement consistent with the system's constraints is zero. This is mathematically expressed as \delta W = \sum_i \mathbf{F}_i \cdot \delta \mathbf{r}_i = 0, where \mathbf{F}_i are the forces acting on each particle and \delta \mathbf{r}_i are the corresponding infinitesimal virtual displacements that satisfy the constraints without violating them. In the framework of generalized coordinates q_k, which parameterize the configuration space while incorporating constraints such as ones, the virtual work is reformulated as \delta W = \sum_k Q_k \delta q_k, where Q_k are the generalized forces conjugate to the coordinates q_k. The generalized forces are defined by Q_k = \sum_i \mathbf{F}_i \cdot \frac{\partial \mathbf{r}_i}{\partial q_k}, capturing the component of the applied forces that contributes to changes in the generalized coordinates. For systems subject to conservative forces derivable from a function V, the generalized forces take the form Q_k = -\frac{\partial V}{\partial q_k}, allowing the principle to connect directly with energy-based formulations. Extending this to dynamic systems, the principle links to , which incorporates inertial effects by requiring \sum_i (\mathbf{F}_i - m_i \mathbf{a}_i) \cdot \delta \mathbf{r}_i = 0, where m_i \mathbf{a}_i are the inertial forces, thus treating as a form of constrained . This approach was developed by in his seminal work Mécanique Analytique (1788), where he employed the principle of to derive in generalized coordinates, bypassing direct resolution of Newtonian forces and constraint reactions. When combined with the expression for in generalized coordinates, the principle yields the Lagrange equations of motion, providing a powerful tool for analyzing complex mechanical systems.

Application to generalized forces

In systems with generalized coordinates, the principle of virtual work is extended to define generalized forces Q_k that account for non-conservative forces acting on the particles of the system. Specifically, for a system of N particles with positions \mathbf{r}_i, the generalized force corresponding to the k-th coordinate q_k is given by Q_k = \sum_{i=1}^N \mathbf{F}_i^{\mathrm{nc}} \cdot \frac{\partial \mathbf{r}_i}{\partial q_k}, where \mathbf{F}_i^{\mathrm{nc}} are the non-conservative forces on the i-th particle, such as or external drives. This expression arises because the virtual work \delta W = \sum_{i=1}^N \mathbf{F}_i^{\mathrm{nc}} \cdot \delta \mathbf{r}_i can be rewritten using \delta \mathbf{r}_i = \sum_k \frac{\partial \mathbf{r}_i}{\partial q_k} \delta q_k, yielding \delta W = \sum_k Q_k \delta q_k. For ideal constraints, the reaction forces associated with perform no virtual work, as the virtual displacements \delta \mathbf{r}_i are chosen to lie in the admissible tangent to the constraint manifold, ensuring their contribution to Q_k vanishes. This formulation extends naturally to non-holonomic systems, where constraints are velocity-dependent and non-integrable, typically of the form \sum_j a_{jk}(q, t) \dot{q}_j = 0. In such cases, the virtual displacements \delta q_k must satisfy \sum_k a_{jk} \delta q_k = 0 for each , ensuring they are to the velocity directions in the . The generalized forces Q_k are then computed similarly, but now only over the non-conservative forces, with reactions again contributing zero under ideal (frictionless) conditions. This approach preserves the structure of the principle while accommodating systems like rolling without slipping or sliding with constraints. Representative examples illustrate the computation of Q_k. In rotational systems, such as a driven by an external \tau about the pivot, the generalized force for the angle coordinate \theta is Q_\theta = \tau, directly reflecting the work done through a virtual \delta \theta. For sliding systems involving , like a block on a rough surface with frictional \mathbf{f} = -\mu N \hat{v}, the generalized force for the position coordinate x becomes Q_x = \mathbf{f} \cdot \frac{\partial \mathbf{r}}{\partial x} = -\mu N if motion aligns with the displacement direction. These cases highlight how Q_k captures dissipative or applied effects not derivable from a potential. The generalized forces enter the via the Lagrange-d'Alembert form, which generalizes Lagrange's equations for systems with non-conservative forces or non-holonomic constraints: \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_k} \right) - \frac{\partial L}{\partial q_k} = Q_k, where L = T - V is the , with T the and V the potential for conservative forces. For non-holonomic cases, additional terms involving Lagrange multipliers \lambda_j may enforce the constraints, modifying the right-hand side to Q_k - \sum_j \lambda_j a_{jk}. This yields a complete dynamical description without explicitly resolving forces. However, the assumes constraints where reactions do no work, limiting its direct application to frictional constraints; in pure , it neglects inertial terms, reducing to conditions \sum_k Q_k \delta q_k = 0.

References

  1. [1]
    [PDF] Lagrange Handout - MIT OpenCourseWare
    Generalized coordinates: The generalized coordinates of a mechanical system are the minimum group of parameters which can completely and unambiguously define ...
  2. [2]
    [PDF] Generalized Coordinates, Lagrange's Equations, and Constraints
    1 Cartesian Coordinates and Generalized Coordinates. The set of coordinates used to describe the motion of a dynamic system is not unique.
  3. [3]
  4. [4]
    [PDF] 2. The Lagrangian Formalism - DAMTP
    Lagrange published his collected works on mechanics in 1788 in a book called “Mechanique. Analytique”. He considered the work to be pure mathematics and boasts ...<|control11|><|separator|>
  5. [5]
    Mécanique analytique : Lagrange, J. L. (Joseph Louis), 1736-1813
    Jan 18, 2010 · Mécanique analytique. by: Lagrange, J. L. (Joseph Louis), 1736 ... PDF download · download 1 file · SINGLE PAGE ORIGINAL JP2 TAR download.
  6. [6]
    [PDF] Lagrangian Dynamics: Generalized Coordinates and Forces Lecture ...
    “Lagrangian approach is simple but devoid of insight.” Both methods can be used to derive equations of motion.
  7. [7]
    Joseph-Louis Lagrange (1736 - 1813) - Biography - MacTutor
    In 1756 Lagrange sent Euler results that he had obtained on applying the calculus of variations to mechanics. These results generalised results which Euler had ...Missing: coordinates | Show results with:coordinates
  8. [8]
    [PDF] Chapter 2. Lagrangian Analytical Mechanics
    where qj are certain generalized coordinates that (together with constraints) completely define the system position. Their number J ≤ 3N is called the ...
  9. [9]
    [PDF] Classical Mechanics - UC Homepages
    Nov 30, 2023 · generalized coordinates, denoted by qi = q1,q2,... qs. (“Proper” in ... Relation to Cartesian coordinates: x1 = r cosϕ, x2 = r sinϕ, x3 ...
  10. [10]
    [PDF] Physics 5153 Classical Mechanics Generalized Coordinates and ...
    This lecture introduces some basic concepts of classical dynamics, and represents the beginning of our study of analytical mechanics. We will start with a brief ...
  11. [11]
    [PDF] Goldstein
    ... generalized coordinates in terms of conventional orthogonal po- sition coordinates. All sorts of quantities may be invoked to serve as generalized coordinates.
  12. [12]
    [PDF] ANALYTICAL MECHANICS
    ... constraints, in both holonomic. (or true) and nonholonomic (or quasi) ... Pfaffian, form f SًB vق ‏ B ¼ 0;. ً2:2:7ق where B ¼ Bًt; rق; B ¼ Bًt; rق are ...
  13. [13]
    [PDF] arXiv:2110.08337v3 [math-ph] 30 Jan 2022
    Jan 30, 2022 · If δξ is a pfaffian form that constitutes an exact differential, then δξ = dξ and the associated Pfaff equation dξ = 0 has as solution ξ = ξ ...
  14. [14]
    [PDF] 8.09(F14) Advanced Classical Mechanics - MIT OpenCourseWare
    running over the constraints. These holonomic or semi-holonomic constraints take the form gα(q, q˙,t) = ajα(q, t)q˙j + atα(q, t) = 0. (1.85) where again ...
  15. [15]
    [PDF] Notes on non-holonomic constraints - CLASSE (Cornell)
    Feb 23, 2013 · Non-holonomic constraints are defined by f(q1, qn, q˙1, qn˙, t)=0, where one must extremize a function f(x, y, ) such that the variation δr is ...
  16. [16]
    [PDF] The Theorem of Frobenius - Reed College
    One inquires whether or not the constraints are integrable, that is, holonomic. For the case just described, one would say that the con- straints are ...
  17. [17]
    [PDF] 5. Nonholonomic constraint - Mechanics of Manipulation
    Theorem 2.8 (Frobenius's theorem):. A regular distribution is integrable if and only if it is involutive. Proof: To prove that an integrable distribution is ...
  18. [18]
    [PDF] Mechanics of Nonholonomic Hybrid Lagrangian Systems
    This constraint is nonholonomic because it cannot be integrated to a con- straint on position. Intuitively, the ball is known to be rolling but that provides ...
  19. [19]
    [PDF] Review talk about nonholonomic dynamics
    Nonholonomic. The constraint is not integrable. Cannot be reduced to semi-holonomic constraints and does not impose restrictions on the configuration space.
  20. [20]
    [PDF] PHY411 Lecture notes on Constraints
    Jan 4, 2019 · Constraints in the form of equation 20 that can be written like equation 23 are holonomic, otherwise they are non-holonomic. Not all constraints ...
  21. [21]
    [PDF] Nonlinear nonholonomic constraints - arXiv
    Oct 21, 2019 · Although the most common technique in nonholonomic problems makes use of the so called quasi– coordinates and quasi–velocities (see, among ...<|control11|><|separator|>
  22. [22]
  23. [23]
    [PDF] inertial mechanics
    Kinetic energy in generalized coordinates is a quadratic form in velocity. The kernel of the quadratic form is the inertia tensor. I(θ) = J(θ)t MJ(θ). Ek ...
  24. [24]
    [PDF] 7 Hamilton's Principle - Lagrangian and Hamiltonian Dynamics
    This last equation suggests that we define the generalized momentum of a particle in the following way: pi = ∂T. ∂ ˙xi. (7.66). It is obviously consistent ...
  25. [25]
    [PDF] Energy Methods: Lagrange's Equations - MIT OpenCourseWare
    generalized coordinates qi are independent, ∂qi = 0 for i = j. ∂qj. Following the approach leading to Equations (2-10), we define a generalized momentum as.
  26. [26]
    Generalized Momenta - Richard Fitzpatrick
    Thus, we conclude that the generalized momentum associated with an ignorable coordinate is a constant of the motion. For example, in Section 9.5, the Lagrangian ...Missing: mechanics | Show results with:mechanics
  27. [27]
    [PDF] The Lagrangian Method
    we call it a conserved momentum. Note that a generalized momentum need not have the units of linear momentum, as the angular-momentum examples below show.
  28. [28]
    [PDF] CHM 532 Notes on Classical Mechanics Lagrange's and Hamilton's ...
    The angle φ is an example of a generalized coordinate. Generalized coordinates are any set of coordinates that are used to describe the motion of a physical ...
  29. [29]
    [PDF] Physics 5153 Classical Mechanics The Hamiltonian
    For these cases we use the Hamiltonian, which is a function of the coordinates and the momentum. ... Invert the generalized momentum, pi = ∂L. ∂ ˙qi , to obtain ˙ ...
  30. [30]
    [PDF] 1 - Chapter 7 Hamilton's Principle - Lagrangian and Hamiltonian ...
    The generalized coordinates are related to the Cartesian coordinates, and transformation rules allow use to carry out transformations between coordinate systems ...
  31. [31]
    [PDF] Bead moving along a thin, rigid, wire. - MIT Mathematics
    Oct 17, 2004 · An equation describing the motion of a bead along a rigid wire is derived. First the case with no friction is considered, and a Lagrangian ...
  32. [32]
    [PDF] AN INTRODUCTION TO LAGRANGIAN MECHANICS Alain J. Brizard
    Jul 7, 2007 · LAGRANGIAN MECHANICS. Figure 2.9: Generalized coordinates for the rotating-pendulum problem. The Lagrangian L = K − U is, therefore, written as.
  33. [33]
    [PDF] Lagrangian Mechanics - Physics Courses
    LAGRANGIAN MECHANICS. Figure 6.3: The double pendulum, with generalized coordinates θ1 and θ2. All motion is confined to a single plane. 6.5.5 The double ...
  34. [34]
    [PDF] Chaos in a double pendulum - James A. Yorke
    9. The initial position was chosen to have large pendula an- gles and so to exhibit chaotic motion.Missing: original | Show results with:original
  35. [35]
    Spherical Pendulum - Richard Fitzpatrick
    This type of pendulum is usually called a conical pendulum, since the string attached to the pendulum bob sweeps out a cone as the bob rotates.Missing: mechanics | Show results with:mechanics
  36. [36]
    [PDF] Generalized coordinates - Purdue Engineering
    Find: T – kinetic energy; identify terms of different type. XYZ – fixed frame xyz – moving frame – attached to the disk at. O' = P.
  37. [37]
    [PDF] Lecture #9 Virtual Work And the Derivation of Lagrange's Equations
    Principle of virtual work: ... → Work done for unit displacement of qi by forces acting on the system when all other generalized coordinates remain constant.
  38. [38]
    [PDF] J. L. Lagrange's early contributions to the principles and methods of ...
    The M~canique Analytique and the basic facts of LAGRANGE'S prior shift from the principle of least action to the principle of virtual velocities are reasonably.
  39. [39]
  40. [40]
    [PDF] Lagrangian Dynamics: Virtual Work and Generalized Forces
    Apr 4, 2007 · If we fix x and y, we can still rotate in a range with θ. # degrees of freedom = # of generalized coordinates: True for 2.003J. True for.
  41. [41]
    [PDF] Equations of motion for general nonholonomic systems from the d ...
    May 16, 2024 · The idea to extend the d'Alembert principle to general nonholonomic systems requires only the definition of the virtual displacements (i. e. ...
  42. [42]
    [PDF] Analytical Dynamics: Lagrange's Equation and its Application
    Use the generalized coordinates, q1 = r, and q2 = θ. 1The negative sign on δV is chosen to reflect that conservative forces may always be written as the ...