Fact-checked by Grok 2 weeks ago

Canonical transformation

A canonical transformation is a coordinate transformation in the of a classical that preserves the form of Hamilton's , mapping old coordinates q_i and momenta p_i to new coordinates Q_i and momenta P_i such that the new K(Q, P, t) generates equations \dot{Q}_i = \frac{\partial K}{\partial P_i} and \dot{P}_i = -\frac{\partial K}{\partial Q_i}. These transformations are fundamental to , as they allow for the simplification of complex Hamiltonians by rendering certain coordinates cyclic—meaning the Hamiltonian does not explicitly depend on them—thereby conserving the corresponding momenta and facilitating the solution of . For instance, in the case of a , a canonical transformation can decouple the equations into independent linear forms, revealing constants of motion directly. Canonical transformations are equivalently characterized by their preservation of the structure, ensuring \{Q_i, Q_j\} = \{P_i, P_j\} = 0 and \{Q_i, P_j\} = \delta_{ij}, which maintains the of . They can be generated systematically using type-specific functions, such as the identity-type generator F_1(q, Q, t) where p_i = \frac{\partial F_1}{\partial q_i} and P_i = -\frac{\partial F_1}{\partial Q_i}, with the new related by K = H + \frac{\partial F_1}{\partial t}. This property also ensures the transformation is volume-preserving in , with the determinant equal to 1, underscoring their role in conserving for incompressible flow in systems. Beyond , canonical transformations underpin quantization procedures, such as Dirac's , where brackets are replaced by commutators to bridge classical and quantum descriptions.

Fundamentals

Notation and phase space

In Hamiltonian mechanics, systems with n degrees of freedom are described using canonical coordinates, consisting of generalized position coordinates q_i (for i = 1, \dots, n) and conjugate momenta p_i, which are defined as p_i = \frac{\partial L}{\partial \dot{q}_i} from the Lagrangian L(q, \dot{q}, t). These coordinates provide a symmetric framework for formulating the dynamics. The is the $2n-dimensional manifold parameterized by the set \{q_1, \dots, q_n, p_1, \dots, p_n\}, equipped with a structure that encodes the geometry of the mechanical system. This structure is captured by the symplectic form \omega = \sum_{i=1}^n dq_i \wedge dp_i, which defines the fundamental relations between coordinates. The dynamics evolve on this manifold according to Hamilton's : \frac{dq_i}{dt} = \frac{\partial H}{\partial p_i}, \quad \frac{dp_i}{dt} = -\frac{\partial H}{\partial q_i}, where H(q, p, t) is the Hamiltonian function, typically expressed as H = \sum_i p_i \dot{q}_i - L. These first-order differential equations govern the time evolution of the system in phase space. A key algebraic tool in this formalism is the Poisson bracket, defined for two functions f and g on phase space as \{f, g\} = \sum_{i=1}^n \left( \frac{\partial f}{\partial q_i} \frac{\partial g}{\partial p_i} - \frac{\partial f}{\partial p_i} \frac{\partial g}{\partial q_i} \right). This bracket quantifies the commutator-like structure of observables and remains invariant under canonical transformations. Notably, canonical transformations preserve the volume of regions in phase space, ensuring that the $2n-dimensional measure is invariant, which follows from the determinant of the transformation Jacobian being \pm 1. This property, known as Liouville's theorem in the context of Hamiltonian flows, underscores the incompressible nature of phase space trajectories.

Definition and basic properties

In Hamiltonian mechanics, a canonical transformation is defined as a change of variables from the original canonical coordinates (q_i, p_i) and momenta to new coordinates (Q_i, P_i) such that the new variables satisfy Hamilton's equations of motion with respect to a transformed Hamiltonian K(Q, P, t), where K(Q, P, t) = H(q(Q, P, t), p(Q, P, t), t) and H is the original Hamiltonian. The specific form of the transformed Hamilton's equations is \dot{Q}_i = \frac{\partial K}{\partial P_i}, \quad \dot{P}_i = -\frac{\partial K}{\partial Q_i} for each index i = 1, \dots, n, mirroring the structure of the original equations \dot{q}_i = \partial H / \partial p_i, \dot{p}_i = -\partial H / \partial q_i. This preservation of the form of Hamilton's equations arises from the underlying symplectic structure of , which canonical transformations maintain, ensuring the remain invariant in structure. Unlike general coordinate transformations, which alter only position variables and may disrupt the Hamiltonian framework, canonical transformations involve both coordinates and momenta simultaneously, thereby sustaining the canonical structure essential for Hamiltonian dynamics. Canonical transformations form a group under composition: the successive application of two such transformations yields another canonical transformation, with the transformation serving as the group element and each transformation having an inverse that is also canonical.

Conditions for canonical transformations

Symplectic condition

The symplectic condition provides the primary geometric criterion for a canonical transformation in classical mechanics, characterizing it as a transformation that preserves the symplectic structure of phase space. In a 2n-dimensional phase space with coordinates (q_1, \dots, q_n, p_1, \dots, p_n), the standard symplectic matrix is defined as J = \begin{pmatrix} 0 & I_n \\ -I_n & 0 \end{pmatrix}, where I_n is the n \times n identity matrix. This matrix encodes the fundamental symplectic form \omega = \sum_{i=1}^n dq_i \wedge dp_i, which defines the Poisson bracket structure and the geometry of Hamiltonian dynamics. For a differentiable (q, p) \mapsto (Q(q, p), P(q, p)), let M denote its matrix, with entries M_{ij} = \partial (Q_j, P_j)/\partial (q_k, p_k). The transformation is if and only if it satisfies the condition M^T J M = J, where M^T is the of M. This matrix equation ensures that the of the form under the transformation remains unchanged, i.e., \omega = dQ \wedge dP, preserving the area-like volumes in subspaces spanned by conjugate pairs. The derivation of this condition follows from requiring the transformed symplectic form to match the original. Consider the differential d\mathbf{z} = M d\mathbf{z}', where \mathbf{z} = (q, p) and \mathbf{z}' = (Q, P); the symplectic form transforms as \omega' = d\mathbf{z}^T J d\mathbf{z} = (d\mathbf{z}')^T M^T J M d\mathbf{z}', which equals \omega if M^T J M = J. A consequence is that \det M = 1, implying the transformation is volume-preserving in the full , consistent with for Hamiltonian flows. This condition guarantees the invariance of Hamilton's equations under the transformation. To see this, suppose the original system obeys \dot{\mathbf{z}} = J \frac{\partial H}{\partial \mathbf{z}}; in the new coordinates, with K(\mathbf{z}') = H(\mathbf{z}(\mathbf{z}')), \dot{\mathbf{z}}' = M \dot{\mathbf{z}} = M J \frac{\partial H}{\partial \mathbf{z}}. By the chain rule, \frac{\partial H}{\partial \mathbf{z}} = M^T \frac{\partial K}{\partial \mathbf{z}'} (assuming time-independent transformation for simplicity). Thus, \dot{\mathbf{z}}' = M J M^T \frac{\partial K}{\partial \mathbf{z}'}. The symplectic condition M^T J M = J implies M J M^T = J, so \dot{\mathbf{z}}' = J \frac{\partial K}{\partial \mathbf{z}'}, preserving the form of the equations. For transformations, the condition linearizes to \delta M^T J + J \delta M = 0, but the finite case applies directly to general maps.

Poisson bracket invariance

In Hamiltonian mechanics, the fundamental Poisson brackets in the original canonical coordinates q_i and momenta p_i are defined as \{q_i, p_j\} = \delta_{ij}, \{q_i, q_j\} = 0, and \{p_i, p_j\} = 0, where \delta_{ij} is the Kronecker delta. These relations encode the symplectic structure of phase space and ensure the consistency of Hamilton's equations. A key condition for a transformation from (q_i, p_i) to new coordinates (Q_k, P_l) to be canonical is that it preserves these fundamental Poisson brackets, meaning \{Q_k, P_l\} = \delta_{kl}, \{Q_k, Q_m\} = 0, and \{P_l, P_n\} = 0. This invariance extends to the Poisson bracket of any two functions on phase space, \{f, g\}_{Q,P} = \{f, g\}_{q,p}, ensuring that the algebraic structure of the theory remains unchanged under the transformation. To verify this condition explicitly, the Poisson bracket in the new variables is computed using the chain rule for the transformation, treating Q_k and P_l as functions of q_i and p_i: \{Q_k, P_l\}_{q,p} = \sum_i \left( \frac{\partial Q_k}{\partial q_i} \frac{\partial P_l}{\partial p_i} - \frac{\partial Q_k}{\partial p_i} \frac{\partial P_l}{\partial q_i} \right). For the transformation to be canonical, this must equal \delta_{kl}. The derivation follows from the general definition of the \{f, g\} = \sum_i \left( \frac{\partial f}{\partial q_i} \frac{\partial g}{\partial p_i} - \frac{\partial f}{\partial p_i} \frac{\partial g}{\partial q_i} \right), applied directly to Q_k and P_l, which shows that the bracket structure is preserved the relations hold in the new coordinates. This condition is mathematically equivalent to the condition on the of the transformation but offers a computationally convenient algebraic tool for explicit verification in practice, especially for low-dimensional systems. As a simple illustration in one dimension (n=1), consider the transformation Q = q and P = p + f(q), where f is an arbitrary of q. The is \{Q, P\} = \frac{\partial Q}{\partial q} \frac{\partial P}{\partial p} - \frac{\partial Q}{\partial p} \frac{\partial P}{\partial q} = (1)(1) - (0) f'(q) = 1, confirming canonicity.

Other bracket invariances

In addition to the Poisson bracket, the Lagrange bracket provides an alternative criterion for verifying transformations. The Lagrange bracket of two functions f and g is defined as [f, g] = \sum_{i=1}^n \left( \frac{\partial f}{\partial q_i} \frac{\partial g}{\partial p_i} - \frac{\partial f}{\partial p_i} \frac{\partial g}{\partial q_i} \right), which coincides with the in standard notation. For the fundamental coordinates, this yields [q_i, p_j] = \delta_{ij}, [q_i, q_j] = 0, and [p_i, p_j] = 0. A transformation to new coordinates Q_k, P_l is if it preserves these relations, meaning the Lagrange brackets computed with respect to the original coordinates satisfy [Q_k, P_l] = \delta_{kl}, [Q_k, Q_l] = 0, and [P_k, P_l] = 0. This preservation condition is mathematically equivalent to the invariance of the , as both stem from the underlying symplectic structure of . Historically, Lagrange brackets were introduced in the early as a in before the widespread adoption of Poisson brackets, offering a direct way to check canonicity without explicitly invoking evolution. They remain useful in specific coordinate systems, such as non-Cartesian generalizations or when emphasizing the algebraic properties of transformations over geometric ones. A more abstract invariance involves the bilinear form \omega = \sum_{i=1}^n dq_i \wedge dp_i, which defines the non-degenerate, skew-symmetric pairing on the . Canonical transformations preserve this form, satisfying \sum dQ_k \wedge dP_k = \sum dq_i \wedge dp_i. This bilinear preservation directly implies (and thus Lagrange) bracket invariance for arbitrary functions f and g, since the brackets measure contractions with respect to \omega. Indirect verification via differentials arises in this context: by ensuring the form remains unchanged under the transformation, one confirms canonicity without computing brackets explicitly for all pairs. Although equivalent to the primary Poisson condition, these bracket and bilinear invariances highlight niche applications, such as in extended spaces or when deriving transformation properties from .

Generating functions

Types 1 and 2

Generating functions provide a systematic to construct canonical transformations by specifying relations between old and new variables through partial derivatives. The approach originates from ensuring the invariance of Hamilton's under the transformation, leading to the total dF = \sum_i p_i \, dq_i - \sum_i P_i \, dQ_i + (K - H) \, dt, where F is the generating function, H(q, p, t) is the old , and K(Q, P, t) is the new one. For Type 1 generating functions, F = F_1(q, Q, t) depends on the old coordinates q, new coordinates Q, and time t. The total differential is dF_1 = \sum_i \frac{\partial F_1}{\partial q_i} \, dq_i + \sum_i \frac{\partial F_1}{\partial Q_i} \, dQ_i + \frac{\partial F_1}{\partial t} \, dt. Equating coefficients with the standard form yields the transformation rules p_i = \frac{\partial F_1}{\partial q_i}, \quad P_i = -\frac{\partial F_1}{\partial Q_i}, \quad K = H + \frac{\partial F_1}{\partial t}. These relations are obtained by identifying the coefficients of dq_i and dQ_i, with the time-dependent term adjusting the Hamiltonian to preserve the canonical structure. Type 1 functions are particularly useful for point transformations, where the new coordinates Q are expressed directly in terms of the old coordinates q, facilitating changes like rotations or scalings in configuration space. For Type 2 generating functions, F = F_2(q, P, t) depends on the old coordinates q, new momenta P, and time t. To align with the standard differential, the form is adjusted such that dF_2 + \sum_i Q_i \, dP_i = \sum_i p_i \, dq_i - \sum_i P_i \, dQ_i + (K - H) \, dt, but the direct relations are derived as p_i = \frac{\partial F_2}{\partial q_i}, \quad Q_i = \frac{\partial F_2}{\partial P_i}, \quad K = H + \frac{\partial F_2}{\partial t}. Here, the identification follows from expanding dF_2 = \sum_i \frac{\partial F_2}{\partial q_i} \, dq_i + \sum_i \frac{\partial F_2}{\partial P_i} \, dP_i + \frac{\partial F_2}{\partial t} \, dt and matching terms, noting that d(Q P) = Q \, dP + P \, dQ contributes to the mixed structure. Type 2 functions are advantageous when the new momenta P are specified in terms of the old variables, such as in transformations involving rescalings or when solving for new coordinates implicitly. A simple example of a Type 1 generating function is the identity transformation in one dimension, F_1(q, Q, t) = q Q, which gives p = Q and P = -q, with K = H if time-independent; this corresponds to a rotation that preserves canonicity. For multiple , the generalization F_1 = \sum_i q_i Q_i yields analogous relations p_i = Q_i and P_i = -q_i. In contrast, the Type 2 function F_2(q, P, t) = \sum_i q_i P_i directly produces the true identity transformation Q_i = q_i, P_i = p_i, and K = H, demonstrating its utility for unchanged variables.

Types 3 and 4

The latter two types of generating functions for canonical transformations, types 3 and 4, extend the foundational approach of types 1 and 2 by depending on the old momenta p_i rather than the old coordinates q_i, facilitating transformations that emphasize momentum dependencies. Type 3 generating functions, denoted F_3(p, Q, t), generate transformations where the old coordinates are expressed as q_i = -\frac{\partial F_3}{\partial p_i} and the new momenta as P_i = -\frac{\partial F_3}{\partial Q_i}, with the new given by K = H + \frac{\partial F_3}{\partial t}. This form is particularly suited for transformations mixing old momenta with new coordinates, allowing direct specification of momentum-to-coordinate mappings while preserving the structure. Type 4 generating functions, denoted F_4(p, P, t), provide a for mixed involving both old and new , with relations q_i = -\frac{\partial F_4}{\partial p_i} and Q_i = \frac{\partial F_4}{\partial P_i}, and the new K = H + \frac{\partial F_4}{\partial t}. These functions are advantageous for scenarios where the transformation preserves or directly maps momentum variables, such as in momentum-preserving maps. Both types derive from the total differential of the in the context of the canonical transformation condition, where dF = \sum_i p_i \, dq_i - \sum_i P_i \, dQ_i + \frac{\partial F}{\partial t} dt (or analogous forms adjusted for the variable dependencies to ensure consistency), leading to the rules through coefficient matching. Sign conventions in these relations ensure the preservation of Hamilton's equations, with the negative signs arising from the orientation of the differentials to maintain the form. Notably, the type 4 generating function represents the Legendre transform of the type 1 function, connecting coordinate-based to fully momentum-based descriptions and proving useful for momentum-preserving maps in dynamical systems. In applications, F_4 finds utility in formulating contact transformations within and , where it supports mappings between ray momenta in optical systems or constrained mechanical configurations. A simple example of a Type 3 generating function is F_3(p, Q, t) = \sum_i p_i Q_i, which gives q_i = -Q_i and P_i = -p_i, with K = H if time-independent; this corresponds to a reflection. For Type 4, F_4(p, P, t) = \sum_i p_i P_i yields q_i = -P_i and Q_i = p_i, demonstrating a momentum-coordinate swap that preserves canonicity.

Limitations and extensions

While the four standard types of generating functions—F₁(q, Q, t), F₂(q, P, t), F₃(p, Q, t), and F₄(p, P, t)—provide a powerful for constructing canonical transformations, they exhibit notable limitations in their applicability. Not all canonical transformations fit neatly into one of these types, as some require mixed dependencies on old and new coordinates or momenta that do not align strictly with the prescribed forms, leading to challenges in invertibility or explicit construction. For instance, transformations that are non-invertible, such as those yielding double-valued solutions in systems like the , can cause the standard types to fail or overlap in ambiguous ways, necessitating careful selection or combination of forms to ensure the condition is preserved. To address these constraints, extensions to more general generating functions F(q, p, Q, P, t) have been developed, defined through the relation dF = \sum_i (p_i \, dq_i - P_i \, dQ_i) + higher-order terms constrained by the need to maintain invariance. These generalized forms allow for broader coverage of transformations that depend simultaneously on all variables, though they remain limited by requirements for local invertibility and the preservation of brackets. In time-independent cases, such a general F can often be separated into components akin to F₁ and F₂ using a Legendre transform, facilitating simplification in Hamilton-Jacobi applications. A key relation in these extensions connects the old Hamiltonian H to the new Hamiltonian K via K - H = \frac{\partial F}{\partial t}, which holds for arbitrary generating functions and underscores the role of explicit time dependence in the transformation. Point transformations, where new coordinates Q_i depend solely on old coordinates q_j and time t (with momenta transforming accordingly to preserve the Lagrangian structure), represent a special case of canonical transformations that can be accommodated within these extended frameworks but often revert to simpler generating function types when the mapping is one-to-one.

Extended phase space transformations

Formulation in extended space

In , to accommodate explicitly time-dependent transformations, the is extended to include time t as an additional canonical coordinate with its conjugate taken as -H, where H is the original . This extended is thus spanned by the variables (q, p, t, -H), forming a $2n+2-dimensional manifold for a with n . A canonical transformation in this extended space maps the original variables to new ones (Q, P, T, -K), where K is the transformed , while preserving the extended structure defined by the fundamental one-form \theta = \sum p_i \, dq_i - H \, dt, with the symplectic 2-form \omega = d\theta. The S for such a satisfies the relation dS = \sum p_i \, dq_i - \sum P_i \, dQ_i - H \, dt + K \, dT, with the new time coordinate typically satisfying T = t to maintain the temporal evolution. This formulation ensures that the Poisson brackets remain invariant and Hamilton's equations hold in the extended variables. The extended space approach uniformly incorporates explicit time dependence into generating functions, reducing time-dependent cases to a form analogous to time-independent ones by treating t and -H as a canonical pair. For instance, the type-1 generating function F_1(q, Q, t) yields p_i = \partial F_1 / \partial q_i, P_i = -\partial F_1 / \partial Q_i, and K = H + \partial F_1 / \partial t, allowing seamless handling of H(q, p, t). A distinctive feature of this framework is its connection to action-angle variables, where the action integrals serve as adiabatic invariants under slow variations of system parameters, preserved by the canonical structure in the extended space. Specifically, the action J = \frac{1}{2\pi} \oint p \, dq remains invariant to leading order in the adiabatic approximation. Furthermore, the transformation must correctly map the hypersurface, defined by the constraint H + p_t = 0 (with p_t = -H), to the corresponding surface in the new variables, ensuring the dynamics lie on the zero-energy shell of the extended . This preservation guarantees that the physical trajectories remain consistent with the original constrained motion.

Conditions and relations

In the extended , which incorporates time as an additional coordinate alongside the standard 2n-dimensional variables (q, p), transformations must preserve the structure in 2n+2 dimensions. The extended condition requires that the matrix S of the transformation satisfies S^T J S = J, where J is the (2n+2) × (2n+2) of the form J = \begin{pmatrix} 0 & I_n & 0 & 0 \\ -I_n & 0 & 0 & 0 \\ 0 & 0 & 0 & 1 \\ 0 & 0 & -1 & 0 \end{pmatrix}, with I_n the n × n ; this ensures the transformation preserves the extended symplectic 2-form ω = ∑ dq_i ∧ dp_i + dH ∧ dt. Poisson brackets in the extended are defined over the coordinates (q, p, t, -H), with the fundamental relation {t, -H} = 1 holding to reflect the role of time in the dynamics. For a canonical transformation to new variables (Q, P, T, -K), preservation of the Poisson bracket structure requires {Q_i, P_j} = δ_{ij}, {T, -K} = 1, and all cross terms such as {Q_i, T} = {Q_i, -K} = {P_j, T} = {P_j, -K} = 0, ensuring the brackets among the original variables are mapped invariantly. These conditions guarantee that the differentials satisfy dt = dT, maintaining the identification of time across the transformation, while the energy variables transform consistently through the relation for the new K(Q, P, T) = H(q(Q, P, T), p(Q, P, T)) + ∂F/∂T, where F is the extended incorporating time dependence. This formulation applies directly to , which acts as a generated by the flow in the extended , preserving the structure and thus the underlying dynamics for time-dependent Hamiltonians.

Infinitesimal canonical transformations

Construction and generators

Infinitesimal canonical transformations arise as the to finite transformations, preserving the structure of to leading order. These transformations are parameterized by a small quantity , such that the new coordinates (Q, P) are related to the old coordinates (q, p) by Q_i = q_i + \epsilon \frac{\partial G}{\partial p_i} and P_i = p_i - \epsilon \frac{\partial G}{\partial q_i}, where G(q, p) is a known as the of the . Equivalently, the infinitesimal changes can be expressed using the as \delta q_i = \epsilon \{q_i, G\} and \delta p_i = \epsilon \{p_i, G\}, highlighting the role of the form in defining the . This form is derived by considering a of the second kind, F_2(q, P, t) = q_i P_i + \epsilon G(q, P, t), which to in \epsilon yields the coordinate relations via the standard rules for transformations: Q_i = \frac{\partial F_2}{\partial P_i} = q_i + \epsilon \frac{\partial G}{\partial P_i} and p_i = \frac{\partial F_2}{\partial q_i} = P_i + \epsilon \frac{\partial G}{\partial q_i}. Since P \approx p to leading order, the partial derivatives with respect to P are approximated by those with respect to p, obtained via a Taylor expansion of the finite transformation around \epsilon = 0 and retaining only linear terms. This construction ensures the transformation satisfies the condition \sum_i (dq_i \wedge dp_i - dQ_i \wedge dP_i) = 0 up to O(\epsilon^2). The generator G uniquely determines the associated Hamiltonian vector field X_G, defined by X_G = \frac{\partial G}{\partial p_i} \frac{\partial}{\partial q_i} - \frac{\partial G}{\partial q_i} \frac{\partial}{\partial p_i}, which governs the infinitesimal flow on . In this sense, G functions as the for this flow, with \epsilon playing the role of an "time" parameter, mirroring the structure of under a true . The set of all such generators forms the of the \mathrm{Sp}(2n, \mathbb{R}), where the Lie bracket of two Hamiltonian vector fields X_{G_1} and X_{G_2} is given by -X_{\{G_1, G_2\}}, with the \{G_1, G_2\} providing the algebra via the relation [X_{G_1}, X_{G_2}] = -X_{\{G_1, G_2\}}. This composition rule ensures that the product of two transformations corresponds to the of their generators, reflecting the group structure of canonical transformations.

Active and passive interpretations

In the active interpretation of an infinitesimal canonical transformation, points in are physically displaced along the flow lines generated by the X_G associated with the generator function G, where the components of the flow satisfy \frac{dq_i}{d\epsilon} = \frac{\partial G}{\partial p_i} and \frac{dp_i}{d\epsilon} = -\frac{\partial G}{\partial q_i}. This view emphasizes the dynamical evolution of the system, treating the transformation as a genuine motion in driven by G. In contrast, the passive interpretation regards the transformation as a mere relabeling or change of coordinates, where the physical points remain fixed while the labels (q, p) are mapped to new coordinates (Q, P) without altering the underlying . Here, scalar functions retain their values at the same physical points, but their functional forms adjust via the coordinate mapping. Both interpretations lead to equivalent mathematical descriptions of the transformation's effect on functions in , though the active view highlights dynamical aspects while the passive stresses coordinate invariance. In the active case, the infinitesimal change in a f at a fixed point is given by \delta f = \epsilon \{ f, G \}, where \epsilon is the infinitesimal parameter and \{ \cdot, \cdot \} denotes the Poisson bracket. In the passive case, the corresponding change arises from the chain rule applied to the coordinate transformation, yielding an analogous expression that preserves the Poisson bracket structure. The interpretations diverge in conceptual emphasis, particularly when transformations generate symmetries of the , as the active view portrays these as physical flows preserving the system's , whereas the passive view sees them as coordinate choices invariant under the action. This distinction becomes relevant in linking conserved quantities to groups, where the active perspective underscores the geometric flow on .

Specific examples

Infinitesimal canonical transformations are exemplified by several fundamental operations in , each associated with a specific function G. provides a key instance, where the is the G = H. The infinitesimal map advances the coordinates by a small time interval \delta t, yielding changes \delta q = \frac{\partial H}{\partial p} \delta t and \delta p = -\frac{\partial H}{\partial q} \delta t. This reflects the flow along Hamilton's equations, integrating to finite-time propagation. A translation in momentum space serves as another concrete example, with the generator G = -\mathbf{P}_0 \cdot \mathbf{q}, where \mathbf{P}_0 is a vector. This produces a uniform shift in while leaving unchanged: \delta \mathbf{q} = 0, \delta \mathbf{p} = -\epsilon \frac{\partial G}{\partial \mathbf{q}} = \epsilon \mathbf{P}_0. Such shifts correspond to changes in the reference frame's velocity without altering spatial coordinates at a fixed instant. Rotations in illustrate a further case, generated by the \mathbf{L} = \mathbf{q} \times \mathbf{p}. For the z-component L_z = x p_y - y p_x, the infinitesimal transformation rotates the coordinates around the z-axis by an angle \epsilon, preserving the symplectic structure and yielding \delta x = -\epsilon y, \delta y = \epsilon x, \delta p_x = -\epsilon p_y, \delta p_y = \epsilon p_x (with vanishing changes in z and p_z). This demonstrates how drives infinitesimal rotations. Boost transformations offer a mixed example, combining shifts with adjustments dependent . The is typically time-dependent, G = \mathbf{v} \cdot (t \mathbf{p} - m \mathbf{q}), producing \delta \mathbf{q} = \mathbf{v} t \epsilon and \delta \mathbf{p} = m \mathbf{v} \epsilon, which implements a velocity . Together with spatial translations (generated by total linear \mathbf{P}), these examples—, translations, rotations, and —generate the ten-dimensional group, underlying the symmetries of non-relativistic . If the is under such a transformation, the G is conserved, serving as a Noether invariant associated with the symmetry.

Applications and interpretations

One-parameter subgroups

In , a one-parameter of canonical transformations consists of a family \{T(\varepsilon)\}_{\varepsilon \in \mathbb{R}} of diffeomorphisms on the , where each T(\varepsilon) is , T(0) is the map, and the family satisfies the group axioms T(\varepsilon + \delta) = T(\varepsilon) \circ T(\delta) for all \varepsilon, \delta \in \mathbb{R}, with closure under inversion given by T(-\varepsilon) = T(\varepsilon)^{-1}. These subgroups are generated by an canonical transformation specified by a function G (the generator) on , linking the abstract group structure to the underlying . The finite transformation T(\varepsilon) acts on phase space functions f via the exponential of the adjoint action in the Lie algebra of Hamiltonian vector fields: T(\varepsilon) f = \exp(\varepsilon \mathrm{ad}_G) f, where \mathrm{ad}_G f = \{G, f\} denotes the adjoint operator defined through the Poisson bracket \{ \cdot, \cdot \}. This Poisson bracket, given by \{f, g\} = \sum_i \left( \frac{\partial f}{\partial q_i} \frac{\partial g}{\partial p_i} - \frac{\partial f}{\partial p_i} \frac{\partial g}{\partial q_i} \right) in canonical coordinates, endows the space of smooth functions with a Lie algebra structure, and the exponential map generates the one-parameter subgroup as the flow of the Hamiltonian vector field X_G associated to G. Such subgroups are always abelian, as the Lie bracket vanishes: [G, G] = \{G, G\} = 0 due to the antisymmetry of the Poisson bracket, ensuring commutativity under composition. This property holds regardless of the specific form of G, as long as it defines a valid Hamiltonian vector field. The explicit form follows from the Baker-Campbell-Hausdorff formula applied to the Lie algebra; for a single generator, it reduces to the power series expansion \exp(\varepsilon \mathrm{ad}_G) f = \sum_{k=0}^\infty \frac{\varepsilon^k}{k!} \mathrm{ad}_G^k f = f + \varepsilon \{G, f\} + \frac{\varepsilon^2}{2!} \{G, \{G, f\}\} + \cdots, derived by considering the infinitesimal change \frac{d}{d\varepsilon} (T(\varepsilon) f) = \{G, T(\varepsilon) f\} with initial condition T(0) f = f. For small \varepsilon, this matches the first-order infinitesimal transformation Q_i = q_i + \varepsilon \frac{\partial G}{\partial p_i}, P_i = p_i - \varepsilon \frac{\partial G}{\partial q_i}, extended iteratively via nested Poisson brackets. These subgroups admit dual interpretations: actively, as integral flows of the X_G solving \frac{d}{d\varepsilon} \Psi(\varepsilon, z) = X_G(\Psi(\varepsilon, z)) with \Psi(0, z) = z; passively, as a parameterized family of coordinate redefinitions preserving the symplectic form. of the flow follows from standard existence and results for ODEs on manifolds, ensuring a well-defined global when the vector field is complete. The structure ensures closure under , as T(\varepsilon) \circ T(\delta) = \exp(\varepsilon \mathrm{ad}_G) \circ \exp(\delta \mathrm{ad}_G) = \exp((\varepsilon + \delta) \mathrm{ad}_G) = T(\varepsilon + \delta) by the properties of the in abelian groups, and inversion holds via T(-\varepsilon) = \exp(-\varepsilon \mathrm{ad}_G).

Motion as transformation

In , the natural time evolution of a mechanical system is interpreted as a canonical transformation that maps initial coordinates to their values at later times. The \Phi_t: (q_0, p_0) \mapsto (q(t), p(t)) is defined such that (q(t), p(t)) satisfies Hamilton's equations \dot{q} = \frac{\partial H}{\partial p} and \dot{p} = -\frac{\partial H}{\partial q}, with initial conditions (q(0), p(0)) = (q_0, p_0). This transformation \Phi_t is canonical for each fixed t, as it is generated by the H through successive canonical transformations along the flow. The canonicity of \Phi_t follows from its preservation of the Poisson bracket structure in phase space. Specifically, for smooth functions f and g on phase space, the flow satisfies \{f \circ \Phi_t, g \circ \Phi_t\}(z) = \{f, g\}(\Phi_t(z)) for any initial point z = (q_0, p_0), ensuring the symplectic form is maintained. To verify this preservation, consider the time derivative of the Poisson bracket along the flow for time-independent functions f and g: \frac{d}{dt} \{f, g\} \circ \Phi_t = \{\{H, f\}, g\} \circ \Phi_t + \{f, \{H, g\}\} \circ \Phi_t. This expression vanishes due to the Jacobi identity for Poisson brackets, \{H, \{f, g\}\} + \{f, \{g, H\}\} + \{g, \{H, f\}\} = 0, which rearranges to show the sum is zero when accounting for the antisymmetry \{a, b\} = -\{b, a\}. Thus, the Hamiltonian flow is symplectic, confirming its status as a canonical transformation. When the Hamiltonian H depends explicitly on time, the family \{\Phi_t\} forms a non-autonomous , meaning it satisfies the property \Phi_{t+s} = \Phi_t \circ \Phi_s only for s, t \geq 0 but fails the full group structure under arbitrary composition due to the time-varying . In the extended , where time t is treated as an additional coordinate conjugate to a new (often set to -H), the become autonomous, and the flow realizes a one-parameter of transformations. This perspective aligns with the broader framework of one-parameter subgroups generated by , applied here to dynamical evolution.

Liouville's theorem

Liouville's theorem states that the volume element in , denoted d^{2n} z = dq_1 \cdots dq_n dp_1 \cdots dp_n, remains invariant under canonical transformations, ensuring that for flows the time derivative of any phase space volume satisfies dV/dt = 0. This preservation arises because the flow generated by the equations distorts the shape of phase space regions but does not change their volume. The proof relies on the fact that the divergence of the X_H, defined by \dot{q}_i = \partial H / \partial p_i and \dot{p}_i = -\partial H / \partial q_i, vanishes identically. Specifically, \nabla \cdot X_H = \sum_{i=1}^n \left( \frac{\partial}{\partial q_i} \left( \frac{\partial H}{\partial p_i} \right) + \frac{\partial}{\partial p_i} \left( -\frac{\partial H}{\partial q_i} \right) \right) = 0, since the equality of mixed partial derivatives implies the terms cancel. This zero divergence guarantees that the of the transformation is unity, preserving volumes. The theorem can be expressed using the Liouville operator \mathcal{L} f = \{ H, f \}, the Poisson bracket with the . For any smooth function f on , the time evolution of its integral yields \frac{d}{dt} \int f \, d^{2n} z = \int \{ H, f \} \, d^{2n} z = 0, where the integral form follows from , assuming boundary terms vanish at infinity. This demonstrates the conservation of phase space integrals under the flow. In , Liouville's theorem implies the conservation of probability densities \rho(q, p, t), satisfying the \partial \rho / \partial t + \{ \rho, H \} = 0, so \rho remains constant along trajectories. In , this volume preservation supports the by ensuring that the invariant measure on allows time averages of observables to equal ensemble averages. The theorem generalizes to weighted volumes through the conservation of integrals \int f \rho \, d^{2n} z for arbitrary weighting functions f and densities \rho, provided the flow preserves the underlying measure.

Modern perspectives

Symplectic geometry framework

In the modern framework of , canonical transformations are abstracted from coordinate-dependent descriptions to the intrinsic geometry of . A is a pair (M, \omega), where M is a smooth manifold and \omega is a closed, non-degenerate 2-form known as the form. A canonical transformation is then realized as a f: (M, \omega) \to (M, \omega), which is a preserving the symplectic structure, ensuring that the geometry of remains invariant under such maps. Central to this framework is the X_H associated to a smooth function H: M \to \mathbb{R}, defined by the contraction equation \iota_{X_H} \omega = -dH, where \iota denotes the interior product and dH is the of H. This generates the local of the , with Hamilton's equations emerging as the integral curves of X_H. The of two functions f, g \in C^\infty(M) is given by \{f, g\} = \omega(X_f, X_g), providing a bilinear, antisymmetric operation that encodes the structure of Hamiltonian vector fields and underlies the structure induced by \omega. A f preserves the form \omega if and only if its satisfies f^* \omega = \omega, which is the defining condition for f to be a . This preservation ensures that the volume form \omega^{\wedge n} (for \dim M = 2n) is invariant, linking canonical transformations to the conservation of volume in , though without delving into dynamical applications here. The Darboux theorem guarantees the existence of local around any point in (M, \omega), where \omega = \sum_{i=1}^n dq_i \wedge dp_i, implying that all manifolds of the same dimension are locally indistinguishable up to , with no local invariants beyond the dimension. In , the for a system with configuration manifold Q is canonically the T^*Q, equipped with the tautological form \omega = -d\theta, where \theta is the canonical 1-form; this structure provides a natural setting for canonical transformations as symplectomorphisms of T^*Q.

Connections to and

In , a canonical that leaves the H invariant under its action defines a of the system. Such transformations preserve the form of Hamilton's equations and the structure of , ensuring that the dynamics remain unchanged. Infinitesimal canonical transformations, generated by a function G(q, p), produce variations \delta q_i = \frac{\partial G}{\partial p_i} and \delta p_i = -\frac{\partial G}{\partial q_i}, which correspond to vector fields when the transformation is a . Noether's theorem in the Hamiltonian framework establishes that if G generates a symmetry, meaning the Poisson bracket \{G, H\} = 0, then the time derivative \frac{dG}{dt} = \{G, H\} + \frac{\partial G}{\partial t} = 0 when H is time-independent (\frac{\partial H}{\partial t} = 0), implying G is conserved along trajectories. This connects symmetries directly to conserved quantities without invoking the , as the invariance of H under the canonical transformation ensures the generator G remains constant. For example, spatial translation symmetries yield linear momentum conservation, while rotational symmetries conserve , tying back to the infinitesimal generators of these transformations. The extends to time-dependent s by embedding the system in an extended that includes time t and the negative -H as additional . In this framework, symmetries are transformations on the extended that preserve an like S_0 = \int (p \, dq - H \, dt), leading to conserved quantities of the form G - \tau H, where \tau is a time transformation parameter, even when H explicitly depends on t. This approach maintains the direct link between symmetry generators and laws. In the perspective, is realized through the momentum map J: M \to \mathfrak{g}^*, which associates elements (symmetry generators) in the dual of the \mathfrak{g} of the to conserved quantities on the manifold M. For a action of a G preserving H, the components J_\xi for \xi \in \mathfrak{g} satisfy \{J_\xi, H\} = 0, ensuring conservation, and the map equivariantly links symmetries to the Noether charges. This formulation unifies the conservation laws arising from continuous in systems.

Historical development

Early origins

The concept of canonical transformations in mechanics traces its roots to early 19th-century developments in , where introduced the in 1809 as a tool for analyzing perturbations in planetary motion. This bracket, which quantifies the interdependence of coordinates and momenta, served as a foundational precursor by providing a structure that later transformations would preserve. Poisson's work emphasized the invariance of certain dynamical relations under changes of variables, laying groundwork for more systematic reformulations of Hamiltonian systems. A pivotal advancement came through William Rowan Hamilton's papers published between 1834 and 1837, where he developed the principal function S, an integral over the that generates transformations between old and new coordinates in dynamical systems. Hamilton's S effectively acted as a , enabling the mapping of trajectories while maintaining the form of the , and it drew inspiration from analogies between and . This approach highlighted how such functions could simplify the solution of mechanical problems by transforming variables in a way that preserved essential physical properties. Building on Hamilton's ideas, formalized canonical transformations in 1837 as an integral component of variational principles in . Jacobi's contributions, particularly in his 1842–1843 lectures and publications, integrated these transformations into the framework, emphasizing their role in deriving from a variational standpoint and extending the utility of generating functions as historical tools for coordinate changes.

Key advancements and contributors

The concept of canonical transformations was first introduced by in 1837 during his work on analytical dynamics, where he developed three key theorems, including the theorem on canonical transformations that preserved the form of the . Jacobi's contributions built on the foundations laid by Lagrange and , emphasizing transformations that maintain the canonical structure of systems, as detailed in his 1837 paper "Ueber die Reduction" and posthumously published lectures from 1866. A significant early application came in 1846 when Charles Delaunay employed canonical transformations extensively in for the Earth-Moon-Sun system, marking the first large-scale use of the method in and advancing lunar motion calculations. Building on Jacobi's ideas, Desboves proved Jacobi's theorem on canonical elements in his 1848 doctoral dissertation, utilizing solutions to the as generating functions for such transformations. William Donkin further advanced the theory in 1854–1855 by proving Jacobi's theorems using brackets and introducing time-dependent generating functions, extending the framework to include explicit time variations in the . In the late 19th century, integrated canonical transformations into a broader mathematical vision of dynamical systems, particularly in , where he emphasized their role in perturbation analysis and stability, influencing subsequent research from 1890 to 1910. applied Hamilton-Jacobi methods, including canonical transformations, in his 1868 dissertation on lunar and planetary theory, demonstrating practical utility in orbital computations. By the early 20th century, Carl Ludwig Charlier extended Jacobi's proofs in his 1902 and 1907 volumes on , applying transformations to intermediate orbits and perturbations. Edmund Taylor Whittaker synthesized these developments in his 1904 A Treatise on the Analytical Dynamics, where he formulated and proved Jacobi's transformation theorem using , providing a rigorous geometric interpretation that bridged and emerging modern views. In the 1920s, the theory transitioned to through the work of , who in 1926 developed canonical transformations for , enabling the quantization of classical brackets into commutators and unifying and wave mechanics. Independently, , , and incorporated canonical transformations into in their 1925–1926 papers, while Jordan's 1927 transformation theory further connected classical canonical methods to quantum operator calculus, resolving inconsistencies between representations.

References

  1. [1]
    [PDF] Chapter 4 Canonical Transformations, Hamilton-Jacobi Equations ...
    Thus we see that Hamilton's principal function S is the generator of canonical transforma- tions of constant (Q, P), and provides a method of obtaining ...
  2. [2]
    [PDF] Physics 5153 Classical Mechanics Canonical Transformations
    We define a canonical transformation, as a transformation that takes the Hamiltonian H(q, p, t) to the Hamiltonian K(Q, P, t) such that they both satisfy ...
  3. [3]
    [PDF] PHY411 Lecture notes – Canonical Transformations
    Sep 27, 2023 · Canonical transformations, defined here as those that preserve the Poisson brackets or equivalently the symplectic 2-form, also preserve ...
  4. [4]
    [PDF] Hamiltonian Mechanics
    Sometimes one can find a canonical transformation that results in a simple Hamiltonian function allowing for an easy solution. In particular, if E0 does not ...
  5. [5]
    [PDF] Chapter 6 Hamilton's Equations - Rutgers Physics
    (Q, P) are canonical variables and the transformation (q, p) → (Q, P) is a canonical transformation. Note that the functions Qi and Pi may depend on time as ...
  6. [6]
    15.3: Canonical Transformations in Hamiltonian Mechanics
    Mar 21, 2021 · Canonical transformations are the foundation of Hamiltonian mechanics; they underlie Hamilton-Jacobi theory and action-angle variable theory.Generating functions · Type 1: F = F 1 ⁡ ( q , Q , t ) · Type 2: F = F 2 ⁡ ( q , P , t ) − Q...
  7. [7]
    [PDF] Section 1.3 Canonical Transformations
    1. Canonical Transformations are symplectic. In other words the matrix of their derivatives is a symplectic matrix. • 2. Canonical Transformations preserve ...Missing: formulation | Show results with:formulation
  8. [8]
    [PDF] canonical transformation theory
    This is an alternative form of the symplectic condition. Apply it to the transformation of canonical forces by pre-multiplying by T t. J and using J. -1 = –J.
  9. [9]
    Mechanics: Volume 1 - L D Landau, E.M. Lifshitz - Google Books
    Jan 15, 1976 · 42 Poisson brackets. 135. 43 The action as a function of the coordinates. 138. 44 Maupertuis principle. 140. 45 Canonical transformations.
  10. [10]
    Lagrange Bracket -- from Wolfram MathWorld
    The Lagrange brackets are anticommutative, [u_l,u_m]=-[u_m,u_l] (Plummer 1960, p. 136). If (q_1,...,q_n,p_1,...,p_n)
  11. [11]
    [PDF] CANONICAL TRANSFORMATIONS
    Proposition 2.13: A transformation (q, p) = C(Q, P) is canonical if and only if it preserves the fundamental Lagrange brackets. Proof. Recall that a canonical ...<|control11|><|separator|>
  12. [12]
    [PDF] CANONICAL TRANSFORMATIONS - Poisson
    The symplectic bilinear form is defined by the relations. [ej, ek]=[dj, dk]=0 , [ej, dk] = δj,k , j, k = 1,...,n . We shall refer to the basis {e1,..., en ...
  13. [13]
    [PDF] C:\Downloaded_files\Arnold V I Mathematical Methods Of Classical ...
    In this book we construct the mathematical apparatus of classical mechanics from the very beginning; thus, the reader is not assumed to have any previous ...
  14. [14]
    None
    ### Summary of Generating Functions Type 1 and Type 2 for Canonical Transformations
  15. [15]
    [PDF] 16. Canonical Transformations - DigitalCommons@URI
    Nov 5, 2015 · lated to each other via Legendre transform. The four basic types of generating functions are. F1(q;Q;t) = F2(q;P;t) −. X j. PjQj. = F3(p;Q;t) +.<|control11|><|separator|>
  16. [16]
  17. [17]
    [PDF] Classical Mechanics - Rutgers Physics
    Nov 17, 2010 · ... extended phase space with a vanishing total time derivative along ... Goldstein, at least. While others often use only the last two ...
  18. [18]
    Extended Phase Space - Washington State University
    Despite the complications presented above, one can still employ extended canonical transformations. For example, transformations of type 2 can be implemented ...
  19. [19]
    [PDF] Canonical transformations of the extended phase space, Toda ...
    As an example, we consider canonical transformations of the extended phase space for the Toda lattices and the Stäckel systems.
  20. [20]
    [PDF] The Adiabatic Invariance of the Action Variable in Classical Dynamics
    Oct 11, 2006 · This new quantity arises naturally within the Hamiltonian framework as follows: a canonical transformation is first per- formed to convert the ...<|control11|><|separator|>
  21. [21]
    Structure and Interpretation of Classical Mechanics - GitHub Pages
    A canonical transformation is a phase-space coordinate transformation and an associated transformation of the Hamiltonian such that the dynamics given by ...
  22. [22]
    [PDF] 4. The Hamiltonian Formalism - DAMTP
    Theorem: The Poisson bracket is invariant under canonical transformations. Con- versely, any transformation which preserves the Poisson bracket structure so ...<|control11|><|separator|>
  23. [23]
  24. [24]
    [PDF] An Introduction to Lie Groups and Symplectic Geometry
    Jul 23, 2018 · This course introduces Lie groups and symplectic geometry, focusing on basic concepts and examples, for students familiar with differential ...
  25. [25]
    [PDF] Projective representation of the Galilei group for classical and ... - arXiv
    Feb 23, 2023 · In Hamiltonian mechanics, the space-time transformations associated with the. Galilei group are realized via canonical transformations.
  26. [26]
    Bracket relations for relativity groups - AIP Publishing
    Nov 1, 2016 · Poisson bracket relations for generators of canonical transformations are derived directly from the Galilei and Poincaré groups of space-time ...
  27. [27]
    [PDF] Canonical transformations: from the coordinate based approach to ...
    Sep 27, 2024 · In summary, the generating function of an explicit infinitesimal canonical transformation leaving the Hamiltonian invariant is a constant of ...
  28. [28]
    Finite and Infinitesimal Canonical Transformations - AIP Publishing
    Sep 1, 1970 · The general relation between the infinitesimal generator of a 1‐parameter subgroup of canonical transformations and the usual finite generatingMissing: one- | Show results with:one-
  29. [29]
    [PDF] ANALYTICAL MECHANICS - Math-Unipd
    Jan 10, 2019 · its flow at time −s constitutes a one-parameter group of canonical transformations. < PROOF. The group properties follow from those of the ...
  30. [30]
    [PDF] INTRODUCTION TO SYMPLECTIC MECHANICS: LECTURE IV
    As we noted in previous subsection the flow of a time-dependent Hamiltonian vector field is not a one-parameter group; this fact sometimes leads to technical.
  31. [31]
    [PDF] Lecture 1 - AFS ENEA
    2. Given the concept of the extended phase space, every system can be de- scribed as autonomous system, i.e. with an Hamiltonian that does not explicitly depend ...
  32. [32]
    4 The Hamiltonian Formalism - DAMTP
    For bounded motion, the θ i are usually scaled so that 0 ≤ θ i < 2 ⁢ π and the coordinates ( θ i , I i ) are called angle-action variables. Liouville's Theorem ...
  33. [33]
  34. [34]
    [PDF] Lecture 3: Equilibrium
    This rigorous result from classical mechanics is called Liouville's theorem. It is easy to prove using Hamilton's equations of motion.1 The point of ...
  35. [35]
    symplectomorphism in nLab
    Jun 28, 2023 · Symplectomorphisms are the homomorphisms of symplectic manifolds. In the context of mechanics where symplectic manifolds model phase spaces.
  36. [36]
    [PDF] Hamiltonian Mechanics and Symplectic Geometry
    Physicist's discussions of Hamiltonian mechanics often assume that one can globally choose “canonical coordinates” on phase space and identify it with (R2n,ω0) ...
  37. [37]
    Hamiltonian vector field in nLab
    Apr 4, 2023 · On symplectic manifolds​​ Definition 1.1. For ( X , ω ) a symplectic manifold, a vector field v ∈ Γ ( T X ) is called a Hamiltonian vector field ...
  38. [38]
    [PDF] 1 Symplectic Geometry In Classical Mechanics - Duke Physics
    Any two symplectic manifolds of the same dimension are locally symplectomorphic, so all their distinguishing features must manifest as global properties. These ...
  39. [39]
    Darboux's theorem (symplectic geometry) - PlanetMath.org
    Mar 22, 2013 · Darboux's theorem implies that there are no local invariants in symplectic geometry, unlike in Riemannian geometry, where there is curvature.
  40. [40]
    cotangent bundle in nLab
    Nov 23, 2017 · Given a differentiable manifold X X , the cotangent bundle T * ( X ) T^*(X) of X X is the dual vector bundle over X X dual to the tangent bundle ...
  41. [41]
    [PDF] Applications of Noether conservation theorem to Hamiltonian systems
    The Noether theorem connecting symmetries and conservation laws can be applied directly in a Hamiltonian framework without using any intermediate Lagrangian ...
  42. [42]
    [PDF] Hamiltonian Systems and Noether's Theorem - UChicago Math
    Aug 27, 2015 · In the third and final section, we will state and prove Noether's theorem in terms of momentum maps and the symmetries defined in section two.
  43. [43]
    Classical Mechanics - Oxford Academic - Oxford University Press
    2.2 Poisson brackets. In 1809 Poisson introduced what has come to be known as the Poisson bracket. This analytic construction was to play a profound role ...
  44. [44]
    [PDF] The Early History of Hamilton-Jacobi Dynamics 1834–1837
    May 2, 2023 · Hamilton would subsequently call S the principal function. ... result, to the effect that a generating function for a canonical transformation is ...
  45. [45]
    [PDF] Canonical transformations from Jacobi to Whittaker - Craig Fraser
    Jun 21, 2022 · ... generating function. Historically, the third part of the theory ... [k, h] is then given as. [k, h] = n. ∑ i=1 d(yi , xi ) d(h, k).<|control11|><|separator|>
  46. [46]
  47. [47]
    Moon-Earth-Sun: The oldest three-body problem | Rev. Mod. Phys.
    Apr 1, 1998 · ... canonical transformations eventually won the field. They were used for the first time on a large scale by Delaunay to find the ultimate ...<|separator|>
  48. [48]
    On the theory of quantum mechanics - Journals
    The new mechanics of the atom introduced by Heisenberg may be based on the assumption that the variables that describe a dynamical system do not obey the ...
  49. [49]
    PAM Dirac and the discovery of quantum mechanics - AIP Publishing
    Mar 1, 2011 · In short, Born and Jordan and Dirac independently discovered canonical quantization, and thereby transformed Heisenberg's scheme into a complete ...